Three infinite families of reflection Hopf algebras
Abstract.
Let be a semisimple Hopf algebra acting on an Artin-Schelter regular algebra , homogeneously, inner-faithfully, preserving the grading on , and so that is an -module algebra. When the fixed subring is also AS regular, thus providing a generalization of the Chevalley-Shephard-Todd Theorem, we say that is a reflection Hopf algebra for . We show that each of the semisimple Hopf algebras of Pansera, and and of Masuoka is a reflection Hopf algebra for an AS regular algebra of dimension 2 or 3.
2010 Mathematics Subject Classification:
16T05, 16E65, 16G10Introduction
Throughout let , and denote the square root of by . A finite subgroup of GL, acting linearly as graded automorphisms on a (commutative) polynomial ring , is called a reflection group if is generated by elements , which act on the vector space with fixed subspace of codimension 1; this condition is equivalent to the condition that all the eigenvalues of are , with the exception of one eigenvalue that is a root of unity (sometimes such elements are called pseudoreflections when the exceptional eigenvalue is not ). Chevalley [9] and Shephard and Todd [22] showed that over a field of characteristic zero, a group is a reflection group if and only if the invariant subalgebra is a polynomial ring, and Shephard and Todd [22] presented a complete classification of the reflection groups into three infinite families (the cyclic groups, the symmetric groups, and the groups for positive integers , , and , where divides ), and thirty-four exceptional groups. Reflection groups have played an important role in many contexts, including in representation theory, combinatorics, commutative ring theory, and algebraic geometry.
There has been interest in extending the Chevalley-Shephard-Todd Theorem to a noncommutative context (replacing the commutative polynomial ring with a noncommutative algebra ), and in [13, Definition 2.2] an analog of a reflection (called a quasi-reflection in that paper) was defined for a graded automorphism of an Artin-Schelter regular (AS regular) algebra that is generated in degree 1 (Definition 1.1). When such an AS regular algebra is commutative, it is isomorphic to a commutative polynomial ring, so this particular noncommutative setting generalizes the classical commutative polynomial algebra case. Moreover, examples suggest that the proper analog of a reflection group for is a group such that the invariant subalgebra is also AS regular. The extended notion of the definition of “reflection” of [13] (which involves “trace functions” rather than eigenvalues) was used in [15] to prove a version of the Chevalley-Shephard-Todd Theorem for groups acting on skew polynomial rings (and a second proof was given in [1]). Among the reflection groups for the skew polynomial ring are the dicyclic groups (also known as binary dihedral groups) generated by and with relations: ; so, for example, the quaternion group of order 8 is a reflection group for . These groups are not among the classical reflection groups.
To extend classical invariant theory further, the group can be replaced by a semisimple Hopf algebra (see [16]) that acts on a noncommutative AS regular algebra , and several extensions of results for the action of a finite subgroup of on have been proved in this context (e.g., [7, 5, 6]). However, it has appeared more difficult to extend the Chevalley-Shephard-Todd Theorem to Hopf actions. To this end we consider pairs , where is an AS regular algebra and is a (finite-dimensional) semisimple Hopf algebra, equipped with an action of on that preserves the grading, and is inner-faithful on (meaning that no non-zero Hopf ideal of annihilates , see Section 1.2), with an -module algebra (so that the coproduct of is used to compute the actions of elements of on products of elements of ). We call a reflection Hopf algebra for ([17, Definition 3.2]) if the ring of invariants is AS regular.
In [14, Examples 7.4 and 7.6] it was shown that the Kac-Palyutkin algebra is a reflection Hopf algebra for both and . In [17] the case of a Hopf algebra of the form , the dual of a group algebra (or equivalently, a group coaction) was considered, and some dual reflection groups were constructed. In [10] the sixteen non-trivial Hopf algebras of dimension sixteen classified by Kashina [12] were considered, and the methods used in this paper were used to determine which are reflection Hopf algebras for AS regular algebras of dimension 2 and 3.
In this paper we consider three infinite families of Hopf algebras: the Hopf algebras of dimension defined by Pansera [20], and the two families and of Hopf algebras of dimension defined by Masuoka [18]. We begin in section 2 by considering the Kac-Palyutkin algebra, which occurs as () in the Pansera construction, as well as () in one of the Masuoka constructions. The Pansera construction is a generalization of the Kac-Palyutkin Hopf algebra and is an extension of the form:
The Hopf algebras and can be viewed as deformations of (see [2]), and are extensions of the form:
The examples of reflection Hopf algebras that we have computed indicate that there are an abundance of examples. The properties that characterize such a pair are not clear, and invite further investigation. One obvious question is:
Question 0.1.
When is a bicrossed product a reflection Hopf algebra for some AS regular algebra ?
The method that is used in this paper is as follows. First, we compute the Grothendieck ring of finite-dimensional -modules for each Hopf algebra . The results are summarized in the following table.
Hopf Algebra | |
---|---|
Theorem 3.5 | |
Theorem 4.3 | |
where is odd | Theorem 5.3 |
where is even | Theorem 6.3 |
Using the fusion relations in the Grothendieck ring of , we construct AS regular algebras on which acts inner-faithfully. In the cases of these three infinite families there are always such AS regular algebras of dimension two or three. The table below lists each of the theorems where the inner-faithful representations of are presented.
Hopf Algebra | Inner-Faithful Reps | Dimension |
---|---|---|
Theorem 3.7 | ||
Theorem 4.5 | ||
where is odd | Theorem 5.5 | |
where is even | Theorem 6.7 |
Using the smallest dimension AS regular quadratic algebras on which acts inner-faithfully, we compute the fixed ring and determine when it is also AS regular. We obtain the following theorem:
Theorem 0.2.
