February 23, 2005
The authors gratefully acknowledge support from the following grants: NSF grants #DMS–0245082 and #DMS–0443476, and NSA grants MSPF-02G-082 and MDA904-01-1-0067. The second author thanks the Mathematical Sciences Research Institute for its support during the writing of this paper.
QUANTUM GROUP ACTIONS, TWISTING ELEMENTS, and DEFORMATIONS OF ALGEBRAS
Georgia Benkart
Sarah Witherspoon
Department of Mathematics, University of Wisconsin, Madison, Wisconsin 53706, USA E-mail address : benkart@math.wisc.edu Department of Mathematics, Texas A&M University, College Station, Texas 77843, USA E-mail address : sjw@math.tamu.edu
-
Abstract.
We construct twisting elements for module algebras of restricted two-parameter quantum groups from factors of their
-matrices. We generalize the theory of Giaquinto and Zhang to universal deformation formulas for categories of module algebras and give examples arising from
-matrices of two-parameter quantum groups.
1 Introduction
The algebraic deformation theory of Giaquinto and Zhang [GZ] laid out a framework for certain types of deformations of associative algebras, namely those arising from actions of bialgebras on algebras. This work generalized known examples of deformations coming from the action of a Lie algebra (equivalently of its universal enveloping algebra). Giaquinto and Zhang illustrated their ideas by constructing a new deformation formula from the action of a noncommutative bialgebra, namely, the universal enveloping algebra of a certain nonabelian Lie algebra. This Lie algebra is spanned by the
matrix units
for
, along with the
diagonal matrices of trace 0, and so it is an abelian extension of a Heisenberg Lie algebra when
, and it is a two-dimensional nonabelian Lie algebra when
. In recent work, Grunspan [G] applied the deformation formula for the
case to solve the open problem of giving an explicit deformation of the Witt algebra.
In their study of deformations and orbifolds, Căldăraru, Giaquinto, and the second author [CGW] constructed a deformation from a noncocommutative (and noncommutative) bialgebra – the Drinfel'd double of a Taft algebra. This was related to unpublished work of Giaquinto and Zhang and was the first known example of an explicit formula for a formal deformation of an algebra arising from a noncocommutative bialgebra.
The present paper began as an attempt to put the example of [CGW] into a general context. The algebra in [CGW] to undergo deformation was a crossed product of a polynomial algebra and the group algebra of a finite group. We expected that generalizations of the Drinfel'd double of a Taft algebra, such as the two-parameter restricted quantum groups
of our paper [BW3] , would act on similar crossed products, potentially leading to deformations. Indeed,
does act on such a crossed product, but we were unable to generalize the deformation formula of [CGW] directly to deform the multiplication in the crossed product via the
-action. Instead, a generalization in a different direction proved more successful in deforming these particular types of algebras (see [Wi] ). However, as we discuss in Sections 3 through 5 in this paper, certain related noncommutative algebras and their crossed products with group algebras also carry a
-action. Since
is a quotient of the infinite-dimensional two-parameter quantum group
, such algebras also have an action of
, and in some cases they admit formal deformations arising from this action.
In order to obtain these examples, we found it necessary to generalize the definitions of twisting elements and universal deformation formulas given in [GZ] , from elements in bialgebras to operators on particular categories of modules. This we do in Section 2. There are known connections between twisting elements and
-matrices of quasitriangular Hopf algebras (see Section 2 for the details). In Section 3, we factor the
-matrix of the quasitriangular Hopf algebra
, where
and
are roots of unity (under a mild numerical constraint), and obtain a twisting element from one of the factors. In the special case of
, we show that this twisting element leads to a universal deformation formula, and via this alternate approach, essentially recover the example of [CGW] .
In Section 4, we continue under the assumption that
and
are roots of unity, but consider the infinite-dimensional quantum group
. We show that an analogue of an
-matrix for
gives rise to a universal deformation formula for certain types of algebras. In such formulas, it appears to be necessary to work with the quantum group
, or more particularly with
, rather than with its one-parameter quotient
, whose relations do not permit the types of actions we describe in Section 5.
For the definitions and general theory in Section 2, we assume only that our algebras are defined over a commutative ring
. For the quantum groups in Sections 3 through 5, we assume
is a field of characteristic 0 containing appropriate roots of unity.
2 Twisting elements and deformation formulas
Let
be an associative bialgebra over a commutative ring
with coproduct
and counit
. An associative
-algebra
is called a left
-module algebra if it is a left
-module such that
|
(2.1)
|
|
(2.2)
|
for all
and
. Here we have adopted the Heyneman-Sweedler convention for the coproduct,
.
