159
Views
0
CrossRef citations to date
0
Altmetric
Research Articles

Drinfeld twists of Koszul algebras

ORCID Icon
Pages 3406-3418 | Received 29 Jun 2023, Accepted 06 Feb 2024, Published online: 22 Feb 2024

Abstract

Given a Hopf algebra H and a counital 2-cocycle μ on H, Drinfeld introduced a notion of twist which deforms an H-module algebra A into a new algebra Aμ. We show that when A is a quadratic algebra, and H acts on A by degree-preserving endomorphisms, then the twist Aμ is also quadratic. Furthermore, if A is a Koszul algebra, then Aμ is a Koszul algebra. As an application, we prove that the twist of the q-quantum plane by the quasitriangular structure of the quantum enveloping algebra Uq(sl2) is a quadratic algebra equal to the q1-quantum plane.

2020 MATHEMATICS SUBJECT CLASSIFICATION:

1 Introduction

In [Citation6], Drinfeld introduced a notion of twist for Hopf algebras and their module algebras. Drinfeld twists have been studied in several different contexts within both mathematics and physics. On the physics side they have, for instance, been used by Vafa and Witten [Citation12] (in the form of cocycle twists; a special case of Drinfeld twists) in work on mirror symmetry, and also in Brouder, Fauser, Frabetti and Oeckl [Citation3] where the operator product and time-ordered products of quantum field theory are shown to be Drinfeld twists of the normal product. Within mathematics, they have also found applications to the study of color Lie algebras [Citation4], non-commutative geometry [Citation1, Citation9] and rational Cherednik algebras and their representation theory [Citation2].

Davies [Citation5] (again in the context of cocycle twists) and Montegomery [Citation11] have found various algebra properties that are preserved by Drinfeld twists, for instance Artin-Schelter regularity. The results of this paper continue this line of work, showing that the properties of being quadratic and Koszul are also preserved under general Drinfeld twists. This offers a new means of showing that an algebra of interest is quadratic (or Koszul), by showing that it is related by twist to another algebra that is known to be quadratic (or Koszul).

In the rest of this section we will briefly recall some basic terminology following Majid [Citation10, Section 2.3]. See also Etingof and Gelaki [Citation7, Section 5.14] for more.

Let k be an arbitrary field, and consider all objects to be linear over k, with =k. If H is a Hopf algebra over k, then let :HHH denote the coproduct and ϵ:Hk the counit. For an H-module A we denote the action of hH on aA as ha. An H-module algebra A is an algebra for which the product map m:AAA is an H-module homomorphism, so hm(ab)=m((h)ab), and the unit 1A satisfies h1A=ϵ(h)1A.

Definition 1.1

([Citation10, Example 2.3.1]). A 2-cocycle of the Hopf algebra H is an invertible μHH satisfying (1.1) (μ1)·(id)(μ)=(1μ)·(id)(μ)(1.1)

A 2-cocycle is said to be counital if (1.2) (ϵid)(μ)=1=(idϵ)(μ)(1.2)

Counital 2-cocycles are also referred to as bialgebra twists in Etingof and Gelaki [Citation7, Definition 5.14.1], and twisting elements in Giaquinto and Zhang [Citation9, Definition 1.2].

The Drinfeld twist of H by μ is defined to be the Hopf algebra Hμ which shares the same algebra structure and counit as H, and has coproduct μ:HμHμHμ,hμ·(h)·μ1. We shall refer to Hμ as a “Hopf algebra twist” in order to differentiate it from the following, which is also called a Drinfeld twist.

Definition 1.2.

The Drinfeld twist of an H-module algebra (A, m) by μ is the Hμ-module algebra (Aμ,mμ) where Aμ=A as k-vector spaces, with the product map mμ:AμAμAμ given by abm(μ1ab) for all a,bAμ.

In Section 2 we consider A to be a quadratic H-module algebra, with the action of H on A being degree-preserving. We show that in this case, the Drinfeld twist of A by a counital 2-cocycle μ is also a quadratic algebra, and we are able to determine the (quadratic) relations of Aμ explicity by means of μ acting on the relations of A. Using this, we prove in Section 3 that if A is a Koszul algebra, then the Drinfeld twist Aμ must be Koszul. This extends to arbitrary Hopf algebras a result of Davies [Citation5, Proposition 4.25], who proved this for the case when H is the group algebra of a finite abelian group. Finally in Section 4 we give several examples which apply these results. In particular we show that the q-quantum plane may be twisted by the quasitriangular structure of Uq(sl2), and the result is the q1-quantum plane. Additionally, we give new proofs that the quantum symmetric and exterior algebras S1(V) and 1(V) are Koszul.

2 Quadratic algebras

Let A be a connected N-graded k-algebra, meaning A=AiNi with Ai·AjAi+j and A0=k. We will assume throughout this paper that A is locally finite-dimensional, meaning that each grading component Ai is finite-dimensional. Now A is also a quadratic algebra if it is generated by its degree 1 elements V:=A1, and has quadratic relations RVV, so that A=T(V)/(R), where T(V) is the tensor algebra over V and (R) is the 2-sided ideal generated by R inside T(V).

Suppose A=T(V)/(R) is both a quadratic algebra, and an H-module algebra, for some Hopf algebra H acting by degree-preserving endomorphisms (i.e. haAi for all hH and aAi). Then, it is easy to see that the degree 1 subspace V:=A1 is an H-submodule of A. Additionally, this action of H on V extends naturally by means of the coproduct of H to make T(V) an H-module algebra.