The following Hopf algebras are reflection Hopf algebras for the given AS regular algebras.
- (1)
- (2)
- (3)
- (4)
- (5)
- (6)
- (7)
For each of fixed rings the product of the degrees of the minimal generators of the invariants is equal to the dimension of the Hopf algebra. The following conjecture is true for actions of reflection groups on a commutative polynomial ring, and in all the group and Hopf action examples we have computed:
Conjecture 0.3.
Let be an AS regular algebra, and a semisimple reflection Hopf algebra for . Then the product of the degrees of any homogeneous minimal generating set of the algebra is .
The paper is organized as follows. Background material is presented in Section 1, the Kac-Palyutkin algebra is discussed in Section 2, the Pansera algebras are discussed in Section 3, the algebras of Masuoka are discussed in Section 4, and the algebras of Masuoka are discussed in Section 5 ( odd) and Section 6 ( even). In Section 7 we note that an action of on can be extended to an Ore extension with , so that the algebras on which a Hopf algebra acts as a reflection Hopf algebra can have arbitrarily large dimension.
1. Background
We follow the standard notation for Hopf algebras, and refer to [19] for any undefined terminology concerning Hopf algebras. For a Hopf algebra , the set of grouplike elements of is denoted .
1.1. AS regular algebras
We consider Hopf algebras that act on AS regular algebras, which are algebras possessing homological properties of commutative polynomial rings.
Definition 1.1.
Let be a connected graded algebra. Then is Artin-Schelter (AS) regular of dimension if it satisfies the conditions below:
-
(1)
has finite global dimension ;
-
(2)
has finite Gelfand-Kirillov dimension;
-
(3)
satisfies the Gorenstein condition, i.e., for some .
Examples of AS regular algebras include skew polynomial rings and Ore extensions of AS regular algebras; the AS regular rings of invariants we find here will either be commutative polynomial rings or Ore extensions of skew polynomial rings. The AS regular algebras of dimensions 2 and 3 have been classified. We will use the following well-known fact (see e.g., [8, Lemma 1.2]) to show that an invariant subring is not AS regular.
Lemma 1.2.
If is an AS regular algebra of GK dimension 2 (resp. 3), then is generated by 2 (resp. 2 or 3) elements .
We will encounter algebras of the form several times in this paper, so we record here the following lemma which identifies a set of monomials in and as a basis.
Lemma 1.3.
Let be a nonzero element of and . Then is AS regular, and the set of monomials with nonnegative integers and forms a -basis of .
Proof.
A straightforward computation shows that a reduced Gröbner basis of the ideal generated by with respect to the graded lexicographic term order is given by and . It follows that each of the elements in the proposed basis are reduced with respect to this term order and are hence linearly independent. Therefore the coefficients of the Hilbert series of are at least those of the polynomial ring in two variables generated in degree 1. For such that , the change of basis and shows that is a quotient of . However, the Hilbert series calculation shows is in fact isomorphic to and is thus AS regular, and hence the proposed basis in fact spans . ∎
1.2. Inner-faithful actions
An -module is inner-faithful if the only Hopf ideal that annihilates is the zero ideal. We record the following result which is due to Brauer [3], Burnside [4] and Steinberg [23] in the case of a group algebra of a finite group, and due to Passman and Quinn [21] in the case of a finite-dimensional semisimple Hopf algebra. We include a proof for the sake of completeness.
Theorem 1.4.
Let be a module over a semisimple Hopf algebra . Then the following conditions are equivalent.
-
(1)
is an inner-faithful -module,
-
(2)
The tensor algebra is a faithful -module,
-
(3)
Every simple -module appears as a direct summand of for some .
Proof.
If is inner-faithful, then [21, Corollary 10] shows that (3) holds. If (3) holds and if , then it follows that is contained in the Jacobson radical of , which is zero. Finally, suppose holds, and let be a Hopf ideal which annihilates . Then for all and for all , one has
If , then since is a Hopf ideal, each summand of contains some tensor factor which is in . It follows that annihilates , hence . ∎
This result motivates the following definition, which we use in several of the proofs regarding inner-faithful representations that follow.
Definition 1.5.
Let and be -modules over a finite-dimensional semisimple Hopf algebra. If appears as a direct summand of for some , we say that generates .
If each simple representation of occurs as a direct summand of then acts faithfully on . If acts faithfully on , then clearly it acts inner-faithfully. In all the examples that we have checked the following conjecture holds.
Conjecture 1.6.
When a semisimple Hopf algebra acts inner-faithfully on an AS regular algebra , it also acts faithfully on .
It is well-known that a two-sided ideal generated by a skew-primitive element is a Hopf ideal. As is skew primitive for , we have the following lemma.
Lemma 1.7.
Let be a Hopf algebra and . Then the two-sided ideal generated by is a Hopf ideal.
1.3. The Grothendieck ring
To determine if a particular representation of a semisimple Hopf algebra is inner-faithful, by Theorem 1.4 it is necessary to compute the decomposition of tensor powers of into irreducible -modules, i.e., to determine the fusion rules in the Grothendieck ring . We are interested in decompositions of -modules, rather than -comodules, which are used in other contexts (e.g., in [18] refers to -comodules).
For each of the Hopf algebras we consider, the irreducible representations are either one-dimensional or two-dimensional. In what follows we shall use the following notation:
Notation 1.8.
Let be a Hopf algebra with a fixed set of algebra generators . Specifying a -dimensional module is equivalent to providing a matrix for each generator of such that the matrices satisfy the relations of .
For one-dimensional representations, we let be the -vector space with basis such that for . For two-dimensional representations, we let be the -vector space with basis and and denote by the matrix that provides the action of on .
Remark 1.9.