Let
be a category of
-module algebras. An element
is a twisting element for
(based on
) if
|
(2.3)
|
|
(2.4)
|
as operators on the
-modules
and
, respectively, for all objects
in
, where
denotes the identity operator. More generally, we allow
to be a formal infinite sum of elements of
, provided that
and the operators on each side of equations ( 2.3 ) and ( 2.4 ) are well-defined operators on the appropriate objects.
The following theorem is essentially the same as [GZ,Thm. 1.3] , but stated in terms of the category
. We include a proof for completeness.
Theorem 2.5.
Let
be an associative algebra in a category
of
-module algebras having multiplication map
, and assume
is a twisting element for
. Let
denote the
-module
with multiplication map
. Then
is an associative algebra with multiplicative identity
.
-
Proof.
Equations ( 2.1 ) and ( 2.3 ) imply that
is the multiplicative identity in
. Associativity in
is equivalent to the identity
of functions from
to
. We prove this identity by applying the associativity of
, ( 2.4 ), and ( 2.2 ) twice:
| |
| |
| |
(This identity may be viewed as a commutative diagram as in [GZ,Thm. 1.3] .) □
Known examples of twisting elements include those arising from quasitriangular Hopf algebras, and we recall these ideas next, with a few details for clarity. A Hopf algebra
is quasitriangular if the antipode
is bijective, and there is an invertible element
such that
|
(2.6)
|
for all
, where
,
|
(2.7)
|
|
(2.8)
|
The notation is standard; for example, if
, then
.
It can be shown that
(see for example [M,Prop. 10.1.8] ). Thus left multiplication of ( 2.7 ) by
and left multiplication of ( 2.8 ) by
yields
which is very similar to ( 2.4 ), but applies to right
-module algebras. For left
-module algebras,
(which equals
[M,(10.1.10)] ) satisfies ( 2.4 ) (as noted in [GZ,pp. 139–40] ). The element
can also be shown to satisfy ( 2.4 ) by the following argument:
Interchange the factors in ( 2.7 ) and ( 2.8 ) to obtain
|
(2.9)
|
|
(2.10)
|
Using the fact that
is itself an
-matrix for
(that is, for
with the opposite coproduct), we see that
(This may also be shown directly.) Thus, multiplying ( 2.9 ) on the right by
and multiplying ( 2.10 ) on the right by
yields
|
(2.11)
|
Again, as
is an
-matrix for
, it satisfies ( 2.3 ) [M,(10.1.11)] . Therefore we have shown the following.
Proposition 2.12.
Let
be a quasitriangular Hopf algebra with R -matrix
. Then
is a twisting element for left
-module algebras.
Now assume
is an associative
-algebra with multiplication map
. Let
be an indeterminate, and extend the scalars for
by setting
.
Here
is the algebra of formal power series in
with coefficients in
. The algebra
is graded with homogeneous components
for each
. Tensor powers of
are also graded in the usual way. A map
is said to have
if
takes
to
for all
.
A formal deformation of
is an associative
-algebra structure on the
-module
for which multiplication takes the form
Each
is assumed to be a
-linear map, extended to be
-linear. Associativity imposes certain constraints on the maps
; for example,
must be a Hochschild two-cocycle:
for all
. For further discussion of the conditions on the maps
, see [GZ,p.141] . Let
be a
-bialgebra,
be a category of
-module algebras of the form
for
-algebras
, and
be a twisting element for
. We say that
is a universal deformation formula for
(based on
) if
|
(2.13)
|
for some elements
of
, which when viewed as transformations on
(for any
in
) satisfy
for each
.
In this case,
has a (new) associative algebra structure with multiplication
and multiplicative identity
by Theorem 2.5 . In addition, this multiplication has the following property: for all
,
Note that our definition of a universal deformation formula generalizes that of [GZ,Defn.1.13] , where
is required to have an expression
with each
, and where
is the category of
-module algebras arising from
-module algebras by extension of scalars. The differences are that here
is not required to satisfy ( 2.3 ) and ( 2.4 ) on module algebras outside of category
, and
is not required to take each algebra
in
to itself, but only to
.
We illustrate such actions in Section 5.
The next example is well-known.
Example 2.14.
Let
be a field of characteristic 0,
be a commutative
-bialgebra, and
be any category of
-module algebras of the form
where
is a
-module algebra. Let
be the space of primitive elements of
and
. Then
is a universal deformation formula for
(see [
GZ,Thm.2.1]
for a proof ).
The following theorem on universal deformation formulas generalizes a consequence of [GZ,Thm.1.3andDefn.1.13] .
Theorem 2.15.
Let
be a
-bialgebra and
be a category of
-module algebras of the form
, where
is a
-algebra. Let
be a universal deformation formula for
based on
. Then for each
in
,
defines a formal deformation of
.
-
Proof.
