Schur–Weyl duality, Verma modules, and row quotients of Ariki–Koike algebras
Abstract.
We prove a Schur–Weyl duality between the quantum enveloping algebra of and certain quotient algebras of Ariki–Koike algebras, which we describe explicitly. This duality involves several algebraically independent parameters and the module underlying it is a tensor product of a parabolic universal Verma module and a tensor power of the standard representation of . We also give a new presentation by generators and relations of the generalized blob algebras of Martin and Woodcock as well as an interpretation in terms of Schur–Weyl duality by showing they occur as a special case of our algebras.
1. Introduction
Schur–Weyl duality is a celebrated theorem connecting the finite-dimensional modules over the general linear and the symmetric groups. It states that, over a field that is algebraically closed, the actions of and on commute and form double centralizers. Several variants of (quantum) Schur–Weyl duality are known, see for example [4, 6, 5, 9, 16, 24] for such variants related to our paper. One particular family of generalizations of interest for us uses a module akin to the one appearing in Schur–Weyl duality, but with an infinite-dimensional module instead of . For example, in [15] it is established a Schur–Weyl duality between and the blob algebra of Martin and Saleur [19] with the underlying module being a tensor product of a projective Verma module with several copies of the standard representation of . We should warn the reader that in [15] the blob algebra was called the Temperley–Lieb algebra of type (see [18] for further explanations).
1.1. In this paper
We consider the tensor product of a parabolic universal Verma module with the -folded tensor product of the standard representation for to establish a Schur–Weyl duality with a quotient of Ariki–Koike algebras. Ariki–Koike algebras were first considered by Cherednik in [10] as a cyclotomic quotient of the affine Hecke algebra of type . These algebras were later rediscovered and studied by Ariki and Koike [3] from a representation theoretic point of view. Independently, Broué and Malle attached in [7] a Hecke algebra to certain complex reflection groups, and Ariki–Koike algebras turn out to be the Hecke algebras associated to the complex reflection groups .
Recall that the Ariki–Koike algebra with parameters and is the -algebra with generators , where generate a finite-dimensional Hecke algebra of type and satisfies , for , and . We consider the semisimple case, where the simple modules of are indexed by -partitions of .
Let be a -tuple of positive integers and be the set of all -partitions of such that for all .
In this paper we introduce the row-quotient algebra , that depends on as the quotient of by the kernel of the surjection
Let be a parabolic Verma module and the standard representation for . In our conventions, is standard and has Levi factor , with and and depends on algebraically independent parameters (see Section 3.2 for more details). Thanks to the braided structure on the category of integrable modules over , we define a left action of on in Section 4. Our main result is:
Theorem A (4.2 and 4.1).
-
•
The action of and of on commute with each other, which endow with a structure of -module.
-
•
The algebra morphism is surjective and factors through an isomorphism
(1) - •
The isomorphism in Equation (1) has several particular specializations (Corollaries 4.3-4.7), some of them recovering well-known algebras:
-
•
If and , then is isomorphic to the Hecke algebra of type .
-
•
If and , then is isomorphic to the Temperley–Lieb algebra of type .
-
•
For such that and for all , then is isomorphic to the Ariki–Koike algebra .
-
•
If is such that and , then is isomorphic to the Hecke algebra of type with unequal and algebraically independent parameters (see [14, Example 5.2.2, (c)]).
- •
In the last case, this gives a new interpretation of the generalized blob algebras in terms of Schur–Weyl duality. We also give a new presentation of as a quotient of Ariki–Koike algebras:
Theorem B (2.15).
Suppose that is semisimple and that for every we have . The generalized blob algebra is isomorphic to the quotient of by the two-sided ideal generated by the element
1.2. Connection to other works
The idea of writing this note originated when we started thinking of possible extensions of our work in [18] to more general Kac–Moody algebras and were not able to find the appropriate generalizations of [15] in the literature. When we were finishing writing this note Peng Shan informed us about [23], whose results are far beyond the ambitions of this article. Nevertheless, we expect our results to be connected to [23, §8] using a braided equivalence of categories between a category of modules for the quantum group and a category of modules over the affine Lie algebra , which is due to Kazhdan and Lusztig [17]. However, the explicit description of the endomorphism algebra of , which was our first motivation towards categorification later on, does not seem to appear anywhere in [23] except in the particular case of our 4.5.
