1 Introduction
1.1
The Iwahori–Hecke algebra $\mathcal {H}_q(W)$ is a q-deformation of the group algebra $KW$ of the Coxeter group W which was later extended to a deformation of a complex reflection group W (see [Reference Broué, Malle and RouquierBMR98]). In particular, it was established independently [Reference Broué and MalleBM93, Reference Ariki and KoikeAK94] by Broué-Malle and Ariki–Koike that a (quantum) algebra, called either the cyclotomic Hecke algebra or the Ariki–Koike algebra, deforms the complex reflection group $G(m,1,d) = C_m \wr \Sigma _d$ (i.e., the wreath product of a cyclic group $C_m$ by a symmetric group $\Sigma _d$ ). The Ariki–Koike algebra affords a theory that is closely connected to the affine Hecke algebra of type A.
For finite groups, there is a well-established theory involving subgroups and their relationship to the ambient group. However, for Hecke algebras, there are very few constructions of natural subalgebras that arise within these algebras. For example, an open problem is to develop a Sylow theory for Hecke algebra when q is specialized to a root of unity. A fundamental construction of Sylow subgroups for the symmetric group involves wreath products so a natural question to ask is whether one can make constructions of wreath products using Hecke algebras. The following problems will constitute the focus of this paper.
-
(A) Developing a theory of wreath products between (quantum) algebras that provides a uniform treatment to all Hecke-like algebras, including but not limited to the Ariki–Koike algebras, the affine Hecke algebras of type A and their (degenerate) variants, Rosso–Savage’s (affine) Frobenius Hecke algebras, affine zigzag algebras and the Hu algebra;
-
(B) Constructing ‘the Hecke algebra’ $\mathcal {H}_{m\wr d}$ that arises from the wreath product $\Sigma _m \wr \Sigma _d$ between the symmetric groups in the sense that the chain of subgroups $\Sigma _m^d \subseteq \Sigma _m \wr \Sigma _d \subseteq \Sigma _{md}$ deforms to a chain of subalgebras $\mathcal {H}_q(\Sigma _m^d) \leq \mathcal {H}_{m \wr d} \leq \mathcal {H}_q(\Sigma _{md})$ , where $\mathcal {H}_{m \wr d}$ affords a standard basis and a bar-invariant basis, both indexed by $\Sigma _m \wr \Sigma _d$ .
Throughout this paper, we consider algebras that are free over a commutative ring K. All tensor products are considered to be over K unless specified otherwise.
1.2 The Hu algebra
Our paper was initially motivated by the work of Jun Hu (see [Reference HuHu02]). Hu constructed an algebra $\mathcal {A}(m)$ in order to state a Morita equivalence theorem between the Hecke algebras of type D and of type A (under an invertibility condition),
In Hu’s first construction, he considered an equation in the type B Hecke algebra of unequal parameters $(1,q)$ that involves the Jucys–Murphy elements and proved that there exists a unique element $H_1$ (called $h_m^*$ therein) in the type A Hecke algebra $\mathcal {H}_q(\Sigma _{2m})$ that solves the equation, without an explicit formula for $H_1$ . The Hu algebra is defined as the subalgebra $\mathcal {A}(m) \subseteq \mathcal {H}_q(\Sigma _{2m})$ generated by $H_1$ as well as the parabolic subalgebra $\mathcal {H}_q(\Sigma _m\times \Sigma _m)$ . The Hu algebra should be regarded as the canonical Hecke algebra for $\Sigma _{m} \wr \Sigma _2$ . Recall that $\Sigma _m \wr \Sigma _2 \subseteq \Sigma _{2m}$ is generated by $\Sigma _m \times \Sigma _m$ together with a ‘thickened braid’ $t_1 := w_{m,m}$ described as below:
It is proved that $H_1$ specializes to $2^m t_1$ at $q=1$ . Moreover, $\mathcal {A}(m)$ admits a presentation with generators $H_1$ and $T_i$ ’s subject to the wreath relations $H_1 T_i = T_{i'} H_1$ , where $i'$ is either $i+m$ or $i-m$ , as well as a ‘quadratic relation’ of the form $H_1^2 = z_{m,m}$ for some central element $z_{m,m} \in Z(\mathcal {H}_q(\Sigma _m \times \Sigma _m))$ .
The key observation for constructing a general wreath product is that one should accommodate quadratic relations of the form: $H^2 = SH+ R$ , where $R,S$ are determined by certain elements in $B \otimes B$ for some algebra B; in contrast, in the usual quadratic relation $H^2 = (q-1)H+q$ , $R,S$ are always scalars in the base ring.
1.3 Quantum wreath product
We address the main problems in Section 1.1 simultaneously by introducing a new notion which we call the quantum wreath product. More specifically, starting from an algebra B, a given integer $d \ge 2$ and a choice Q of parameters, our quantum wreath product produces an algebra $B \wr \mathcal {H}(d) = B \wr _Q \mathcal {H}(d)$ generated by (lower level) tensor algebra $B^{\otimes d}$ together with (upper level) Hecke-like generators $H_1, \dots , H_{d-1}$ (see Definition 3.1.1). Other salient features are listed below.
-
(A) This procedure $B \mapsto B \wr _Q \mathcal {H}(d)$ that produces an associative algebra can be thought of as a quantization of the wreath product $G\mapsto G \wr \Sigma _d$ . In particular, as long as B is a deformation of the group algebra of a finite group G, our quantum wreath product $B \wr \mathcal {H}(d)$ is a deformation of the group algebra of $G \wr \Sigma _d$ .