Let be a semisimple Hopf algebra. In this paper, we search for AS regular algebras on which acts (inner-faithfully) using the following procedure. Let be a graded algebra generated in degree 1. If is an (inner-faithful) -module, we extend the action of to using the coproduct of . This action passes to if and only if is an -submodule of .
In this paper we only study actions on quadratic algebras, hence we may also assume . Therefore the possible relations for algebras on which acts are governed by the decomposition of into simple -modules, and these are further restricted by those relations that give AS regular algebras. Lemma 1.2 aids us in identifying algebras that are not AS regular in many of our examples.
1.4. Generating invariants
The following lemmas are useful in finding minimal generating sets for some subrings of invariants.
Lemma 1.10.
Let and be commuting elements of a ring, and let with .
-
(1)
is in the subalgebra .
-
(2)
is in the subalgebra .
Proof.
The proofs are by induction on . We give the proof for (2), the proof for (1) being similar.
When is odd,
and the result follows by induction. When is even the proof is similar, with an extra middle term . Hence the result holds in all cases. ∎
Lemma 1.11.
Let be the following ring
for some and some positive integer . Then the element
where are nonnegative integers not all zero, is contained in the subalgebra generated by the elements
Proof.
If then the lemma follows from Lemma 1.10, therefore we can generate for all . Now we prove that is generated by induction on . If then the element is one of the generators. Otherwise assume that the elements have been constructed, then the equality
shows that can be constructed using the claimed generators.
So far we have proved that the elements and are generated for all . Now we show that all the elements of the form are generated. Indeed if is even then and if is odd then .
Now we prove by induction that can be generated. If then is one of the generators. Recall that the element has already been constructed and assume that has been constructed, then the equality
shows that can be constructed.
To generate we observe that
To show that is generated we just notice that if is even then and if is odd then .
To conclude the proof we show that can be generated. If is even then , if is odd then . ∎
Remark 1.12.
The proof of the previous Lemma shows that if is even then is in the subalgebra generated by .
Remark 1.13.
Adopting the notation of Lemma 1.11, we notice that in the element is in the subalgebra generated by .
Remark 1.14.
Adopting the notation of Lemma 1.11, we notice that in the element is in the subalgebra generated by .
Lemma 1.15.
In the ring the element
is in the subalgebra .
Proof.
We first notice that one can generate , indeed
Therefore is generated for all even (and it is zero when is odd).
Now we show by induction that and can be generated. We first check it for , indeed
We now assume that and have been constructed for . We consider the case even first and show how to generate . Indeed
We can assume that is odd, otherwise the only extra term in the previous summation is , which we have already proved may be generated. In this case
and therefore can be generated. If is odd then
and we are done by induction.
Now we show that one can generate . If is even then
If is odd then
To see that one may obtain for , notice that
∎
2. The Kac-Palyutkin Hopf algebra
Let be the Kac-Palyutkin algebra of dimension 8; it is the smallest dimensional semisimple Hopf algebra that is neither commutative nor cocommutative (nor is it a twist of such a Hopf algebra). As an algebra, is generated by and with relations
The coproduct in given by
while the counit is defined by
and the antipode is the anti-automorphism given by
Using Notation 1.8, there are four one-dimensional representations of , namely:
There is a unique two-dimensional irreducible representation given by the matrices:
Using the coproduct of we compute that under this action
One can check that as an -module,
so that all irreducible -modules occur as direct summands of or . Hence, is an inner-faithful -module. Moreover, the basis elements of the one-dimensional summands indicated above give us the possible quadratic algebras on which acts.
In summary, we see that will act inner-faithfully on if as an -module, and is any one of the four basis elements of the one-dimensional -modules occurring as a direct summand in as listed above. Moreover, in each case the algebra is AS regular of dimension 2. The table below gives the relation and the corresponding fixed ring in each of these cases. One can check that in each case there is a copy of in the algebra , so that the action is actually faithful.
Case | Relation | Fixed Ring | |
---|---|---|---|
(a) | commutative hypersurface | ||
(b) | |||
(c) | |||
(d) |
Summarizing we have the following theorem.
Theorem 2.1.
The Kac-Palyutkin Hopf algebra acts (inner-)faithfully on the AS regular algebras and on with fixed subring a commutative polynomial ring, and hence is a reflection Hopf algebra for each of these three AS regular algebras of dimension two.
3. The Hopf algebras of Pansera
In [20] D. Pansera defined an infinite family of semisimple Hopf algebras of dimension that act inner-faithfully on certain quantum polynomial algebras. When , this Hopf algebra is the -dimensional semisimple algebra defined by Palyutkin [11], which was discussed in the previous section. We begin by reviewing the construction of these algebras. Fix an integer .
Let be the direct product of two cyclic groups of order and let denote , a primitive th root of unity. A complete set of orthogonal idempotents in the group algebra is given by , where
similarly, we define by
Let denote the automorphism of given by , and define the element by
Note that is a right Drinfel’d twist of ([20, Lemma 2.10]). Letting denote the multiplication map on , one may show that is invertible in , and that . Finally, define as the factor ring of the skew polynomial extension of :
In [20] it is shown that is a Hopf algebra, with vector space basis
where the Hopf structure of is extended to by setting:
3.1. The Grothendieck ring
We now fix a square root of , and denote it by . When is odd, is a primitive th root of unity. To give the irreducible representations of , we first record a lemma.
Lemma 3.1.
Let be a primitive root of unity. Then for all , one has the equality
Proof.
The lemma follows from the following string of equalities:
where the last equality follows from the fact that an root of unity different from 1 is a root of the polynomial . ∎
Below, we again use Notation 1.8 to describe the representations of .
Proposition 3.2.
Let . Then the one-dimensional vector space with -action given by is an -module, and all one-dimensional -modules are of this form. Furthermore, for all , the two-dimensional vector space with -action given by
is an -module.