Let
. By Theorem 2.5 ,
provides an associative algebra structure on
. For
, we may write
where
,
are
-linear functions. The coefficient of
in
is
. As
, the coefficient of
in
is 0 for each
. Therefore we have
where
This is necessarily well-defined for each
, since
is assumed to be a well-defined operator on
. As
,
act linearly on
for each
, the same is true of the functions
,
, and thus
is bilinear. □
3 Twisting elements from finite quantum groups
Drinfel'd doubles of finite-dimensional Hopf algebras provide a wealth of examples of quasitriangular Hopf algebras. In particular, the two-parameter restricted quantum groups
of [BW3] are Drinfel'd doubles under some mild assumptions on
and
. We discuss implications of this for twisting and deforming algebras later on. The restricted quantum group
is a quotient of the unital associative
-algebra
, which we introduce next.
Let
denote an orthonormal basis of a Euclidean space
with an inner product
. Set
and
. Then
is a finite root system of type A
with
a base of simple roots. Let
be a field. Choose
with
. The algebra
is the unital associative
-algebra generated by
, and
(
), subject to the following relations.
-
(R1)
The
all commute with one another and
-
(R2)
and
-
(R3)
and
,
-
(R4)
-
(R5)
-
(R6)
-
(R7)
The algebra
becomes a Hopf algebra over
with
,
group-like and
| |
| |
| |
Let
be the group algebra generated by all
,
, and let
(respectively,
) be the subalgebra of
generated by all
(respectively,
).
We define
The algebra
has a triangular decomposition
, and as discussed in [BKL] , the subalgebras
,
respectively have monomial PBW (Poincaré-Birkhoff-Witt) bases
|
(3.1)
|
|
(3.2)
|
The spaces
and
are Hopf subalgebras of
.
We view
as a Hopf algebra with the opposite coproduct
. Then there is a Hopf pairing
given by
|
(3.3)
|
for
, and all other pairings between generators are 0. Pairings between more complicated expressions can be computed from those values by applying the coproducts
of
and
of
according to the following rules:
(3.4)
-
(i)
-
(ii)
-
(iii)
for all
and
.
For the rest of this section, we assume that
and
are roots of unity. Let
be a primitive
th root of unity,
be a primitive
th root of unity, and
be the least common multiple of
and
. We also assume that
contains a primitive
th root of unity
and that
,
, where
are nonnegative integers.
It is shown in [BW3] that all
,
,
, and
(
) are central in
. The ideal
generated by these elements is in fact a Hopf ideal [BW3,Thm.2.17] . This is evident from the following expressions for the coproduct given in [BW3,(2.24)andtheensuingtext] :
|
(3.5)
|
|
(3.6)
|
where
and
. As
is a Hopf ideal, the quotient
|
(3.7)
|
is a Hopf algebra, called the restricted two-parameter quantum group. Furthermore,
is finite-dimensional, as can be readily seen from the PBW-bases ( 3.1 ) and ( 3.2 ).
The algebra
is graded by the root lattice
of
by assigning
Since the generators of
are homogeneous,
inherits the grading. Note that by ( 3.3 ) and ( 3.4 ), if
and
with
, then
.
Let
denote the set of monomials in
having each
appear with exponent at most
. Identifying cosets in
with their representatives, we may assume
is a basis for the subalgebra of
generated by the elements
.
The following proposition will allow us to use the pairing ( 3.3 ) on the quotient algebra
.
Proposition 3.8.
The ideal
is contained in the radical of the pairing
of
. Thus there is an induced pairing on the quotient
.
-
Proof.
For the group-like elements, note that for each pair
,
| |
| |
as
and
are
th roots of 1. Thus
is in the radical of
, and similarly for
.
Now consider
and let
be any monomial in
of degree
. Such a monomial is the only type that potentially has a nonzero pairing with
. As
and
, we may write
for some
. By ( 3.4 )(iii) and ( 3.5 ),
Again as
and
, each term above is 0. □
We will need to apply the following results from [BW3] , where the corresponding pairing differs from ( 3.3 ) by nonzero scalar multiples. This difference does not affect the results. (See the proof of [BW3,Thm.4.8] where a relevant adjustment is made.)
Proposition 3.9.
[
BW3,(5.8)andLem. 4.1]
Let
be a primitive
th root of unity, and suppose
and
. Let
be the subalgebra of
generated by
and
be the subalgebra of
generated by
. Then the Hopf pairing
on
satisfying ( 3.3 ) and ( 3.4 ) is nondegenerate if
|
(3.10)
|
(that is, the first expression in the parentheses is relatively prime to
, the least common multiple of the orders of
and
as roots of 1).
We will use the definition of the Drinfel'd double
given in [BW3] :
as a coalgebra, where
denotes the dual Hopf algebra with the opposite coproduct. As an algebra,
and
become subalgebras of
under identifications with
and
, respectively, and
Proposition 3.11.