Now a fact that will be very useful throughout the paper is that the space of relations R is an H-submodule of T(V). This can be seen as follows. Note that a quotient of an H-module algebra by an ideal can only be an H-module algebra too if the ideal is also an H-submodule. Since the H-module algebra structure of A arises from quotienting the H-module algebra T(V) by the ideal (R), we see that (R) must also be an H-submodule of T(V). Additionally, since R is precisely the subspace of degree 2 elements in (R), it must be fixed under the degree-preserving endomorphisms of H. And so R is an H-submodule of T(V).

We now give our first main result.

Theorem 2.1.

Let A=T(V)/(R) be a quadratic H-module algebra, where H is a Hopf algebra acting by degree-preserving endomorphisms. If μ is a counital 2-cocycle of H, then the Drinfeld twist Aμ is a quadratic algebra of the form T(V)/(Rμ) where Rμ:={μr|rR}.

Proof.

Firstly let us show Aμ is a connected N-graded algebra under the same grading as on A. Since Aμ=A as vector spaces, Aμ forms an N-graded vector space under the grading of A. Let m and mμ denote the products on A and Aμ respectively. Now Aμ is a graded algebra if mμ(vw)Ai+j for all vAi and wAj. But this follows since mμ:=m(μ1), and the action of H is degree-preserving, so μ1vwAiAj. Aμ is also connected as (Aμ)0=A0=k.

Note that, as the action of H is degree-preserving, Rμ is a subspace of VV. Also, since A=Aμ as N-graded vector spaces, V:=A1 may be viewed as the subspace of degree 1 elements of Aμ. Therefore, by showing Aμ=T(V)/(Rμ), we will have proven that Aμ is a quadratic algebra, as required.

First we prove Aμ is generated by V by induction on the degree of elements of Aμ. Consider a degree 2 element xA2. Since A is generated by V, x=im(viwi) for some vi,wiV. Therefore, x=im(viwi)=im(μ1(μviwi))=imμ(μviwi)

Since the action of H is degree preserving, μviwiVV, and so x can be expressed as a linear combination of mμ-products of elements of V, as required.

Now let k2, and suppose every element of degree k in Aμ can be expressed as a linear combination of mμ-products of elements of V. We show that this implies the elements of degree k + 1 have a similar such decomposition. Consider xAk+1. Since A is generated by V, we may decompose x as a linear combination of terms which are the m-products of k + 1-elements in V, i.e. is of the form m(v1m(v2m())), for some v1,,vk+1V. Suppose μ=μ1μ2 for some μ1,μ2H. Then (2.1) m(v1m(v2m()))=mμ([μ1v1][μ2m(v2m())])(2.1)

Now μ2m(v2m())Ak, and so by the induction hypothesis this can be decomposed into a linear combination of mμ-products of k-elements of V. Therefore m(v1m(v2m())) can be decomposed using mμ-products of k + 1-elements, and this implies the same holds for x too. This concludes the induction argument proving Aμ is generated by V.

There now exists a natural surjective algebra homomorphism ϕ:T(V)Aμ. We show ker(ϕ)=(Rμ) to complete the proof that Aμ is a quadratic algebra. Note that for the natural embedding i:VA, we have R=ker(m°(ii)). As i is an H-module homomorphism, ii is an HH-module homomorphism, and we see mμ°(ii)=m°(ii)(μ1). So rker(mμ°(ii)) if and only if μ1rR, and this holds if and only if rRμ. So (Rμ)ker(ϕ), and therefore Aμ=T(V)/ker(ϕ) is seen to be a quotient of T(V)/(Rμ).

To show that Aμ is equal to T(V)/(Rμ), we will apply a dimensional argument. Note that all dimensions in the following are finite due to our assumption that A is locally finite-dimensional. Now, since Aμ is a quotient of T(V)/(Rμ), the dimension of each grading component of Aμ must be less than or equal to the dimension of the same degree component of T(V)/(Rμ), i.e. (2.2) dim(Aμ)idim(T(V)/(Rμ))i(2.2)

Next we will consider the Drinfeld twist of T(V)/(Rμ) by the cocycle μ1, which leads us to another dimensional inequality (see (2.3)). First we must check that we can indeed take such a twist. Note that this twist is over the Hopf algebra Hμ, rather than the Hopf algebra H that we have used so far. It is a simple exercise to check μ1 is a counital 2-cocycle of the Hopf algebra twist Hμ, and we show next that T(V)/(Rμ) is an Hμ-module algebra.

Since Hμ=H as algebras, the H-action on V defines an Hμ-action on V. This extends, by means of the twisted coproduct μ, to make T(V) an Hμ-module algebra. Now Rμ is an Hμ-submodule of T(V) since, if hHμ and rRμ, where r=μr for some rR, we have hr=(μ·(h)·μ1)(μr)Rμwhere we use the fact that (h)rR, since we proved above that R is an H-submodule of the H-module algebra T(V). So we have established T(V) is an Hμ-module algebra, and Rμ is an Hμ-submodule, and therefore T(V)/(Rμ) is an Hμ-module algebra as we required.