Acknowledgments
We would like to thank Steen Ryom-Hansen for comments on an earlier version of this paper. The authors would also like to thank the referee for his/her numerous, detailed, and helpful comments. The authors were supported by the Fonds de la Recherche Scientifique - FNRS under Grant no. MIS-F.4536.19.
2. Ariki–Koike algebras, row quotients and generalized blob algebras
We recall the definition of Ariki–Koike algebras and define some quotients which will appear as endomorphism algebras of modules over a quantum group. As a particular case we recover the generalized blob algebras of Martin and Woodcock [20] and we obtain a presentation of these blob algebras that seems to be new.
2.1. Reminders on Ariki–Koike algebras
Fix once and for all a field and two positive integers and and choose elements and . We recall the definition of the Ariki–Koike algebra introduced in [3], which we view as a quotient of the group algebra of the Artin–Tits braid group of type .
Definition 2.1.
The Ariki–Koike algebra with parameters and is the -algebra with generators , the relation
the cyclotomic relation
and the braid relations
Remark 2.2.
We use different conventions than [3]. In order to recover their definition, one should replace by , by , and by .
As in the type Hecke algebra, for any we can define unambiguously by choosing any reduced expression of .
It is shown in [3] that the algebra is of dimension and a basis is given in terms of Jucys–Murphy elements, which are recursively defined by and .
Theorem 2.3 ([3, Theorem 3.10, Theorem 3.20]).
A basis of is given by the set
Moreover, the center of is generated by the symmetric polynomials in .
We end this section with a semisimplicity criterion due to Ariki [2], which in our conventions takes the following form.
Theorem 2.4 ([2, Main Theorem]).
The algebra is semisimple if and only if
2.2. Modules over Ariki–Koike algebras
In this section, we suppose that the algebra is semisimple. In [3], Ariki and Koike gave a construction of the simple -modules, using the combinatorics of multipartitions.
2.2.1. -partitions and the Young lattice
A partition of of length is a non-increasing sequence of integers summing to . A -partition of is a -tuple of partitions such that . Given a -partition its Young diagram is the set
whose elements are called boxes. We usually represents a Young diagram as a -tuple of sequences of left-aligned boxes, with boxes in the -th row of the -th component.
Example 2.5.
The Young diagram of the -partition of is
A box of is said to be removable if is the Young diagram of a -partition , and in this case the box is said to be addable to .
Example 2.6.
The removable boxes of the -partition below are depicted with a cross
With respect to the above the definitions, we will also use the evident notions of adding a box to a Young diagram or removing a box from a Young diagram.
We consider the Young lattice for -partitions and some sublattices. It is a graph with vertices consisting of -partitions of any integers, and there is an edge between two -partitions if and only if one can be obtained from the other by adding or removing a box.
Example 2.7.
The beginning of the Young lattice for -partitions is the following:
If we fix , we then define as the set of -partitions such that . We will also consider the corresponding sublattice of the Young lattice.
Example 2.8.
For and , the beginning of the Young lattice for -partitions with and is the following:
We end this subsection with the notion of a standard tableau of shape where is a -partition of . Such a standard tableau is a bijection such that for all boxes and we have if and or and . Giving a standard tableau of shape is equivalent to giving a path in the Young lattice from the empty -partition to the -partition .
Example 2.9.