[
BW3,Thm.4.8]
Assume
and
, where
is a primitive
th root of unity, and suppose that ( 3.10 ) holds. Then there is an isomorphism of Hopf algebras
, where
is the Drinfel'd double of the Hopf subalgebra
of
generated by
.
In general, whenever
, then
is quasitriangular with
-matrix,
where
runs over a basis of
and
runs over the dual basis of the dual space
, which can be identified with the Hopf subalgebra
of
generated by
(
) but with the opposite coproduct
[BW3,Lem.4.1] .
To illustrate this result in a very special case, take
, a primitive
th root of unity, and
. Then
is isomorphic to the Drinfel'd double of
when
and
are relatively prime. The quotient of
by the ideal generated by the elements
for
is a one-parameter restricted quantum group related to those which have played a significant role in the study of algebraic groups in the case
, a prime (see for example, [AJS] ).
Lemma 3.12.
Assume
is a Drinfel'd double as in Proposition 3.11 . Then its
-matrix factors,
|
(3.13)
|
with
, where
runs over the basis
of
-power truncated monomials, and
, where
runs over the basis
for all
of the group algebra generated by the
.
-
Proof.
Let
denote elements of
dual to the vectors of
with respect to this pairing so that
for all
. For example, if
, then
is a linear combination of
and
. Similarly, let
denote the basis of
dual to
, where
is the group generated by the
. By the triangular decomposition of
, the elements
form a basis of
(see for example, [BW3,(2.16)] ). Moreover using
and
, the properties of the pairing in ( 3.4 ), and degree considerations, we have
| |
| |
| |
| |
| |
This calculation shows that
is the dual basis element of
relative to the pairing. Thus,
| |
| |
□
It follows from Lemma 3.12 that
, where
=
and
. Equation ( 2.11 ) holds for
, that is
|
(3.14)
|
We will show that the factor
alone satisfies this equation and so is a twisting element.
Let
be the category of left
-module algebras. Let
be the group generated by the
and
the group generated by the
(
). Since
generate the finite abelian group
of order a power of
, and
contains a primitive
th root of 1, each
decomposes into a direct sum of common eigenspaces (or weight spaces) for the
. That is,
for some group homomorphisms
, where
and
for all
,
. We will use this decomposition to show that
is itself a twisting element for
.
Theorem 3.15.
Let
. The factor
of the opposite
of the
-matrix for
is a twisting element for the category of left
-module algebras.
-
Proof.
Let
be a left
-module algebra, and apply both sides of ( 3.14 ) to
, for weights
of
. Note that
acts as the scalar
on
. Applying the left side of ( 3.14 ) to
, we have
| |
| |
since each term of
when applied to
changes the weight of
and the weight of
by multiplication by functions which are inverses of one another, with a net effect of no change at all in the weight. Similarly, applying the right side of ( 3.14 ) to
, we have
| |
| |
Now
is itself an
-matrix for the group algebra
, because our constructions show that this group algebra is the Drinfel'd double of
. Similar calculations to those above prove that
for all weights
. (Note this implies that
is a 2-cocycle for the dual group to
.) Moreover, as
is invertible, none of these values is zero. Thus we may cancel the factors
and
from the above expressions and apply ( 3.14 ) to obtain
as desired. We also have
by the definition of
. □
Examples 3.16.
We will give some examples of
-module algebras, which may be twisted by
according to Theorems 2.5 and 3.15 . First let
be the natural
-dimensional module for
; that is,
has a basis
and
| |
| |
| |
| |
where
. Because each
acts as 0,
becomes a
-module via the induced action. This action extends to a representation of
on the tensor algebra
, where the action on
is by
for
. When
, then
, and the action is given by the counit. By definition then,
is a
-module algebra.
Let
be the (two-sided) ideal of
generated by all elements of the form
Then
is homogeneous,
, where
. In fact,
is a
-submodule of
: A computation such as in [
BW2,Prop.5.3]
shows that
is a
-submodule of
, and as
, the result follows. Thus, the quantum plane
, with
if
, is isomorphic to the
-module algebra
, under the identification
.
Let
Then the
-module action is given by
| |
| |
| |
| |
where
Other quotients of
provide examples as well: Choose a positive integer
, and let
be the ideal of
generated by the
-submodule of
having as generators all
. As an abbreviation, we will write
, and similarly leave out the tensor product symbol in the notation for words in
. For example, if
and
, the
-submodule of
generated by
and
has basis
. If
and
, the
-submodule of
generated by
and
has basis
.