We can now consider the Drinfeld twist of T(V)/(Rμ) by μ1. In direct analogy to the argument used for Aμ, one may show that V generates (T(V)/(Rμ))μ1, i.e. use the fact T(V)/(Rμ) is generated by V to express elements of (T(V)/(Rμ))μ1 as linear combinations of T(V)/(Rμ)-products of elements of V. Then one rewrites each T(V)/(Rμ)-product as a linear combination of (T(V)/(Rμ))μ1-products.

We therefore have a surjective map ϕ:T(V)(T(V)/(Rμ))μ1. It is easy to show that (R)ker(ϕ), and so we find (T(V)/(Rμ))μ1 is a quotient of A. This implies the following inequality in the dimensions of the grading components of degree i, (2.3) dim((T(V)/(Rμ)μ1)idim(A)i(2.3)

But as discussed at the start of the proof, twisting preserves the grading on algebras, so (2.4) dim(A)i=dim(Aμ)i,dim(T(V)/(Rμ))i=dim((T(V)/(Rμ)μ1)i(2.4)

Combining inequalities (2.2), (2.3), and (2.4) we find dim(Aμ)i=dim(T(V)/(Rμ))i. But since Aμ was a quotient of T(V)/(Rμ), we deduce the algebras must be equal. □

3 Koszul algebras

3.1 Statement of the main theorem

Let A be a connected, N-graded, and locally finite-dimensional, k-algebra. Define A+:=Ai>0i, so A/A+k, and we call the corresponding quotient map ϵ:Ak the augmentation map. Now k is an A-module via ϵ, and A is called a Koszul algebra if there is a linear graded free resolution of k as an A-module (see Witherspoon [Citation14, Definition 3.4.3]). If A is Koszul, then it is a standard fact that it is also quadratic, so A=T(V)/(R) for V = A1 and R a subspace of VV.

The next result establishes that Koszulity is preserved under Drinfeld twists. This generalizes to arbitrary Hopf algebras a result of Davies [Citation5, Proposition 4.25] who proved this for the case when the Hopf algebra H is the group algebra of a finite abelian group. The rest of the section is dedicated to proving this result.

Theorem 3.1.

Let A=T(V)/(R) be a Koszul H-module algebra, where H is a Hopf algebra acting by degree-preserving endomorphisms. If μ is a counital 2-cocycle of H, then the Drinfeld twist Aμ is a Koszul algebra given by T(V)/(Rμ) where Rμ:={μr|rR}.

3.2 Plan for the proof

It follows immediately from Theorem 2.1 that Aμ is a quadratic algebra of the form T(V)/(Rμ). Using this we can construct a complex K(Aμ), which, by Witherspoon [Citation14, Theorem 3.4.6], is a resolution of k as an Aμ-module precisely when Aμ is a Koszul algebra. We therefore seek to show K(Aμ) is a resolution to complete the proof. We start by considering the Koszul resolution K(A) of A, and twist this resolution using a functor of Giaquinto and Zhang. This produces a new resolution of k as an Aμ-module, which we denote K(A)μ. We then construct an isomorphism of complexes between K(A)μ and K(Aμ), the existence of which implies K(Aμ) is a resolution k as an Aμ-module, as required.

3.3 The Koszul resolution of A

Let us start by defining the Koszul resolution of k as an A-module. It is given by K(A)=Kn0n(A), where (3.1) K0(A)=A,K1(A)=AV,Kn(A)=Ai+j=n2(ViRVj)for n2.(3.1)

The differentials dn are induced by the canonical embedding of K(A) into the Bar resolution B(A) of k, where Bn(A)=A(n+1) and (3.2) dn(a0an)=(1)nϵ(an)a0an1+i=0n1(1)ia0aiai+1an(3.2) for a0,,anA and ϵ is the augmentation map of A. Kn(A) is a left A-module under multiplication on the leftmost tensor leg. It is also a left H-module by restricting the natural action of H on Bn(A)=A(n+1) (using the coproduct of H) onto Kn(A). It is clear for the n = 0 and n = 1 cases that Kn(A) is closed under this H-action. For n2, we showed above that V and R are H-submodules of T(V), and so each can be seen as an H-submodule of A and AA respectively. Therefore, in this case, Kn(A) is an intersection of H-submodules of Bn(A), so is an H-submodule itself.

3.4 The Giaquinto and Zhang twisting functor

Let A‐Mod be the category of all left A-modules. Take the category (H,A)‐Mod to be the subcategory of A‐Mod whose objects M are also left H-modules such that the following holds for all mM,hH, aA, (3.3) h(am)=(h(1)a)(h(2)m)(3.3) where (h)=h(1)h(2). The morphisms are those of A‐Mod that are also H-module homomorphisms.

So far, Drinfeld twists have provided a mechanism for deforming the Hopf algebra H and the H-module algebra A. Giaquinto and Zhang [Citation9, Theorem 1.7] extends this by defining a way to twist a module in (H,A)‐Mod into a module in (Hμ,Aμ)‐Mod. This new twist defines an equivalence of categories (H,A)‐Mod(Hμ,Aμ)‐Mod. We show next that the Koszul resolution K(A) can be defined within (H,A)‐Mod, and describe its image under this twisting functor of Giaquinto and Zhang.