Proof.
Let be a one-dimensional representation generated by . The relations and tell us that and must act on as an th root of unity. The relation tells us that and must act on as the same root of unity . Indeed, to understand the action of on , Lemma 3.1 implies that , and hence must act on by multiplication by . This action obviously satisfies all the other relations.
Let be a two-dimensional vector space generated by elements and over which acts as stated in the theorem. We prove that this action satisfies the relations of the algebra. Again, the only relation that is not obviously satisfied is the relation involving .
Again using Lemma 3.1, on a vector of the element acts as the matrix
It just remains to notice that . ∎
Remark 3.3.
Note that one has , and that the representation is reducible. Indeed, a straightforward computation shows
(3.1) |
We also abuse notation and read the subscripts of modulo , since the action of depends only on the subscripts modulo .
Theorem 3.4.
The representations for and for are a complete set of irreducible representations of .
Proof.
The fact that is irreducible follows from the fact that one may not simultaneously diagonalize and when . Furthermore, the representations considered are distinct since the matrices have different spectra.
Hence, we have found one-dimensional representations and two-dimensional irreducible representations of . These are all the irreducible representations since
Theorem 3.5.
The ring has the fusion rules given below, where and the subscripts on the right-hand side of each equality are to be read modulo as mentioned in Remark 3.3.
(3.2) |
Proof.
A computation shows that the action on is as follows:
The first two equalities in the statement of the theorem follow immediately from the previous table. We show the computation for :
where the fourth equality follows from Lemma 3.1. The other equalities follow similarly. ∎
Proposition 3.6.
The ring is isomorphic to .
Proof.
A similar argument to the one used in the proof of Theorem 3.4 shows that all the one-dimensional representations of are of the form (where is defined in a manner similar to ) and all the two-dimensional irreducible representation are of the form with and
Now it is just a matter of proving that the multiplication in is the same as in . A computation shows that the action on is as follows
which gives
The other products are checked similarly. ∎
Theorem 3.7.
If is an algebra generated by in degree 1 where then the action of on is inner-faithful if and only if for .
Proof.
Let be a Hopf ideal of such that . We denote by the group algebra , and think of it as a Hopf subalgebra of generated by and . Then is a Hopf ideal of . By [20, Lemma 1.4] there is a normal subgroup of such that
(3.3) |
Since there is such that . By 3.3 , hence
This implies that
i.e., the vector is in the kernel of the matrix
If then is injective hence and , which implies . By [20, Lemma 2.12], which means that the action is inner-faithful. If is not injective then consider a nonzero vector such that then and so the action is not inner-faithful. ∎
3.2. Inner-faithful Hopf actions of on AS regular algebras and their fixed rings
Recall that and (which is a primitive th root of unity when is odd). Let act on a quadratic algebra with two generators and for . Then by Theorem 3.5
Then by Remark 3.3 decomposes into one-dimensional representations whose basis element could be taken to be the relation in an algebra that acts upon, namely
Using either of these basis elements as the relation in and recalling that shows that the Hopf algebra acts on the AS regular algebras of dimension 2
Remark 3.8.
When is even the representation decomposes as a sum of two representations of dimension one if , giving extra choices for a relation in . But by Theorem 3.7 these actions are inner-faithful if and only if . The conditions that is even, , and can be simultaneously true only if . Therefore, the “extra” inner-faithful algebra actions can occur only if , which means the algebra is the Kac-Palyutkin algebra, which was analyzed in Section 2.
We first consider the algebra , which we denote by . We want to compute the fixed ring and determine when it is AS regular.
Lemma 3.9.
The action on the monomials of is given by
(3.4) |
Proof.
We first prove that
(3.5) |
It is clear if . We assume it is true for and prove it for :
where the last equality follows from Lemma 3.1. Similarly one proves
(3.6) |
In particular, one has
and for all . This implies that is a fixed element for all , is fixed when is odd and is fixed when is even.
Theorem 3.10.
If acts inner-faithfully by (i.e., if for ) on
where , then the fixed subring is
Hence is a reflection Hopf algebra under this action for when .
Proof.
Let be an element of fixed by the action of . Then
and
so in order for to be fixed by the action of and we must have
which is equivalent to
but this kernel is zero because , hence .
Now we assume odd, the case even is similarly proved. A fixed element must be of the form
(3.8) |
and by using (3.4) we get
because
In order for an element of the form (3.8) to be fixed by the action of we must have for all . Hence a fixed element has the form
It follows from Lemma 1.10, by setting and , (since is odd so and commute), that this invariant ring is generated by .
It remains only to prove that the generators of the fixed ring are algebraically independent. This follows because they form a regular sequence in the commutative Cohen-Macaulay ring . ∎
When one checks similarly that the action of on monomials in is as given in the following lemma.
Lemma 3.11.
The action of on monomials of is given by
When is even the computations above and Lemma 3.11 show that is the same for as for . Using an argument similar to the even case, one can show that when is odd an invariant must have the form
We may rewrite the previous expression as
Applying Lemma 1.15 with and shows that and generate the invariants. The subring of generated by and is the subring invariant under the transposition [15, Example 3.1], which is not AS regular [15, Theorem 1.5(2)].
We summarize these cases in the following theorem.
Theorem 3.12.
When acts by on
for , inner-faithfully (i.e., for ) then
-
(1)
when is even, the invariant ring is
and is a reflection Hopf algebra for ,
-
(2)
when is odd, the invariant ring is
which is not AS regular.