The standard tableau
of shape correspond to the path
2.2.2. Constructing the simple modules
We present the construction of simple modules of the Ariki–Koike algebra following [3, Section 3]. This construction is similar to the classical construction of simple modules of the symmetric group, the Hecke algebra of type or of the complex reflection group . This construction describes explicitly the action of the Ariki–Koike algebra on a vector space. For a -multipartition of , we set
where the sum is over all the standard tableaux of shape . Ariki and Koike gave an explicit action of the generators on the basis of given by the standard tableaux. The action of is diagonal with respect to this basis:
where is such that . The action of is more involved and depends on the relative positions of the numbers and in the tableau :
-
(1)
if and are in the same row of the standard tableau , then ,
-
(2)
if and are in the same column of the standard tableau , then ,
-
(3)
if and neither appear in the same row nor the same column of the standard tableau , then will act on the two dimensional subspace generated by and , where is the standard tableau obtained from by permuting the entries and . The explicit matrix is given in [3] and we will not need it.
Proposition 2.10 ([3, Theorem 3.7]).
If is any -multipartition of , the space is a well-defined -module and it is absolutely simple. A set of isomorphism classes of simple -modules is moreover given by , for running over the set of -partitions of .
The action of the Jucys–Murphy elements is also diagonal in the basis of standard tableaux:
(2) |
where . A useful consequence of 2.10 is the following: if is a simple -module and is a common eigenvector for with eigenvalues as in (2) for some standard tableau of shape , then is isomorphic to .
From the explicit description of the modules , using the standard inclusion , it is easy to see that for any -partition of we have
where the sum is over all -partition of whose Young diagram is obtained by deleting one removable box from the Young diagram of . The branching rule of the inclusions is therefore governed by the Young lattice of -partitions.
2.3. Row quotients of and generalized blob algebras
We now define the row quotients of which will appear later as endomorphism algebras of a tensor product of modules for .
Definition 2.11.
Let and recall that the algebra is assumed to be semisimple, which implies that , the product being over all -partitions of . Recall also that is the set of -partitions of with -th component of length at most .
The -row quotient of , denoted , is the quotient of by the kernel of the surjection
Remark 2.12.
If for all then .
Similar to the case of , we have inclusions and the branching rule is governed by the corresponding truncation of the Young lattice of -partitions.
2.3.1. Generalized blob algebras
In the particular case where for all , we recover the definition of the generalized blob algebras [20, Equation (14)], which we denote by . Under a mild hypothesis on the parameters, we give a presentation of .
We consider the following element of :
This element may look cumbersome, but can be better understood thanks to the following lemma:
Lemma 2.13.
The two-sided ideal of generated by is equal to the two-sided ideal generated by
Proof.
A simple computation in shows that
We therefore conclude using the fact that and that commutes with . ∎
We now investigate which -modules factor through the quotient by the two-sided ideal generated by .
Proposition 2.14.
The element acts by zero on if and only if for every such that for all .
Proof.
Suppose that and are such that with for all . Then there exist a tableau of shape such that and are in the first two columns of the -th component of the Young diagram of . By definition of , the generator acts on by multiplication by . The Jucys–Murphy element acts on by multiplication by whereas the Jucys–Murphy element acts on by multiplication by . Therefore, thanks to 2.13, does not act by zero on .
It remains to check that acts by zero on with whenever for all . Let be a standard tableau of shape . If and are in the same component of the tableau , then either and are in the same row and acts on by multiplication by , either and are in the same column and acts on by multiplication by . The second case is possible only if there exists such that and then acts by zero. If and are in two different Young diagrams and acts on by , where (resp. ) is such that (resp ). In both cases, acts by zero. ∎
Theorem 2.15.
Suppose that is semisimple and that for every we have . The generalized blob algebra is isomorphic to the quotient of by the two-sided ideal generated by .
3. Quantum , parabolic Verma modules and tensor products
We recall the definition of the quantum enveloping algebra of , and we also recall some basic properties of its modules, e.g. concerning parabolic Verma modules.
3.1. The quantum enveloping algebra of
Let be an indeterminate. The following definition of is over the field , but, via scalar extension, we will also consider it over a field containing without further notice.
Definition 3.1.
The quantum enveloping algebra is the -algebra with generators and , for and with the following relations:
and the quantum Serre relations
We endow it with a structure of a Hopf algebra, with comultiplication , counit and antipode given on generators by the following:
Set as the subalgebra generated by , and as the subalgebra generated by .