If
is a primitive third root of 1, and
,
, then
acts on the quotient of
by the ideal generated by
,
, which is the down-up algebra
in the notation of [
BR]
. This example is related to one appearing in the work of Montgomery and Schneider [
MS]
on skew derivations and actions of the double of the Taft algebra.
The family of down-up algebras
(where
is an arbitrary primitive
th root of unity) is especially interesting, as the finite-dimensional modules for these algebras are completely reducible [
CM]
. This family includes the universal enveloping algebra
for
, the universal enveloping algebra
of the Lie superalgebra
for
, and the algebras appearing in this example for
a primitive third root of 1.
Example 3.17.
Here we generalize the last example by allowing
to be arbitrary. Again we suppose
,
for
a primitive third root of 1, but take
for any
. Consider the subspace
of
spanned by the following elements:
| |
| |
| |
| |
From the formulas below it is easy to see that
is a
-submodule of
:
| |
| |
for all
; and using the fact that
, we have
| |
| |
| |
| |
| |
| |
| |
| |
The restricted quantum group
acts on the quotient algebra
obtained by factoring out the ideal generated by
. In the quotient, each pair
(
) generates a down-up algebra
.
Example 3.18.
As a special instance of Theorem 3.15 , consider the algebra
for which
(Note we omit subscripts when discussing
, as there is only one of each type of generator.) We know from Theorem 3.15 that this is a twisting element, but this also may be verified directly quite easily. This quantum group has as a quotient
and in
, the images of
and
commute by relation (R4). Thus the defining relations of
are all homogeneous with respect to powers of
.
Consequently, the following is a twisting element based on
, as may be checked directly:
Thus
is a universal deformation formula for the category of
-module algebras arising from
-module algebras by extension of scalars. In fact, there is a one-parameter family of such deformation formulas, given by
, for any
. The choice
yields precisely the universal deformation formula of [
CGW,Lem. 6.2]
that was applied to a certain
-module algebra
(given by a crossed product of a polynomial ring with a finite group) to obtain a formal deformation of
.
However, it is an accident due to the choices
,
for
that this technique produces a universal deformation formula directly, as relation (R4) of
is not homogeneous with respect to the powers of the
(nor of the
), and in general, there is no reasonable quotient in which it becomes so. In the next section we will remedy this situation by returning to the infinite-dimensional Hopf algebra
and choosing an action of
that itself incorporates the indeterminate
. Using this infinite-dimensional Hopf algebra necessitates a more complicated (but related) construction of a twisting element.
4 Twisting elements from infinite quantum groups
We consider the infinite-dimensional quantum group
, defined in the previous section, where
and
are roots of unity. We adopt the same notation as before, so that
is the least common multiple of the orders of
and
as roots of
, and
is the root lattice of
. We will also assume ( 3.10 ) holds, so that we may use Proposition 3.9 . Let
be a commutative
-algebra, and extend coefficients for
to
:
. (For the applications in the next section,
will be an extension of the field
by powers of an indeterminate.) As in [BW1] , we will be interested in weight modules, this time for
. We will consider only
-modules that are free as
-modules. For every
, define the corresponding algebra homomorphism
by
|
(4.1)
|
which may be extended to yield an algebra homomorphism from
to
. As
and
are roots of unity, the set
is finite.
Let
be a
-module, assumed to be free as a
-module. If as a
-module,
decomposes into the direct sum of eigenspaces
for algebra homomorphisms
, we say
acts semisimply on
. The homomorphisms
such that
are called the weights of
. Note that
|
(4.2)
|
by (R2), (R3), and ( 4.1 ). It follows that if
is a simple
-module with a nonzero weight space
, then
acts semisimply on
, since the (necessarily direct) sum of the weight spaces is a submodule. In fact, if
, any cyclic
-module generated by a weight vector
, then
acts semisimply on
, and all the weights of
are of the form
(
).
Let
be the category of unital
-modules
that are free as
-modules and satisfy the following conditions:
-
(
)
acts semisimply on
;
-
(
)
For each
(
), both
and
annihilate
.
Note that
is closed under direct sums and quotients, and it follows from ( 3.5 ) and ( 3.6 ) that
is closed under tensor products.
Examples of modules
satisfying (
2) are the
-modules (see the previous section) with scalars extended to
, which can be viewed as
-modules.
However we will also be interested in some modules in category
that do not correspond to
-modules, where neither
nor
acts as the identity. We remark that (
2) has been included so that certain operators will be well-defined. It is not necessarily true that
act nilpotently on all finite-dimensional modules as happens in the non-root of unity case [BW1,Cor.3.14] . The argument used there fails, as
is finite.