4. The Hopf algebras of Masuoka
The -dimensional Hopf algebras and for were defined by Masuoka in [18, Definition 3.3]. Note that is the Kac-Palyutkin algebra considered in Section 2. Let be the group algebra of a cyclic group of order 2; is identified with its dual , in which and are the idempotents and . The Hopf algebras and are defined as algebras over , with a central Hopf subalgebra (with group-like). The Hopf algebra is generated as an algebra over by the two elements and with the relations:
The coproduct, counit and antipode in are:
is defined in the same way, except the relation is replaced by the relation .
Next we compute the Grothendieck rings of the irreducible modules of . We note that the Grothendieck rings given in [18] are for the irreducible comodules. The irreducible modules for these algebras are all one-dimensional or two-dimensional.
4.1. The Grothendieck ring
The next proposition is straightforward.
Proposition 4.1.
The one-dimensional representations of are of the form . The irreducible two-dimensional representations are
where is a primitive th root of unity, .
Notation 4.2.
We denote by and the (reducible) two-dimensional representations given by
A straightforward computation shows
The next two results are also straightforward.
Theorem 4.3.
The Grothendieck ring has the following fusion rules:
when is even, and
when is odd. In addition
We notice that the dihedral group has the same representations as . Comparing the fusion rules of the two Grothendieck rings we notice that there is an isomorphism between them defined as and . We have proved the following
Proposition 4.4.
The ring is isomorphic to .
We use the previous proposition to analyze the inner-faithful representations of by reducing the problem to analyzing the inner-faithful representations of .
Theorem 4.5.
Let be a -algebra generated in degree 1 by and with . The action of on is inner-faithful if and only if for .
Proof.
By Proposition 4.4, a representation generates (see Definition 1.5) if and only if it generates , so from now on we will work with the group algebra and will be a -algebra generated in degree 1 by and with over which acts.
Let be a Hopf ideal of such that with . Then by [20, Lemma 1.4] there is a normal subgroup of such that . Since is not trivial then neither is , hence there is a element in of the form
or
We first deal with the case . In this case and hence it annihilates . But this element acts on as the matrix
which is never zero, a contradiction. Hence all the nontrivial elements in must be of the form with . As a result there is an element in of the form with . The element acts on as
This matrix is zero if and only if if and only if for some which is equivalent to .
Hence if the action is not inner-faithful then . If then choose between 1 and such that is a multiple of . The Hopf ideal generated by annihilates and hence the action is not inner-faithful.
∎
4.2. Inner-faithful Hopf actions of on AS regular algebras and their fixed rings
Noting when has one-dimensional summands, by Theorems 4.3 and 4.5 the Hopf algebra acts inner-faithfully on the AS regular algebras of dimension 2 for ,
for , a primitive th root of unity.
We first set , , and calculate .
Remark 4.6.
Theorem 4.7.
If acts on inner-faithfully by (i.e. for ), then
and hence is a reflection Hopf algebra for .
Proof.
It is easy to check that and are fixed.
If an element is fixed by then it must be of even degree, hence, by Lemma 1.3, it is of the form
A computation shows
Setting yields and setting yields . Both identities, combined, lead to . Hence setting and using the fact that is odd we deduce that must have the form
This element is generated by the claimed elements by Lemma 1.10 since is central in .
It remains only to prove that the generators of the fixed ring are algebraically independent. Let , and . Then the algebra is isomorphic to since the latter algebra is a commutative domain (as the element is irreducible by Eisenstein’s criterion), and the former algebra has GK dimension two. It follows that is Cohen-Macaulay. Since and form regular sequence in a Cohen-Macaulay algebra, they are algebraically independent. ∎
Next we consider , and show that is not a reflection Hopf algebra for .
Theorem 4.8.
The fixed ring for the inner-faithful action of on by (i.e. for ) is
Furthermore, this ring is not AS regular.
Proof.
The argument is similar to the proof of Theorem 4.7. Any invariant must have even degree so
The action of is
As before, this implies that for some integer . Thus,
Using Remark 1.14 by setting , we see that the generators above indeed generate the fixed ring. To see that they are all necessary, note that the form of an invariant above implies that for even, the Hilbert series of begins as:
and for odd, it begins as
Therefore it is clear one needs the generators and . The subalgebra generated by these invariants is zero in degree when is even and spanned by a power of in degree when is odd. In either case, is not generated by the other two generators, hence all three are necessary. Since has dimension 2, if the invariant ring were AS regular, it would also be AS regular of dimension 2, and hence by Lemma 1.2 the invariant ring is not AS regular. ∎
5. The Hopf algebras of Masuoka for odd
Recall that the Hopf algebras were defined in Section 4 as follows. The group algebra of a cyclic group of order 2 is identified with its dual , in which and are the idempotents and . The Hopf algebras are defined as algebras over , with a central Hopf subalgebra (with group-like), and generated over by the two elements and with the relations :
the coproduct, counit and antipode in are:
5.1. The Grothendieck ring ( odd)
In this section represents a primitive th root of unity (not a th root of unity, as it was in the case of ). The next proposition is straightforward.
Proposition 5.1.
Let be odd. The one-dimensional representations of are of the form and . The two-dimensional irreducible representations are
where is a primitive th root of unity, and .
Notation 5.2.
We denote by the (reducible) two-dimensional representation given by
with . A straightforward computation shows
Theorem 5.3.
The ring when is odd has the following fusion rules:
Proof.
A computation shows that the action on is
from which the first equality follows. The other equalities are proved similarly. ∎
Proposition 5.4.
When is odd, the ring isomorphic to .
Proof.
As algebras, and are isomorphic, hence and are isomorphic as abelian groups. Abusing notation, we denote the irreducible representations of in the same way we denoted the ones of . The multiplication is clearly the same in both rings since in this case act on as group-likes for both algebras. The table for the action of on is
which gives the following decomposition
hence the product is the same. The other products are similarly checked. ∎
Theorem 5.5.