We denote by the weight lattice of with -basis given by the fundamental weights where . We denote by the root lattice with -basis given by the simple roots where . Denote by the set of positive roots, by the set of dominant weights for , that is with . We also endow with the standard non-degenerate bilinear form: . The symmetric group acts on by permuting the coordinates and leaves the bilinear form invariant. Finally, let be the half-sum of the positive roots.
We will often work with extensions , where the ’s are indeterminates and we also extend the bilinear form to .
3.2. Weights and parabolic Verma modules
Suppose that our field contains the field and let be an -module over the ground field . An element is said to be a weight vector if , where is the corresponding weight. The module is said to be a weight module if the action of the elements is simultaneously diagonalizable. A highest weight module is a weight module such that , where is a weight vector such that for .
It is well-known that finite dimensional weight -modules of type are parameterized by the set of dominant weights.
In this paper, we will be interested in modules over the field , where is an indeterminate (recall that is formal and so is also formal). Moreover, we only consider type modules, where the weights are of the form
for some and for all .
We now turn to parabolic Verma modules. Let be a standard parabolic subalgebra of with Levi factor , where and . Denote by the set , where , so that is generated by and for and and is generated by and for , and . Denote by the set of dominant weights for . We identify the set with the dominant weights of by the following map
For a dominant weight , we have an simple integrable finite dimensional -module of highest weight
Indeed, one can check that for any . We turn this -module into a -module by setting for all . Then the parabolic Verma module is
It is a highest weight module of highest weight . If , then we will simply denote this module by and its highest weight by .
Lemma 3.2.
For any , the parabolic Verma module is simple.
Proof.
Since for any the scalar product is not an integer, as one easily checks, the claim follows.∎
Remark 3.3.
If the parabolic subalgebra is the Borel subalgebra of upper triangular matrices, we have and the parabolic Verma module is the universal Verma module. The adjective universal means that any parabolic Verma module can be obtained from by specialization of the parameters.
In the rest of this article, all dominant weights will satisfy for all , and it will be convenient to identify such a weight with the corresponding -partition in . We will use the same notation to denote the -partition or the corresponding dominant weight.
We also denote by the standard representation of of dimension . Explicitly, this is a highest weight module with highest weight , it has as a basis and the action of is given by
3.3. Tensor products and branching rule
As is a Hopf algebra, its category of modules can be endowed with a tensor product. Explicitly, given and two modules over a ground ring , the action of the generators on is given using the comultiplication: for all and , one have
We will write instead of to simplify the notations. Since we will be interested in the endomorphism algebra of , we start by understanding the decomposition of this module.
Proposition 3.4.
For any , there is an isomorphism of -modules
where the sum is over all whose Young diagram is obtained from the Young diagram of by adding one addable box.
Proof.
We start by showing that has a filtration given by the as in the statement. First, we have the following tensor identity:
Noticing that is an exact functor from the category of finite dimensional -modules to the category of -modules, it remains to show that
where the sum is over all whose Young diagram is obtained from the Young diagram of by adding one addable box. This follows from the usual branching rule for -modules.
To show that the sum is direct, we use arguments from the infinite-dimensional representation theory of Lie algebras. We consider the usual category for [21, Chapter 4]. We then show that each lie in a different block of the category , which then implies that the sum is direct.
First, as is a quotient of the universal Verma module , these two modules share the same central character. Therefore and are in the same block if and only if the central characters afforded by and are the same. But these central characters are equal if and only if and are in the same orbit for the dot action of the symmetric group, which is the usual action of the symmetric group shifted by the sum of simple roots .
We obtain that and are in the same block if and only if there exists such that
Now, suppose that and are in the same block. Since the dot action satisfies , we deduce that so that lies in . Then, writing , we find that for every . Since both and are dominant weights, we deduce that for every . Indeed, each orbit for the dot action contains a unique dominant weight.
Hence if , the parabolic Verma modules and are in different blocks of the category . ∎
Using the previous proposition and induction, one shows the following corollary.
Corollary 3.5.