Let us first check that (3.3) holds on Kn(A) for all n0. For K0(A)=A, let hH,a,bA, then h(ab)=h(a·b)=(h(1)a)·(h(2)b)=(h(1)a)(h(2)b)as required. For n1, let mKn(A)A(n+1). It is sufficient to show (3.3) holds when m=am for some aA,mAn, since it will then follow for a general element of Kn(A) by linearity. So, h(am)=h(a·am)=(h(1)a·a)(h(2)m) =(h(1)(1)a)·(h(1)(2)a)(h(2)m) =(h(1)a)·(h(2)(1)a)(h(2)(2)m) =(h(1)a)(h(2)m) where we use coassociativity of H in the third equality. So (3.3) is satisfied and Kn(A) is an object in (H,A)‐Mod. It is standard that the differentials dn of the Koszul resolution are A-module homomorphisms, but we note that they are also H-module homomorphisms. Indeed, on inspecting (3.2), we see that as a map, (3.4) dn=i=0n1(1)iidimidni1+(1)nidnϵ(3.4) where m denotes the product map of A and ϵ is the augmentation map of A. Since m and ϵ are H-module homomorphisms, it follows that dn also is. Therefore the whole Koszul resolution K(A) can be defined within the category (H,A)‐Mod.

We now apply the functor of Giaquinto and Zhang [Citation9, Theorem 1.7]. Under this functor the A-modules Kn(A) are twisted into Aμ-modules Kn(A)μ in the following way: let Kn(A)μ=Kn(A) as k-vector spaces, and equip Kn(A)μ with the action μ:AμKn(A)μKn(A)μ,μ=(μ1)

Here we use the fact that A and Kn(A) are H-modules, so AKn(A) is naturally an HH-module. Therefore μ1 defines a k-linear endomorphism of AKn(A), and also of AμKn(A)μ, since these are equal as k-vector spaces.

We must also apply the twisting functor to the differentials in the Koszul resolution. Giaquinto and Zhang do not explicitly describe how their functor behaves on morphisms, so for completeness, we prove here that we can take it as mapping each morphism in (H,A)‐Mod to itself. Suppose V,W(H,A)‐Mod with the actions of A on V and W denoted by V and W respectively. Let ϕ:VW be an A-module homomorphism, so W°(idAϕ)=ϕ°V. Then ϕ is also an Aμ-module homomorphism VμWμ. Indeed μW°(idAϕ)=W°(idAϕ)(μ1)=(ϕ°V)(μ1)=ϕ°μVwhere we use the fact idAϕ is an H-module homomorphism in the first equality. ϕ is also an Hμ-module homomorphism VμWμ, since Hμ acts on Vμ and Wμ in exactly the same way that H acts on V and W respectively. Therefore ϕ can also be viewed as a morphism in (Hμ,Aμ)‐Mod from Vμ to Wμ, and so it makes sense to let the functor of Giaquinto and Zhang act identically on morphisms. Therefore each differential dn of the Koszul resolution K(A) will be sent by the functor of Giaquinto and Zhang to itself.

To summarize, the result of applying the twisting functor to K(A) is a complex K(A)μ of Aμ-modules which share the same underlying vector spaces, and differentials, as those on K(A). Since K(A) is a resolution of k as an A-module, we see that the new “twisted” complex K(A)μ must be a resolution of k as an Aμ-module.

3.5 The Koszul complex of Aμ

Using Theorem 2.1 we know that Aμ is a quadratic algebra of the form T(V)/(Rμ). We can therefore construct a complex K(Aμ) completely analogously to the Koszul resolution for A given in (3.1), and we describe this explicitly next. Let K(Aμ)=Kn0n(Aμ) where K0(Aμ)=Aμ,K1(Aμ)=AμV, Kn(Aμ)=Aμi+j=n2(ViRμVj)for n2.

This is a complex of Aμ-modules, where the differentials are inherited from the Bar resolution B(Aμ), where Bn(Aμ)=(Aμ)(n+1) and (3.5) dnμ=i=0n1(1)iidimμidni1+(1)nidnϵ(3.5) where we use the same augmentation map ϵ as on A, since Aμ and A have the same grading.

Note that by Witherspoon [Citation14, Theorem 3.4.6], K(Aμ) is a resolution of k as an Aμ-module precisely when Aμ is a Koszul algebra. We use this to prove our result, showing that K(Aμ) is indeed a resolution. To do so we construct an isomorphism of complexes from K(A)μ to K(Aμ), and use the fact established in the last section that K(A)μ is a resolution of k as an Aμ-module. To construct this isomorphism of complexes we first define a sequence of elements which turn out to generalize counital 2-cocycles.

3.6 Higher counital 2-cocycles

First we introduce some notation: for i{1,,n}, let i(n):HnHn+1,i(n)=idid, where appears in the i-th position. By convention let 0(n)()=1() and n+1(n)()=()1.

Recall that a counital 2-cocycle μ must satisfy the 2-cocycle Equationequation (1.1). This equation may be rewritten in the above notation as: (3.6) 3(2)(μ)·1(2)(μ)=0(2)(μ)·2(2)(μ)(3.6)

It turns out that μ can be viewed as just one step in a sequence of “higher counital 2-cocycles”, where the next two lemmas justify this terminology. For n0, let fnH(n+1) be defined as f0=1,f1=μ and for n2, (3.7) fn=(μ1n1)·i=0n2(1(n)°1(n1)°°1(ni))(μ1ni2)(3.7) where ° denotes composition, and all products are taken in the algebra H(n+1). Notice that f2=(μ1)·(id)(μ), which is just the left hand side of the 2-cocycle Equationequation (3.6). The following lemma can be viewed as saying that each of the elements fn satisfies a “higher” version of the 2-cocycle Equationequation (3.6):

Lemma 3.2.