Let be a graded -algebra generated in degree 1 by and with . The action of , odd, on is inner-faithful if and only if and for .
Proof.
If then the Hopf ideal generated by annihilates and hence the action is not inner-faithful by Lemma 1.7. From now on we will assume . By 5.4 a representation generates if and only if it generates , so from now on we will work with the group algebra with , and will be a graded -algebra generated in degree 1 by and with on which acts.
Let be a Hopf ideal of such that with . Then by [20, Lemma 1.4] there is a normal subgroup of such that . Since is not trivial then neither is . Furthermore, since acts as a scalar matrix, there must be an integer with such that either or with . We may now conclude as in the proof of Theorem 4.5 that .
Hence, if the action is not inner-faithful then . If , then choose between 1 and such that is a multiple of , so that the Hopf ideal generated by annihilates , and hence the action is not inner-faithful. ∎
5.2. Inner-faithful Hopf actions of ( odd) on AS regular algebras and their fixed rings
Recall that we are assuming that is odd, and that is a primitive th root of unity. The Hopf algebras act on the AS regular algebras of dimension 2
We denote the algebra by and the Hopf algebra by , and we compute the fixed ring .
Theorem 5.6.
If , odd, acts on
for by inner-faithfully (i.e., for ), then
Hence is a reflection Hopf algebra for this action on when .
Proof.
A straightforward calculation shows that and are invariant. More generally, if an element is fixed by then every monomial in it must have even total degree. By Lemma 1.3, any homogeneous element in of even degree must be of the form
The action of and on the basis used in the expression above is:
Setting gives , and setting gives for all . It follows that implies . Since is odd and , we have . Using the fact that is an root of unity, we have that an invariant has the form
Theorem 5.7.
If , odd, acts on
for by , inner-faithfully (i.e., for ), the fixed subring is
Furthermore, the ring is not AS regular.
6. The Hopf algebras of Masuoka for even
6.1. The Grothendieck ring ( even)
The next proposition is straightforward.
Proposition 6.1.
If is even then the one-dimensional irreducible representations of are of the form , , and . The two-dimensional irreducible representations are
where is a primitive th root of unity, and .
Notation 6.2.
Similar to Notation 5.2, we define the following (reducible) two-dimensional representations below, where :
A straightforward computation shows that:
The proof of the following theorem is similar to that of Theorem 5.3.
Theorem 6.3.
The ring when is even has the following fusion rules:
∎
Remark 6.4.
In order to make the decompositions above more palatable, we introduce a final piece of notation. For , we set . Then the tensor product decompositions above, given without reference to a basis, become the following for all , where the subscripts are read modulo :
(6.1) | |||||
Note that the group of one-dimensional representations is isomorphic to the dihedral group of order 8, where the quarter rotations are given by and , the center is generated by , and the reflections are and . Since the Grothendieck ring is not commutative, it cannot be isomorphic to the Grothendieck ring of a group, and therefore an approach different from those used in the earlier examples is needed.
Theorem 6.5.
When is even, none of the representations of for is inner-faithful.
Proof.
The two-dimensional representations with positive exponent cannot be inner-faithful because they cannot generate a two-dimensional irreducible representation with a negative exponent. A necessary condition for to be inner-faithful is , since all the two-dimensional irreducible representations appearing as direct summands of a tensor power of have an index that is an integer combination of and ; if were inner-faithful then one of these combinations would be equal to 1, showing that . Hence we assume that .
By the first equality in (6.3), the two-dimensional summands of are of the form for . Therefore, we obtain (at most) one-half of the two-dimensional irreducible representations of from . ∎
Before characterizing the three-dimensional inner-faithful representations, we prove a lemma which aids us in our reductions.
Lemma 6.6.
When is even, a representation of is inner-faithful if and only if it generates all the irreducible two-dimensional representations up to a sign and and for some .
Proof.
One implication is obvious. The fusion rules show that by decomposing , , and , generates all the one-dimensional irreducible representations. By tensoring an irreducible two-dimensional representation by we change its exponent. It follows that is inner-faithful. ∎
Theorem 6.7.
When is even, the inner-faithful three-dimensional representations of are given as follows, where and :
-
(1)
,
-
(2)
,
-
(3)
,
-
(4)
, provided
-
(5)
, provided .
Proof.
In this proof, we make frequent use of the relations appearing in equations (6.1). Let be an inner-faithful three-dimensional representation of . Certainly must have a two-dimensional irreducible summand, so we let .
If and , then the second isomorphism in equations (6.1) shows that does not assist in generating new irreducible two-dimensional representations, so Theorem 6.5 gives a contradiction. If and , then one must have that in order to obtain an irreducible two-dimensional representation with subscript 1, since tensoring with does not change the subscript. Cases (1) and (2) then follow by Lemma 6.6.
Therefore we may assume that . For to be inner-faithful, the first and third isomorphisms in equations (6.1) show that . Hence either , or is even and . In the latter case, only generates two-dimensional representations of the form , which all have even subscripts. Since is odd, tensoring with only generates representations of the form where is odd, and hence cannot yield a pair of irreducible two-dimensional representations with the same subscript but different exponent. Therefore, we have that , and hence we are in a situation where all two-dimensional representations are generated up to sign.
If , then one must have , otherwise does not generate any ‘negative’ representations. The case indeed gives an inner-faithful representation since generates all ‘positive’ two-dimensional representations and tensoring with gives one of the ‘negative’ two-dimensional representations required by Lemma 6.6.