There is an isomorphism
where is the number of paths from the empty -partition to in the Young lattice of -multipartitions.
3.4. Braiding and an action of the Artin-Tits group of type
The quantized enveloping algebra (or rather a completion of the tensor product with itself) contains an element, called the quasi--matrix, which is a crucial tool in defining a braiding on a subcategory of the -modules. Since there are several possible braidings, we make our choice explicit and refer to [8, 10.1.D] for more details.
In a completion of , we define an element by
where and being the root vectors associated to a positive root . If and are two type weight modules over the ground ring where act locally nilpotently, induces an isomorphism of vector spaces . We then define a morphism of -modules
by
where is the flip and is the map if and are of respective weights and . This endows the category of type weight modules on which acts locally nilpotently with a braiding. In particular, we have the hexagon equation:
Let be the Artin-Tits braid group of type . It has the following presentation in terms of generators and relations:
Using the braiding, we define the following endomorphisms of :
Pictorially, one can represent these endomorphisms as
Proposition 3.6.
The assignment defines an action of on the module which commutes with the action.
Proof.
The fact that is a -morphism follows by definition of . The fact that the defining relations of are satisfied follows from the embedding of the braid group of type into the braid group of type [15, Lemma 2.1]. ∎
Finally, we end this section with a lemma due to Drinfeld [13, Proposition 5.1 and Remark 4) below] computing the action of the double braiding on highest weight modules, which is related with the action of the ribbon element.
Lemma 3.7.
Let and be highest weight modules of respective highest weight and such that . Then the double braiding restricted to acts by multiplication by the scalar
4. The endomorphism algebra of
The aim of this section is to prove the main result of this paper. We first explain why inherits an action of the Ariki–Koike algebra from the action of the braid group of type . It is a classical result that the eigenvalues of are and : the action of the braiding on is
Moreover, using 3.7, we easily compute the eigenvalues of the endomorphism in order to show that the action of factors through the Ariki–Koike algebra.
Lemma 4.1.
The eigenvalues of on are equal to
Proof.
Let be the highest weight of . The decomposition of is given in 3.4:
where is the -partition of whose only non-zero component is the -th one and is equal to . The highest weight of being , the action of on is given by
and we check that
∎
By the definition of the Ariki–Koike algebra, Proposition 3.6 and the previous lemma we thus get an action of the Ariki–Koike algebra for the parameters on . Therefore, the assignment defines a morphism of algebras
Theorem 4.2.
-
•
The algebra morphism is surjective and factors through an isomorphism
-
•
There is an isomorphism of -module
Proof.
The first part of the theorem follows immediately from the second part and the definition of the row-quotient .
Using 3.5 and the fact that acts on by -linear endomorphisms, we see that
for some -modules . Since the multiplicity of in is given by the number of paths in the Young lattice from the empty -partition to the -partition , we have . Showing that is a submodule of will end the proof of the second part of the theorem.
Let be a standard Young tableau of shape and denote by . Denote by the -partition of obtained by adding the boxes labeled by to in the chosen standard tableau to the empty -partition. We now choose a highest weight vector of weight such that for all we have
Using the branching rule, one see that such a vector exists and is unique up to a scalar. Let us show that this vector is a common eigenvector of the Jucys–Murphy elements . It is easy to see that the action of the Jucys–Murphy element on is given by the double braiding . By 3.7, we obtain that acts on by multiplication by
Indeed, lies in the summand of . But where so that
since the component of on is . Therefore, acts on by multiplication by
Therefore, the submodule spanned by is isomorphic to and then is a submodule of . ∎
4.1. Some particular cases
We finish by giving some special cases of 4.2 in order to recover various well-known algebras. The two first special cases involve the well-known situation without a parabolic Verma module: it suffices to note that, if , then is the trivial module.
Corollary 4.3.
If the parabolic subalgebra is and , then the endomorphism algebra of is isomorphic to Hecke algebra of type .
Corollary 4.4.
If the parabolic subalgebra is and , then the endomorphism algebra of is isomorphic to Temperley–Lieb algebra of type .