For all nN and i{1,,n},fn=(1i1μ1ni)·i(n)(fn1).

Proof.

Let fni:=(1i1μ1ni)·i(n)(fn1). Firstly it is clear that f1=f11. Now let us check fn=fn1 for n2. By definition, fn1=(μ1n1)·1(n)(fn1), and using (3.7), 1(n)(fn1)=1(n)(μ1n2)·1(n)(i=0n3(1(n1)°°1(ni1))(μ1ni3)) =i=0n2(1(n)°1(n1)°°1(i+2))(μ1i)

Therefore fn1=fn as required. To finish proving the lemma we show that, for each n2,fni=fni+1 for all i{1,,n1}. We do so by performing induction on n. The base case n = 2 holds since f21=f22 is precisely the 2-cocycle Equationequation (3.6). Now suppose the hypothesis holds for n=k1, and let us show it holds for n = k. Take some i{1,,n1}. We know that fk1=fk11, and now by the induction hypothesis we have (3.8) fk1=fk1i=(1i1μ1ki1)·i(k1)(fk2)(3.8)

Therefore, fki+1=(1iμ1ki1)·i+1(k)((1i1μ1ki1)·i(k1)(fk2)) =(1iμ1ki1)·i+1(k)(1i1μ1ki1)·i+1(k)(i(k1)(fk2)) =(1i1[0(2)(μ)·2(2)(μ)]1ki1)·i+1(k)(i(k1)(fk2)) =(1i1[3(2)(μ)·1(2)(μ)]1ki1)·i+1(k)(i(k1)(fk2)) =(1i1μ1ki)·i(k)(1i1μ1ki1)·i+1(k)(i(k1)(fk2))where in the first equality we use the definition of fki+1, and insert the expression (3.8) for fk1. Next we use the fact that i+1(k) is an algebra homomorphism. In the 3rd equality we take the product of the first two terms in the previous line, and in the 4-th equality we apply (3.6). Finally, by coassociativity of H, i+1(k)°i(k1)=i(k)°i(k1), and therefore the final expression above reduces to fki, as required. □

Recall that μ also satisfies the counital Equationequation (1.2), i.e. (ϵid)(μ)=1=(idϵ)(μ), where ϵ is the counit of H. The following lemma proves that the elements fn satisfy a generalized notion of counitality,

Lemma 3.3.

For all n1 and i{0,,n},(idiϵidni)(fn)=fn1.

Proof.

The n = 1 case is precisely (1.2). For arbitrary n1 and i{0,,n}, we can apply Lemma 3.2 to express fn as (1i1μ1ni)·i(n)(fn1). Then (idiϵidni)(fn)= (idiϵidni)(1i1μ1ni)·(idiϵidni)(i(n)(fn1))

Now (idiϵidni)(1i1μ1ni)=1i1(idϵ)(μ)1ni=1nby the counitality of μ. Also, (idiϵidni)(i(n)(fn1))=(idi1[(idϵ)°]idni)(fn1) and this is equal to fn1 since by the counit axiom for H, (idϵ)°=id. □

3.7 Defining the isomorphism of complexes F

Using the higher counital 2-cocycles constructed in the last section we will now define an isomorphism of complexes between K(A)μ and K(Aμ). From this we will deduce that K(Aμ) is a resolution, and therefore that Aμ is a Koszul algebra.

Recall that the complex K(A) embeds into the Bar resolution B(A), so in particular Kn(A) is a subspace of An+1, for all n0. Additionally, Kn(A)μ=Kn(A) as k-vector spaces, and therefore Kn(A)μ is also a subspace of An+1. Since A is an H-module algebra, An+1 is naturally an Hn+1-module. Therefore each fnHn+1 given in (3.7) has a well-defined action on the space Kn(A)μ.

Define F:K(A)μK(Aμ) to be the corresponding sequence of k-linear maps, i.e. for each n0, Fn is the map given by fn acting on Kn(A)μ. The first few maps of F are depicted in the diagram below,

We check in Section 3.8 that the image of Fn is indeed inside Kn(Aμ), and in Section 3.9 that the diagram above commutes, i.e. dnμFn=Fn1dn for all n1. Finally in Section 3.10 we show each Fn has a k-linear inverse. These facts are sufficient to deduce that K(Aμ) is a resolution of k as an Aμ-module, which is discussed in Section 3.11.

Remark 3.4.

It is possible to show that the k-linear maps Fn are also Aμ-module homomorphisms, and therefore that F is an isomorphism of complexes. However it is not required that Fn be an Aμ-module homomorphism in order to prove K(Aμ) is a resolution, so we won’t include proof of this fact here. For brevity though, we will still refer to F as being a chain map, or, isomorphism of complexes.

3.8 Checking im(Fn)Kn(Aμ)

For the n = 0 case, F0=id, and as vector spaces, K0(A)μ=A=Aμ=K0(Aμ), so the image of F0 is equal to K0(Aμ), as required.