Next, suppose . Then generates . Since is odd, if is odd, then is also odd, and if is even, then is also even. Therefore all irreducible two-dimensional representations generated by with even (respectively, odd) subscripts have positive (respectively, negative) exponents. If , then is odd, so that while does not help generate any new two-dimensional representations, does. This gives us case (5) in the table above. Similarly when , is even, the roles of and are reversed, giving us case (4). ∎
6.2. Inner-faithful Hopf actions of ( even) on AS regular algebras and their fixed rings
When is even then, by Theorem 6.3, acts on the following AS regular Ore extensions of dimension 3, where and ; recall that is a primitive th root of unity .
Note that in each case is an automorphism of the coefficient ring.
Before we analyze these actions, we prove a lemma which helps shrink the search space for invariants.
Lemma 6.8.
Let be a finite dimensional semisimple Hopf algebra and let be an algebra which is a domain. Suppose that acts linearly on the Ore extension such that the action on restricts to an -action on , and that is a one-dimensional -module. If is not a direct summand of as an -module, then there are not any nonzero elements in of the form , where .
Proof.
Let be an -invariant. Let be the -submodule of generated by . Then the -submodule of is isomorphic to . Since is an invariant, the trivial representation must appear as a summand of . It follows that the inverse (in ) of must appear as a direct summand of . Since has finite order in and is a domain, we must therefore have that appears as an -direct summand of . ∎
Proposition 6.9.
For any of the actions of ( even) on the Ore extensions given above, all invariants involve only even powers of .
Proof.
In all cases, the hypotheses of Lemma 6.8 are met for all odd . ∎
Theorem 6.10.
The fixed rings for the inner-faithful actions of , even (see Theorem 6.7), on the previous algebras are:
where is the induced automorphism. Furthermore, ( even) is a reflection Hopf algebra for , , , but not for and .
Proof.
By Proposition 6.9, we may assume all invariants in all cases involve only even powers of . Also, since the actions in each case are graded we may assume all invariants are homogeneous.
Now consider an invariant , written in the usual basis of :
A computation shows that
Setting and yields and . This implies that . After relabeling the coefficients, we have that must therefore be of the form
Lemma 1.10 shows that can be generated by the invariants . The fixed ring is AS regular since generate a commutative polynomial ring, and is an Ore extension, where is the induced automorphism, and hence the fixed ring is AS regular.
Since is an Ore extension of an algebra of the form appearing in Lemma 1.3, we will use the basis (where are nonnegative integers and is or ). Since the power of must be even, the action of on the base of the Ore extension shows that the total degree of an invariant must be even as well.
Since is fixed, it is enough to find the elements in the subalgebra generated by and that are fixed by and . A computation shows that the action of on the following monomials, where is odd and is arbitrary, is:
(6.2) |
Now suppose is an invariant in and alone:
Using (6.2) we see that in order for to be invariant must be congruent to zero modulo . After reindexing the coefficients, may be written as:
Hence Lemma 1.10 shows that all invariants in are generated by and . The fixed ring is AS regular since generate a commutative polynomial ring, and is an Ore extension, where is the induced automorphism.
For one can show by an argument similar to that for that an invariant must have the form
Using Remark 1.13 with , we see that and generate the invariant subring. To see the proposed generators are all necessary, note that the algebra is bigraded by setting the bidegree of and to be and the bidegree of to be . By the above description of the invariants, the bigraded Hilbert series of in bidegree less than in lexicographic order is given by
where we use to represent bidegree and to represent . Therefore the subalgebra generated by and is spanned by in bidegree and is zero in bidegree , hence the generators and are both necessary. By Lemma 1.2 the invariant ring is not AS regular.
For , again using Proposition 6.9 and the definition of the action one may show that an invariant must be of the form
Now we conclude that and generate the invariant subring by applying Remark 1.13 with . That the proposed generators are necessary follows from the same bigraded Hilbert series argument as in the case, again by Lemma 1.2 this algebra is not AS regular.
For the algebra , we may conclude as for the algebra since the base of the Ore extension as well as the action on it are identical, is invariant, and the total degree of an invariant must be even. ∎
Theorem 6.11.
Let . The fixed rings for the inner-faithful actions of , even (see Theorem 6.7) on are not AS regular:
Hence , even, is not a reflection Hopf algebra for any of the algebras .
Proof.
To determine the invariants, we proceed as in Theorem 6.10, keeping in mind that all invariants must involve only even powers of by Proposition 6.9.
The action of on is given by
for even, and so the invariants are of the form
where are nonnegative integers, with for some , and is even. Relabeling our coefficients, one may write the previous display as
Now we apply Remark 1.12, with .
The action of on the even degree monomials of the base ring of is given as follows. When is an even integer and is arbitrary:
(6.3) |
Since is invariant and due to the action of , we need only consider which elements in of even degree are fixed. Therefore, let
be an invariant in only and where each is odd. The actions of and imply that . Writing for some integer , and again relabeling our coefficients, we have
We conclude by using Remark 1.12 with .
For one shows that:
An element is an invariant if and only if and the element has the form
which can be written as
Now one concludes using Lemma 1.11 by setting .
For the algebra , we may conclude as for the algebra since the base of the Ore extension as well as the action on it are identical, is invariant, and the total degree of an invariant must be even.
That each of the proposed generating sets for and is minimal follows from Hilbert series arguments as in the proofs of Theorems 4.8 6.10. For and , note that all three generators in and alone are necessary, as well as at least one generator involving a . Therefore Lemma 1.2 shows that they are not AS regular.
∎
7. Extension rings
We conclude this paper with a result which allows one to extend a Hopf action on an algebra to an action on an Ore extension of . This allows one to extend the Hopf actions in the earlier sections to algebras of larger GK dimension.
Proposition 7.1.
Let be a Hopf algebra that acts linearly on an algebra , and let be an automorphism of which is also an -module homomorphism. Let be the trivial -module. Then the action of extends to , and . Further, if the action of is inner-faithful on then it is inner-faithful on .