We now turn to special cases where is a strict subalgebra of . The following corollary follows from 2.12.
Corollary 4.5.
For such that and for all , the endomorphism algebra of is isomorphic to the Ariki–Koike algebra .
The Hecke algebra of type with unequal parameters appears when we work with a standard parabolic subalgebra with Levi factor .
Corollary 4.6.
If the parabolic subalgebra is such that , and , then the endomorphism algebra of is isomorphic to the Hecke algebra of type with unequal and algebraically independent parameters.
Finally, the last special case is a generalization of the case of [15], where we recover the generalized blob algebra.
Corollary 4.7.
If the parabolic subalgebra is the standard Borel subalgebra of , that is and for , then the endomorphism algebra of is isomorphic to the generalized blob algebra .
5. Some remarks on the non-semisimple case
This paper deals with the semisimple case, where the decomposition of as the sum of simple modules is a crucial tool to compute its endomorphism algebra. Non-semisimple situations appear if is no longer an indeterminate in the base field but a root of unity. If and the parameters appearing in the highest weight of are no longer algebraically independent, a non-semisimple situation may also appear. Indeed, the parabolic Verma module might not be simple anymore as it is readily seen from the case of . It is then natural to ask whether it is possible to extend the Schur–Weyl duality to the non-semisimple case. Let us remark that if is not a root of unity and if for all then the behavior is similar to the one described in the previous sections.
In order to define the action, we use an “integral version” of the algebras and and of the module , compatible with the specialization at a root of unity.
We start with the Ariki–Koike algebra. The definition given in Section 2.1 is valid for any field and any choice of parameters. Concerning the algebra , we consider Lusztig’s integral from over , see [8, Section 9.3]. It is also known that the quasi--matrix is an element of (a completion of) . Then for a base field and any , the quantum group is defined as , where we see as a -module via the morphism sending to .
The parabolic Verma module is a highest weight module and we choose a highest weight vector. We then have at our disposal an integral version, which is the submodule generated over by the highest weight . Its specialization at will still be denoted . Similarly, we have a version at of the standard module , which has a well-known integral form.
Since the quasi--matrix lies in the Lusztig’s integral form of the quantum group, we can similarly use the braiding to define the endomorphisms of the -module . As in the semisimple case, we have:
Proposition 5.1.
Let be a field, and . Then the assignment is a morphism of algebras from to . The parameters of the Ariki–Koike algebra are still given by 4.1.
It is more difficult to understand the image of map , or even better to describe the image and the kernel of the map. In [15] Iohara, Lehrer and Zhang studied the particular case of and (this corresponds to and ) and proved that if is an indeterminate in and that for , , then the map is surjective [15, Proposition 5.11].
In order to extend the Schur–Weyl duality form the semisimple case to a non-semisimple case, a classical strategy [12, 1] is to argue that the dimensions of the various algebras, such as or , are independent of the base field .
Following the arguments of [1], a first step would be to determine whether the parabolic Verma module is tilting in an appropriate category of infinite dimensional -modules. Since is tilting and the tensor product of tilting modules is tilting, having being tilting would mean that is. Since the space of endomorphisms of a tilting module is flat, its dimension does not depend on the base field .
Concerning , its definition is valid over the ring and it is known that the basis given in 2.3 is a basis over this ring. This implies that the dimension of the algebra is independent of the field and the choice of and of .
Therefore, if is tilting in an appropriate category of infinite dimensional -modules, the map would be surjective for any base field .
If we want to consider the row-quotients of , one must first give a definition which does not rely on the semisimplicity of the algebra so that the map factors through and then study the existence of an integral basis of .
Let us stress that these arguments depend heavily on being tilting and on the existence of an integral basis of . One may need some extra assumptions on the field , as for example being infinite, or on the parameters of the parabolic Verma module. This non-semisimple behavior deserves further study, which was outside the scope of this paper.
References
- [1] H. H. Andersen, C. Stroppel, and D. Tubbenhauer. Cellular structures using -tilting modules. Pacific J. Math., 292(1):21–59, 2018.