For any two modules
and
in
, we will construct a
-module homomorphism
by Jantzen's method (see [J,Ch. 7] ). The main difference between the two-parameter version of Jantzen's method (see [BW1,Sec. 4] ) and what we are about to do here is that now we are assuming both
and
are roots of unity, a case excluded from consideration in [J, BW1] . As a consequence, we will not know whether
is invertible. However we will make the necessary adjustments to show that most of the arguments of [BW1,Secs. 4,5] , in particular those needed to obtain a twisting element, apply in this context.
The desired function
will be the composition of three
-linear functions
,
, and
, where
is the map that interchanges tensor factors as before, and
and
are as follows. (Ultimately we will show that
is a twisting element, accounting for the choice of notation.) As
satisfies (
1), by ( 4.2 ) there are algebra homomorphisms
such that
, where
is a
-submodule having weights contained in
. (For example, we may take the sum over a set of representatives
of cosets of
in
.) For the purposes of this section, it will suffice to deal with each summand
separately. Assume
and
are such modules. Let
and
,
. Set
where
is given by
with
,
,
,
, and the Hopf pairing is as in ( 3.3 ), ( 3.4 ). Thus
is the function
of [BW1,(4.2)] as restricted to the root lattice
, and it can be shown that
generalizes the function
there by looking at cosets of the weight lattice modulo
. In particular, the following hold:
|
(4.3)
|
The definition of
is similar to that of
in [BW1,Sec.4] . We will construct
as a sum of elements of
, which then may also be considered elements of
. The subalgebra
of
generated by
and
(
) may be decomposed as
where
, and
.
For each
, let
be the linear span of all PBW basis elements in
(see ( 3.1 )) in which the power of each
is less than
. Similarly define
. By Propositions 3.8 and 3.9 , the spaces
,
are nondegenerately paired under the assumption that ( 3.10 ) holds. We define
where
,
,
is a basis for
, and
the dual basis for
. Note that if
, then
, and if
, we will also set
for convenience.
As
, for all
we have
where
means
. For each
, there are elements
and
such that
| |
| |
where
(respectively,
) is a sum of terms involving products of more than one
in the second factor (respectively, in the first factor). Similarly, if
, we define
and
by
| |
| |
The following identities from [BW1] hold in our context, as the proof consists of calculations that do not use properties of the parameters
and
.
Lemma 4.5.
[
BW1,Lem.4.6,Lem.4.8]
For all
,
, and
, the following hold:
-
(i)
.
-
(ii)
.
-
(iii)
.
For all
,
, and
, the following hold:
-
(iv)
.
-
(v)
.
-
(vi)
.
Since the spaces
and
are nondegenerately paired with dual bases
and
, for each
and
, we have
|
(4.6)
|
The next lemma is a modified version of [BW1,Lem.4.10] or [J,Lem.7.1] . The main distinction is that here the identities are as operators on modules rather than as elements of
.
Lemma 4.7.
Let
and
. Then the following relations hold for operators on tensor products of pairs of modules in category
:
-
(i)
and
.
-
(ii)
.
-
(iii)
.
-
Proof.
We will verify that (ii) holds. Identity (iii) is similar to (ii), and (i) is immediate from (R2) and (R3). In the calculations below, we use Lemma 4.5 (iv)–(vi), ( 4.6 ), and the fact that on a module in category
any element of
acts as its projection in
.
| |
| |
| |
| |
| |
| |
| |
| |
| |
| |
□
Notice that (ii) and (iii) of the above lemma hold even in the cases where
but
, because in these cases both sides of each equation annihilate modules in category
.
The following is a modification of [BW1,Thm.4.11] or [J,Thm.7.3] .
Theorem 4.8.
Let
and
be modules in category
. Then the map
is a homomorphism of
-modules.
-
Proof.
We must prove that the action of each generator of
commutes with the map
. By Lemma 4.7 ,
and
commute with
. We will check this for
and leave the similar calculation for
as an exercise. We may assume
and
for some algebra homomorphisms
. Let
,
. By ( 4.3 ),
| |
| |
| |
| |
| |
Now we may replace
by
or
,
by
, and
by
. Thus we obtain the following expression to which we apply Lemma 4.7 (ii):
| |
| |
| |
□
We will need the following relation [BW1,(5.1)] :
|
(4.9)
|
for all
and
. A similar identity was proven and used in our Lemma 3.12 .
Lemma 4.10.
If
and
, then the following hold as relations of operators on a tensor product of two modules from category
:
-
(i)
,
-
(ii)
.
-
Proof.
As in [BW1,Lem.5.2] , we may write
for some scalars
, this time as an identity of operators on a tensor product of modules from category
. For all
, and
, by ( 3.4 ) and ( 4.9 ), we have
The proof of (ii) is similar. □
Now let
,
, similarly
, and
, where
denotes
applied to the
tensor slots. Recall the notation
for
introduced in Section 2, where the notation
was reserved for
.
Lemma 4.11.