When n = 1, we have as vector spaces K1(A)μ=AV=AμV=K1(Aμ). Now V was defined to be the degree 1 component A1 of A, and since the H-action on A is degree-preserving, V is an H-submodule of A. Therefore AV is an HH-submodule of AA, so the action of μ on K1(A)μ will be closed. Due to the equality of vector spaces K1(A)μ=K1(Aμ), we can view the image of μ as being in K1(Aμ).

For n2, we inspect how fn in (3.7) acts on an element aKn(A)μ, and show the result lies in Kn(Aμ). Recall that, Kn(A)μ=Ai+j=n2(ViRVj)Kn(Aμ)=Aμi+j=n2(ViRμVj)

If aKn(A)μ, then aAViRVn2i for all i{0,,n2}. Similarly, we can show fnaKn(Aμ) by showing fnaAViRμVn2i for all i{0,,n2}, where we have used here the fact Aμ=A as vector spaces. Fix an arbitrary i{0,,n2}. By Lemma 3.2, fn=(1i+1μ1ni2)·i+2(n)(fn1), and so, fna=(1i+1μ1ni2)(i+2(n)(fn1)a)

Let us inspect where we land after i+2(n)(fn1) acts on a. First note aAViRVn2i, and since A and V are H-modules, the legs of i+2(n)(fn1) hitting A or V will remain in these spaces. Recall also that R is an H-submodule of T(V), so (h)rR for all hH and rR. Since the in i+2(n)=idid hits the R in AViRVn2i, we may apply the fact that R is an H-submodule of T(V), to deduce i+2(n)(fn1)aAViRVn2i

Now when 1i+1μ1ni2 acts on an element of AViRVn2i we clearly land in AViRμVn2i. Since i was arbitrary in {0,,n2}, we find fnaKn(Aμ), and therefore the image of Fn is indeed inside Kn(Aμ) as required.

3.9 F commutes with differentials

Let us now check that the maps Fn make the above diagram commute, i.e. dnμFn=Fn1dn for all n1. We inspect the result of each side of this condition on an element a0anKn(A)μ: Fn1dn(a0an) =fn1(i=0n1(1)iidimidni1+(1)nidnϵ)(a0an) =i=0n1(1)i(idimidni1)(i+1(n)(fn1)a0an) +(1)nfn1(idnϵ)(a0an)where in the first equality we apply the definition of the map dn given in (3.4). In the second equality we use the fact that (idimidni1) is an Hn-module homomorphism to pull the fn1 inside, resulting in the i+1(n)(fn1) term. Now, using the definition of dnμ in (3.5), dnμFn(a0an)=i=0n1(1)i(idimμidni1)(fna0an) +(1)n(idnϵ)(fna0an) and notice that (idimμidni1)=(idimidni1)((1iμ11ni1))

By Lemma 3.2, fn=(1iμ1ni1)·i+1(n)(fn1), and therefore (1iμ11ni1)·fn=i+1(n)(fn1). This nearly concludes the proof that dnμFn(a0an)=Fn1dn(a0an). It only remains to show (idnϵ)(fna0an)=fn1(idnϵ)(a0an)

As A is an H-module algebra, we have h1=ϵ(h)1 where ϵ is the counit of H. Therefore the degree 0 component A0k of A has the structure of a trivial H-module, i.e. hλ=ϵ(h)λ for all hH and λA0. Since the augmentation map ϵ maps into k, we find hϵ(x)=ϵ(h)ϵ(x). Using this, and the fact ϵ is an H-module homomorphism, we find (idnϵ)(fna0an)=fn(idnϵ)(a0an) =(idnϵ)(fn)(idnϵ)(a0an) =fn1(idnϵ)(a0an)where we apply the “higher counitality” of fn (Lemma 3.3) in the final equality to say fn1=(idnϵ)(fn).

3.10 The inverse chain map

Finally it is clear from the fact that μ is invertible that each map Fn has a k-linear inverse. In particular, let Fn1:=fn1, where f01=1,f11=μ1 and for n2, (3.9) fn1=(i=0n2(1(n)°1(n1)°°1(i+2))(μ11i))·(μ11n1)(3.9)

Although it is not required in this proof, it can be checked that F1 are Aμ-module homomorphisms, and it is clear that they will also commute with differentials. Therefore F is an isomorphism of complexes, as claimed.

3.11 Kn(Aμ) is a resolution of k as an Aμ-module

Recall that the complex K(A)μ, which arose from applying the functor of Giaquinto and Zhang to the Koszul resolution Kn(A), is a resolution of k as an Aμ-module. This means that the following complex is exact,

(3.10)

We wish to show that K(Aμ) is also a resolution of k as an Aμ-module, as this is equivalent to Aμ being a Koszul algebra. Hence we must show the following is exact,

In particular we wish to show im(dn+1μ)=ker(dnμ) for n1 and im(d1μ)=ker(ϵ).

We showed in Section 3.9 that F commutes with differentials, so for all n0,dn+1μ=Fndn+1Fn+11. Additionally every map Fn is a k-linear isomorphism, and so we find im(dn+1μ)=Fn(im(dn+1)). Also, for n1,dnμ=Fn1dnFn1, and we see that ker(dnμ)=Fn(ker(dn)). Therefore, for n1, im(dn+1μ)=Fn(im(dn+1))=Fn(ker(dn))=ker(dnμ)where in the second equality we apply the fact (3.10) is exact so im(dn+1)=ker(dn) for n1. Finally im(d1μ)=F0(im(d1)), but F0=id, so im(d1μ)=im(d1)=ker(ϵ), again using the fact (3.10) is exact. Therefore K(Aμ) is a resolution of k as an Aμ-module, and Aμ is a Koszul algebra.