- [2] S. Ariki. On the Semi-simplicity of the Hecke Algebra of . J. Algebra, 169(1):216–225, 1994.
- [3] S. Ariki and K. Koike. A Hecke Algebra of and Construction of its Irreducible Representations. Adv. Math., 106(2):216–243, 1994.
- [4] S. Ariki, T. Terasoma, and H. Yamada. Schur–Weyl reciprocity for the Hecke algebra of . J. Algebra, 178(2):374–390, 1995.
- [5] M. Balagović, Z. Daugherty, I. Entova-Aizenbud, I. Halacheva, J. Hennig, M. S. Im, G. Letzter, E. Norton, V. Serganova, and C. Stroppel. The affine VW supercategory. Sel. Math. New Ser., 26, 2020.
- [6] H. Bao, W. Wang, and H. Watanabe. Multiparameter quantum Schur duality of type . Proc. Amer. Math. Soc., 146(8):3203–3216, 2018.
- [7] M. Broué and G. Malle. Zyklotomische Heckealgebren. Number 212, pages 119–189. 1993. Représentations unipotentes génériques et blocs des groupes réductifs finis.
- [8] V. Chari and A. Pressley. A guide to quantum groups. Cambridge University Press, Cambridge, 1994.
- [9] V. Chari and A. Pressley. Quantum affine algebras and affine Hecke algebras. Pacific J. Math., 174(2):295–326, 1996.
- [10] I. V. Cherednik. A new interpretation of Gelfand-Tzetlin bases. Duke Math. J., 54(2):563–577, 1987.
- [11] Z. Daugherty and A. Ram. Two boundary Hecke Algebras and combinatorics of type . 2018, arXiv: 1804.10296v1.
- [12] S. Doty. Schur-Weyl duality in positive characteristic. In Representation theory, volume 478 of Contemp. Math., pages 15–28. Amer. Math. Soc., Providence, RI, 2009.
- [13] V. Drinfeld. On almost cocommutative hopf algebras. Leningrad Math. J., 1(2):321–342, 1990.
- [14] M. Geck and N. Jacon. Representations of Hecke algebras at roots of unity, volume 15 of Algebra and Applications. Springer-Verlag London, Ltd., London, 2011.
- [15] K. Iohara, G. I. Lehrer, and R. B. Zhang. Schur–Weyl duality for certain infinite dimensional -modules. 2018, arXiv: 1811.01325v2.
- [16] M. Jimbo. A -analogue of , Hecke algebra, and the Yang-Baxter equation. Lett. Math. Phys., 11(3):247–252, 1986.
- [17] D. Kazhdan and G. Lusztig. Tensor structures arising from affine Lie algebras i-iv. J. Amer. Math. Soc., 6-7:905–947, 949–1011,335–381,383–453, 1993-1994.
- [18] A. Lacabanne, G. Naisse, and P. Vaz. Tensor product categorifications, Verma modules, and the blob algebra. 2020, arxiv: 2005.06257v1.
- [19] P. Martin and H. Saleur. The blob algebra and the periodic Temperley-Lieb algebra. Lett. Math. Phys., 30(3):189–206, 1994.
- [20] P. Martin and D. Woodcock. Generalized blob algebras and alcove geometry. LMS J. Comput. Math., 6:249–296, 2003.
- [21] V. Mazorchuk. Lectures on algebraic categorification. QGM Master Class Series. European Mathematical Society (EMS), Zürich, 2012.
- [22] D. Rose and D. Tubbenhauer. HOMFLYPT homology for links in handlebodies via type A Soergel bimodules. 2019, arXiv: 1908.06878v1.
- [23] R. Rouquier, P. Shan, M. Varagnolo, and E. Vasserot. Categorifications and cyclotomic rational double affine Hecke algebras. Invent. Math., 204(3):671–786, 2016.
- [24] M. Sakamoto and T. Sakamoto. Schur–Weyl reciprocity for Ariki–Koike algebras. J. Algebra, 221(1):293–314, 1999.