The following holds as an identity of operators on a tensor product of three modules from category
:
.
-
Proof.
Let
and assume
,
, and
. Let
,
,
. The left side of (i) applied to
is
| |
| |
| |
| |
| |
On the other hand, by ( 4.3 ),
| |
| |
| |
Re-summing over
, and replacing
by
, we see that (i) holds. □
The next result is the quantum Yang-Baxter equation for
.
Theorem 4.12.
Let
. Then
as maps from
to
.
-
Proof.
If
is a permutation of
, let
denote the corresponding permutation of three tensor factors, that is
.
In particular, if
equals the transposition
, we write simply
, as
is just the same as
applied to tensor slots
and
in that case.
Note that
for all
, and that
by two applications of ( 4.3 ). We apply these identities, Theorem 4.8 , and Lemma 4.11 to obtain the following:
| |
| |
| |
| |
| |
| |
| |
| |
| |
□
We will need one more lemma in order to obtain the hexagon identities, from which it will follow that
is a twisting element for
.
Lemma 4.13.
The following are identities of operators on a tensor product of three modules from
:
-
(i)
.
-
(ii)
.
-
(iii)
-
(iv)
-
Proof.
We will check (ii) and (iv); the proofs of (i) and (iii) are similar. By Lemma 4.10 and ( 4.6 ), considering operators on modules we have
| |
| |
| |
| |
which proves (ii).
Let
and
,
,
. By ( 4.3 ),
| |
| |
| |
| |
which proves (iv). □
Next we will prove the hexagon identities.
Theorem 4.14.
Let
. Then the following are identities of maps from
to
(respectively,
):
-
(i)
.
-
(ii)
.
-
Proof.
We will prove (ii). The proof of (i) is similar. Let
. By Lemma 4.13 , we have
| |
| |
| |
| |
| |
| |
□
Finally we show that
is a twisting element for
.
Theorem 4.15.
Let
be modules in category
. Then for
as defined in ( 4.4 ) we have
as operators on
. Thus
is a twisting element for any subcategory of
consisting of
-module algebras.
-
Proof.
In Theorem 4.14 , multiply (i) by
on the right, multiply (ii) by
on the right, and apply Theorem 4.12 to obtain the identity
as functions from
to
, for any
. It may be checked that this is equivalent to the identity stated in the theorem, using ( 4.3 ). Thus ( 2.4 ) holds, and ( 2.3 ) is immediate from the definition of
. □
We remark that all of the above arguments apply equally well to
, if
is not a root of unity and our
-module algebras are (possibly infinite) direct sums of modules arising from the category
modules of [BW1,Sec. 4] by shifting weights by some functions
. The function
there merely needs to be replaced by our more general function
here. The element
from that paper provides another example of a twisting element similar to the well-known examples arising from
-matrices of one-parameter quantum groups.
5 Deformation formulas from infinite quantum groups
Let
, and assume throughout this section that
and
are both roots of unity with
satisfying condition ( 3.10 ). Let
, Laurent polynomials in
with finitely many negative powers of
. Let
be any finite-dimensional
-module such that the extension
is a
-module in category
. Let
be the tensor algebra of
over
, considered as a
-module algebra. As
is closed under tensor products and infinite direct sums,
is also in category
. Therefore Theorem 4.15 may be applied to twist the multiplication of
.
We wish instead to go further and produce a deformation formula that may be applied to quotients of
to obtain formal deformations. We define a new action of
on
denoted
, that is
-linear, by specifying
| |
| |
for all
, and extending to
via the
-module algebra conditions ( 2.1 ), ( 2.2 ). (We are just forcing the generators of
to act as scalar multiples of their original actions, the scalars being powers of
.) The powers of
arising in the new action
are chosen so that the defining relations of
are preserved.
One way to obtain this module structure is to let
act on the tensor product
via coproducts, making
a weight module with the single weight
so that
,
annihilate
(resulting in the above action on the tensor product
via their coproducts). Observe that this new
-module structure on
yields a module in category
, as the only change is in the weights
.
Theorem 5.1.
The operator
defined in ( 4.4 ) is a universal deformation formula for any subcategory of
(over
) consisting of
-module algebras for which ( 2.13 ) holds.
-
Proof.
This follows immediately from Theorem 4.15 and the definition of a universal deformation formula for a category. □
Of course, the
-module algebras
constructed above satisfy the hypotheses of the theorem, so by Theorem 2.15 ,
defines a formal deformation of
.
If
is any finite-dimensional vector space, the Hochschild cohomology groups
are
for
(e.g. see [W,Prop.9.1.6] ). Thus all formal deformations of
are equivalent to the trivial one, and so we are interested in applying Theorem 5.1 to proper quotients. For example, we may take the truncated tensor algebra obtained by letting
be a fixed positive integer and taking the quotient by the ideal
. The second Hochschild cohomology group of a truncated tensor algebra is nontrivial. (See for example [C,Lem.4.1] , where we need to identify our
with Cibils'
,
the quiver with one vertex and
loops, and
the ideal of the path algebra
generated by the loops.)