4 Examples

Next we demonstrate Theorems 2.1 and 3.1 with several examples. In particular, in Section 4.1, we use Theorem 2.1 to determine the result of twisting the quantum plane by the quasitriangular structure of the quantum enveloping algebra Uq(sl2). Then in Section 4.2 we apply Theorem 3.1 to provide a new proof of the Koszulity of the quantum symmetric and exterior algebras S1(V) and (V)1.

4.1 The quantum plane and Uq(sl2)

For a non-zero qk, the q-quantum plane is given by A=kx,y/(xyqyx). In the following we explain why we can twist A by the quasitriangular structure R of the quantum enveloping algebra Uq(sl2), and we find via Theorem 2.1 that AR is the quadratic algebra equal to the q1-quantum plane kx,y/(xyq1yx).

First recall that a quasitriangular structure on a Hopf algebra H is an invertible RHH satisfying: (4.1) τ°(h)=R·(h)·R1,hH,(id)(R)=R13·R23,(id)(R)=R13·R12(4.1) where τ is the transposition map vwwv,R12=R1,R23=1R and R13 has 1 inserted in the middle leg. Quasitriangular structures are known to satisfy the quantum Yang-Baxter equation [10, Lemma 2.1.4]: R12·R13·R23=R23·R13·R12, and from this, one can check that a quasitriangular structure is also a counital 2-cocycle, i.e. satisfies (1.1) and (1.2). For this reason it makes sense to use quasitriangular structures to perform Drinfeld twists.

Let us introduce the quantum enveloping algebra Uq(sl2). We now suppose q also satisfies q21. Then Uq(sl2) is defined to be the k-algebra generated by E,F,K,K1, and satisfying the relations KE=q2EK,KF=q2KF,EFFE=KK1qq1,KK1=K1K=1.

Uq(sl2) may additionally be given the structure of a Hopf algebra via the following: (E)=1E+EK,ϵ(E)=0,S(E)=EK1 (F)=K1F+F1,ϵ(F)=0,S(F)=KF (K±1)=K±1K±1,ϵ(K±1)=1,S(K±1)=K1

Now the quantum plane A=kx,y/(xyqyx) is, by construction, a quadratic algebra. Additionally, we may define a representation of Uq(sl2) on the degree 1 homogeneous subspace V=spank{x,y} of A as follows: Ex=0,Fx=y,K±1x=q±1x, Ey=x,Fy=0,K±1y=q1y

It is well-known (see, for instance, [Citation10, Exercise 9.1.13]) that this action extends to make A a Uq(sl2)-module algebra. Note that, by construction, this action is degree-preserving.

In the terminology of Vlaar [Citation13, Theorem 6.7], Uq(sl2) has a quasitriangular structure R “up to completion” - meaning that R does not lie in Uq(sl2)Uq(sl2), but rather in a completion of this space. Despite this technicality, R still satisfies axioms (1.1) and (1.2) of a counital 2-cocycle. However, since R is not in Uq(sl2)Uq(sl2), we must check that there is still a well-defined action of R on AA in order for us to define the twist of A by R.

Etingof [Citation8, Remark 3.41] states that given two representations of Uq(sl2), say ρ:Uq(sl2)End(V) and ρ:Uq(sl2)End(W), which are locally nilpotent (i.e. for all vV or W, there exists some nN such that Env=0), we have that (ρρ)(R) is a well-defined operator on VW. Therefore, if the action of Uq(sl2) on A is locally nilpotent then it will follow that R has a well-defined action on AA. The action on A is indeed locally nilpotent, and this can be seen from how E acts on a general basis element of A: (4.2) Exayb=[b]qxa+1yb1,where [b]q:=qbqbqq1(4.2)

It is a simple exercise to check (4.2), first by showing Exayb=xa(Eyb), and then using induction (in the degree b), to show Eyb=[b]qxyb1. From (4.2), we see Eb+1xayb=0, and so Uq(sl2) indeed acts locally nilpotently on A. Hence R has a well-defined action on AA and we may construct the Drinfeld twist AR.

Finally we can apply our first main result, Theorem 2.1. The conditions of the theorem are met since the action of Uq(sl2) on A is degree-preserving. We deduce that AR is a quadratic algebra, and is given by kx,y/(R(xyqyx)). Vlaar [Citation13, Equation 6.37] tells us how R acts on VV explicity, and for the basis vectors xy and yx it is, Rxy=q1/2(xy+(qq1)yx),Ryx=q1/2yxn Therefore Rxyqyx=q1/2(xyq1yx) and so R(xyqyx) is equal to the ideal (xyq1yx). Hence AR is the q1-quantum plane.

4.2 Symmetric and exterior algebras

Here we will consider k=C, and in each example we will take the Hopf algebra H to be the group algebra CT, where T is the finite abelian group given by (C2)n=t1,,tn|ti2=1,titj=tjti for some n2. The counit and coproduct of H are given by ϵ(t)=1 and (t)=tt for all tT. Let, μ=1j<inμij,μij=12(11+ti1+1tjtitj)

By [Citation2, Lemma 4.5], μ is a counital 2-cocycle of H=CT. Note also that μ=μ1.