Proof.
Recall that the trivial module satisfies . To prove that acts on , we need only check that the set of relations from the Ore extension are closed under the action. To see this, let and . Write . Then one has
To see that , suppose that is in . Then one has
Since form a basis of as an -module, the result follows. The inner-faithful claim is clear. ∎
Remark 7.2.
Note that of course any scalar map is a Hopf module map, and if is connected graded -algebra with algebraically closed and generated in degree one by an irreducible -module, these will be the only graded -module endomorphisms of .
Remark 7.3.
In order to provide an action of on the Ore extension , the hypothesis that is a trivial module is stronger than what is actually required. Indeed let be any one-dimensional module and let be the algebra map associated to the representation. Then the proof above shows that acts on provided is -linear and for all , one has . If we let denote the unit of , then this condition is equivalent to saying that , where denotes the convolution product of the algebra .
This holds if is the trivial -representation given by the counit (since is the identity of ), or if is a commutative or cocommutative Hopf algebra. One may also check that many of the one-dimensional representations of the Hopf algebras considered in this paper also satisfy this condition. However, one may check that the representation of ( even) does not.
Lastly we note that of course the claim regarding the fixed ring of these more general actions no longer holds, and there can often be many more elements that are fixed by the action of than just (where is the order of the representation in the Grothendieck ring of ). Determining the invariant subrings of such actions would be of significant interest.
Acknowledgement: The authors thank the referee for several helpful suggestions.
References
- [1] Y. Bazlov and A. Berenstein, Mystic reflection groups, SIGMA Symmetry Integrability Geom. Methods Appl. 10 (2014), Paper 040, 11. MR 3210595
- [2] J. Bichon and S. Natale, Hopf algebra deformations of binary polyhedral groups, Transform. Groups 16 (2011), no. 2, 339–374. MR 2806496
- [3] R. Brauer, A note on theorems of Burnside and Blichfeldt, Proc. Amer. Math. Soc. 15 (1964), 31–34. MR 0158004
- [4] W. Burnside, Theory of groups of finite order, Dover Publications, Inc., New York, 1955, 2d ed. MR 0069818
- [5] K. Chan, E. Kirkman, C. Walton, and J. Zhang, McKay correspondence for semisimple Hopf actions on regular graded algebras, I, J. Algebra 508 (2018), 512–538.
- [6] by same author, McKay correspondence for semisimple Hopf actions on regular graded algebras. II, J. Noncommut. Geom. 13 (2019), no. 1, 87–114. MR 3941474
- [7] K. Chan, E. Kirkman, C. Walton, and J. J. Zhang, Quantum binary polyhedral groups and their actions on quantum planes, J. Reine Angew. Math. 719 (2016), 211–252. MR 3552496
- [8] J. Chen, E. Kirkman, and J. J. Zhang, Rigidity of down-up algebras with respect to finite group coactions, J. Pure Appl. Algebra 221 (2017), no. 12, 3089–3103. MR 3666738
- [9] C. Chevalley, Invariants of finite groups generated by reflections, Amer. J. Math. 77 (1955), 778–782. MR 0072877
- [10] L. Ferraro, E. Kirkman, W. F. Moore, and R. Won, Semisimple reflection Hopf algebras of dimension sixteen, arXiv:1907.06763 (2019).
- [11] G.I. Kac and Palyutkin. V.G., Finite ring groups, Trudy Moskov. Mat. Obsc 15 (1966), 224–261, (Russian).
- [12] Y. Kashina, Classification of semisimple Hopf algebras of dimension 16, J. Algebra 232 (2000), no. 2, 617–663. MR 1792748
- [13] E. Kirkman, J. Kuzmanovich, and J. J. Zhang, Rigidity of graded regular algebras, Trans. Amer. Math. Soc. 360 (2008), no. 12, 6331–6369. MR 2434290
- [14] by same author, Gorenstein subrings of invariants under Hopf algebra actions, J. Algebra 322 (2009), no. 10, 3640–3669. MR 2568355
- [15] by same author, Invariants of -skew polynomial rings under permutation representations, Recent advances in representation theory, quantum groups, algebraic geometry, and related topics, Contemp. Math., vol. 623, Amer. Math. Soc., Providence, RI, 2014, pp. 155–192. MR 3288627
- [16] by same author, Invariant theory of finite group actions on down-up algebras, Transform. Groups 20 (2015), no. 1, 113–165. MR 3317798
- [17] by same author, Nakayama automorphism and rigidity of dual reflection group coactions, J. Algebra 487 (2017), 60–92. MR 3671184
- [18] A. Masuoka, Cocycle deformations and Galois objects for some cosemisimple Hopf algebras of finite dimension, New trends in Hopf algebra theory (La Falda, 1999), Contemp. Math., vol. 267, Amer. Math. Soc., Providence, RI, 2000, pp. 195–214. MR 1800713
- [19] S. Montgomery, Hopf algebras and their actions on rings, CBMS Regional Conference Series in Mathematics, vol. 82, Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 1993. MR 1243637
- [20] D. Pansera, A class of semisimple Hopf algebras acting on quantum polynomial algebras, Rings, modules and codes, Contemp. Math., vol. 727, Amer. Math. Soc., Providence, RI, 2019, pp. 303–316. MR 3938158
- [21] D. S. Passman and D. Quinn, Burnside’s theorem for Hopf algebras, Proc. Amer. Math. Soc. 123 (1995), no. 2, 327–333. MR 1215204
- [22] G. C. Shephard and J. A. Todd, Finite unitary reflection groups, Canadian J. Math. 6 (1954), 274–304. MR 0059914
- [23] R. Steinberg, Complete sets of representations of algebras, Proc. Amer. Math. Soc. 13 (1962), 746–747. MR 0141710