Example 5.2.
Next we consider some examples that are group crossed products.
Let
be the natural
-dimensional module for
, defined in Section 3. We will specify an action of
on a crossed product
, where
is the abelian group
on generators
, written multiplicatively.
Choose arbitrary
th roots of unity
(
), and let an action of
on
as automorphisms be defined by
for
and
. The smash product
is the vector space
with the multiplication
for
and
. Define the following action of
on the smash product
:
| |
| |
| |
| |
where
acts trivially on
. (It is possible to modify this example to involve nontrivial actions of the
,
on
.) It may be checked that
is a
-module algebra. Note that
is in category
, which may be seen by decomposing
as an algebra into a direct sum of copies of
(as
is a finite abelian group), and partitioning the elements of each
accordingly, (or simply by letting
be the ring of coefficients). If elements of
are assigned degree 0, it may be verified that
satisfies equation ( 2.13 ), and so yields a formal deformation of
. This example is similar in some respects to that given in [
CGW]
, although in that example it was possible to take a further quotient of
, namely a polynomial algebra, because of the special nature of the parameters.
Just as for
, formal deformations of
are necessarily infinitesimally trivial, as the second Hochschild cohomology group is again trivial. However, we may again truncate the tensor algebra and consider the resulting crossed product with
as a
-module algebra and corresponding deformations.
References
-
H.H. Andersen, J.C. Jantzen, and W. Soergel, Representations of quantum groups at a
th root of unity and of semisimple groups in characteristic
: Independence of
, Astérisque 220 Société Math. de France 1994.
-
G. Benkart, S.-J. Kang, and K.-H. Lee, On the center of two-parameter quantum groups, submitted.
-
G. Benkart and T. Roby, Down-up algebras, J. Algebra, 209 (1998), 305–344; Addendum 213 (1999), 378.
-
G. Benkart and S. Witherspoon, Two-parameter quantum groups and Drinfel'd doubles, Algebr. Represent. Theory, 7 (2004), 261–286.
-
G. Benkart and S. Witherspoon, Representations of two-parameter quantum groups and Schur-Weyl duality, “Hopf Algebras: Proceedings from an International Conference held at DePaul University”, Bergen, Catoiu, and Chin, eds., 65–92, Lecture Notes in Pure and Appl. Math. 237 Dekker, New York, 2004.
-
G. Benkart and S. Witherspoon, Restricted two-parameter quantum groups, “Representations of Finite Dimensional Algebras and Related Topics in Lie Theory and Geometry”, 293–318. Fields Inst. Commun. 40, Amer. Math. Soc., Providence, RI, 2004.
-
A. Căldăraru, A. Giaquinto, and S. Witherspoon, Algebraic deformations arising from orbifolds with discrete torsion, J. Pure Appl. Algebra 187 (2004), 51–70.
-
P.A.A.B. Carvalho and I.M. Musson, Down-up algebras and their representation theory, J. Algebra 228 (2000), 286–310.
-
C. Cibils, Rigidity of truncated quiver algebras, Adv. Math. 79 (1990), no. 1, 18–42.
-
A. Giaquinto and J. Zhang, Bialgebra actions, twists, and universal deformation formulas, J. Pure Appl. Algebra 128 (1998), 133–151.
-
C. Grunspan, Quantizations of the Witt algebra and of simple Lie algebras in characteristic
, J. Algebra 280 (2004), 145–161.
-
J. C. Jantzen, Lectures on Quantum Groups, vol. 6, Graduate Studies in Math., Amer. Math. Soc., Providence, 1996.
-
S. Montgomery, Hopf Algebras and Their Actions on Rings, CBMS Conf. Math. Publ., vol. 82, Amer. Math. Soc., Providence, 1993.
-
S. Montgomery and H.J. Schneider, Skew derivations of finite-dimensional algebras and actions of the double of the Taft Hopf algebra, Tsukuba J. Math. 25 (2001), no. 2, 337–358.
-
C. A. Weibel, An Introduction to Homological Algebra, Cambridge University Press, 1994.
-
S. Witherspoon, Skew derivations, Hopf algebras and deformations of group crossed products, submitted (http://www.math.tamu.edu/~sjw/pub/skew.pdf).
Department of Mathematics, University of Wisconsin, Madison,Wisconsin 53706, USA E-mail address : benkart@math.wisc.edu Department of Mathematics, Texas A&M University, College Station,Texas 77843, USA E-mail address : sjw@math.tamu.edu