Consider an n-dimensional C-vector space V with a fixed basis x1,,xn. For an n × n-matrix q=(qij) satisfying qii = 1 and qijqji=1, the corresponding quantum symmetric algebra is defined as Sq(V):=T(V)/(xixjqijxjxi|1i,jn).

Likewise the quantum exterior algebra is given as (V)q:=T(V)/(xixj+qijxjxi|1i,jn). In the following examples we will be interested in the case when q=(1), where (1)ij={1,if i=j1,otherwise

The quantum symmetric and exterior algebras are known to be Koszul, but we show that this can also be deduced as an application of Theorem 3.1.

Example 4.1

(Twisting the symmetric algebra.). Recall that the symmetric algebra S(V)C[x1,,xn] is a Koszul algebra (see Witherspoon [Citation14, Example 3.4.11]), and by definition is equal to T(V)/(R) for R=spanC{xixjxjxi|1i,jn}. Additionally H=CT acts on V via tixj=(1)δijxj, and this extends to make S(V) an H-module algebra. In particular the action on a monomial is given by tix1a1xiaixnan=(1)aix1a1xiaixnan

Note that this action of H on S(V) is degree-preserving, and therefore we may apply Theorem 3.1 to deduce that the Drinfeld twist S(V)μ is a Koszul algebra isomorphic to T(V)/(Rμ), where Rμ=spanC{μ(xixjxjxi)|1i,jn}

By [Citation2, Corollary 5.8], for ij,μ(xixjxjxi)=xixj+xjxi, so S(V)μ=S1(V). Therefore we find that Theorem 3.1 provides a new proof that the quantum symmetric algebra S1(V) is Koszul.

Example 4.2

(Twisting the exterior algebra). The exterior algebra is also Koszul, and is equal to T(V)/(R) for R=spanC{xixj+xjxi|1i,jn}. The action of H=CT on V given by tixj=(1)δijxj also extends by algebra automorphisms to the exterior algebra (V), making (V) an H-module algebra. By construction this action is degree-preserving, and so, by Theorem 3.1, we find (V)μ is a Koszul algebra. Now (V)μ=T(V)/(Rμ) for Rμ=spanC{μ(xixj+xjxi)|1i,jn} =spanC{xixi,xixjxjxi|ij}where we apply [Citation2, Corollary 5.8] in the second equality. Therefore (V)μ=(V)1, and we have found another proof that (V)1 is Koszul.

Acknowledgments

I wish to thank Yuri Bazlov for many helpful discussions.

Additional information

Funding

Research for this paper was supported by the Engineering and Physical Sciences Research Council, UK.

References

  • Aschieri, P., Schenkel, A. (2014). Noncommutative connections on bimodules and drinfeld twist deformation. Adv. Theor. Math. Phys. 18(3):513–612. DOI: 10.4310/ATMP.2014.v18.n3.a1.
  • Bazlov, Y., Berenstein, A., Jones-Healey, E., Mcgaw, A. (2022). Twists of rational Cherednik algebras. Q. J. Math. 74(2):511–528. DOI: 10.1093/qmath/haac033.
  • Brouder, C., Fauser, B., Frabetti, A., Oeckl, R. (2004). Quantum field theory and Hopf algebra cohomology. J. Phys. A: Math. General 37(22):5895. DOI: 10.1088/0305-4470/37/22/014.
  • Chen, X.-W., Silvestrov, S. D., Van Oystaeyen, F. (2006). Representations and cocycle twists of color Lie algebras. Algebras Represent. Theory 9(6):633–650. DOI: 10.1007/s10468-006-9027-0.
  • Davies, A. (2017). Cocycle twists of Algebras. Commun. Algebra 45(3):1347–1363. DOI: 10.1080/00927872.2016.1178271.
  • Drinfel’d, V. G. (1986). Quantum Groups. Proc. Int. Congress Math. 1:798–820.
  • Etingof, P., Gelaki, S., Nikshych, D., Ostrik, V. (2016). Tensor Categories. Volume 205 of Mathematical Surveys and Monographs. Providence, RI: American Mathematical Society.
  • Etingof, P., Semenyakin, M. (2021). A brief introduction to quantum groups. arXiv:2106.05252.
  • Giaquinto, A., Zhang, J. J. (1998). Bialgebra actions, twists, and universal deformation formulas. J. Pure Appl. Algebra 128(2):133–151. DOI: 10.1016/S0022-4049(97)00041-8.
  • Majid, S. (1995). Foundations of Quantum Group Theory. Cambridge, UK: Cambridge University Press.
  • Montgomery, S. (2004). Algebra properties invariant under twisting. In: Caenepeel, S., Van Oystaeyen, F., eds. Hopf Algebras in Noncommutative Geometry and Physics. Boca Raton, FL: CRC Press, p. 239.
  • Vafa, C., Witten, E. (1995). On orbifolds with discrete torsion. J. Geometry Phys. 15(3):189–214. DOI: 10.1016/0393-0440(94)00048-9.
  • Vlaar, B. (2020). LMS Autumn Algebra School 2020, Lecture notes: Introduction to Quantum Groups. https://www.icms.org.uk/sites/default/files/documents/events/Bart%20Vlaar.pdf. [Online; accessed 31-May-2023].
  • Witherspoon, S. J. (2019). Hochschild Cohomology for Algebras. Volume 204 of Graduate Studies in Mathematics. Providence, RI: American Mathematical Society.