-
(B) In general, B does not need to be finite-dimensional nor commutative. For instance, in the Hu algebra case, the base algebra $B= \mathcal {H}_q(\Sigma _d)$ is not commutative. In another case when $B = K[X^{\pm 1}]$ is the Laurent polynomial ring, there is a quantum wreath product $K[X^{\pm 1}] \wr \mathcal {H}(d)$ that realizes the affine Hecke algebra of type A.
-
(C) We are able to determine the necessary and sufficient conditions on the choice of parameters for $B \wr \mathcal {H}(d)$ to admit bases of the ‘right’ size (see Theorem 3.3.1). Our proof of the basis theorem relies on a key lemma having a similar flavor as Kashiwara’s grand loop induction (see Lemma 5.2.2).
In [Reference EliasE22], Elias proved a variation of our basis theorem for his Hecke-type categories (see Section 4.2.5, Section 5.4) in the following sense: (1) Elias’ proof uses Bergman’s diamond lemma for algebras over commutative rings, that is, the statement is valid for those quantum wreath products whose base algebras are commutative. Thus, Elias’ basis theorem does not imply our basis theorem. (2) His basis theorem is focused on describing the minimal set of ambiguities, which corresponds to the first step of our proof. In contrast, we proceed and then determine the conditions on the choice of parameters so that these ambiguities are resolvable. Such a determination of coefficients was not pursued in [Reference EliasE22]. (3) In contrast to Elias’ proof which is based on the diamond lemma, we provide a Humphreys-type proof based on an abstract version of the regular representation.
In Table 1, we provide a list of known examples that manifest themselves as quantum wreath products which are deformed from group or monoid algebras.
In the literature, Kleshchev–Muth first introduced in [Reference Kleshchev and MuthKM19] the rank n affinization as a general approach to certain variants of degenerate affine Hecke algebras. Their main example are the affine zigzag algebras. Savage introduced the affine wreath algebras in [Reference SavageSa20] as another unifying approach to some other variants of degenerate affine Hecke algebras, including Wan–Wang’s wreath Hecke algebras [Reference Wan and WangWW08]. Motivated by the endomorphism algebras in the (quantum) Frobenius Heisenberg categories, Rosso and Savage later developed in [Reference Rosso and SavageRS20] a quantum version which they call the quantum affine wreath algebras (or affine Frobenius Hecke algebras). Their theory recovers the action of the quantum Heisenberg category of [Reference Brundan, Savage and WebsterBSW20] on module categories for Ariki–Koike algebras. Also, their theory applies to the affine Sergeev algebra (i.e., the degenerate affine Hecke–Clifford algebra) and Evseev–Kleshchev’s super wreath product algebra appearing in their proof [Reference Evseev and KleshchevEK17] of Turner’s conjecture.
There have been papers written by Banica, Bichon, Lemeux and Tarrago on ‘free wreath product quantum groups’. Our construction of the quantum wreath product differs from this earlier notion. These authors considered the wreath product of a Hopf algebra with a quantum permutation group. Our quantum wreath product is the wreath product of an associative algebra with Hecke type algebras.
1.4 Additional features of quantum wreath products
We provide a brief summary of some additional significant properties of our quantum wreath product.
-
(A) Our theory not only provides a natural generalization of the theory of Rosso–Savage, and of Kleshchev–Muth (see Sections 4.3–4.4) but also has the advantage of being able to treat a multitude of Hecke-like algebras and their affinizations (degenerate or not) as a result of taking a single quantum wreath product. In particular, our general definition of quantum wreath product encompasses the three definitions in [Reference SavageSa20, Reference Rosso and SavageRS20] for their unifying theory. Furthermore, our theory applies to the Hu algebras, which are not covered by previous theory due to its intriguing quadratic relation.
-
(B) Another important feature of having a quantum wreath product realization is that the Schur–Weyl duality (i.e., double centralizer property) almost come for free, in the sense that the construction of $B \wr \mathcal {H}(d)$ already encodes necessary information for a Schur duality. By combining Theorem 7.3.1, Proposition 3.4.1 and Lemma 7.4.2, we obtain a machinery to establish a Schur duality for $B \wr \mathcal {H}(d)$ in the following sense, as long as the base algebra B is symmetric:
Note that Theorem 7.3.1 also applies in the following scenario, which turns out to be useful for the study of the Hu algebras:
Furthermore, we obtain a Schur functor from the existence of the splitting; see Section 7.2.
1.5 Bar-invariant basis arising from $\Sigma _m \wr \Sigma _2$
In this paper, we obtain a new explicit formula for $H_1$ as a sum of $2^m$ terms in which every term is a quantization of $t_1 := w_{m,m}$ after normalization. For example,
and
As a consequence, we discovered that $\mathcal {A}(m)$ is closed under the bar-involution, has a bar-invariant basis, in which a positivity with respect to the dual canonical basis is observed. In contrast, it is not known whether groups of the form $\Sigma _m \wr \Sigma _2$ afford a canonical basis theory since they are generally not complex reflection groups.
Moreover, with the explicit formula for $H_1$ , we are able to construct the generalized Hu algebra denoted by $\mathcal {H}_{m \wr d}$ . It is not clear at this time whether these algebras can be written as a quantum wreath product for $d \geq 3$ .
1.6
In the sequel, we will use the representation theory of $\mathcal {H}_{m\wr 2}$ to solve the type D analog of a conjecture by Ginzburg–Guay–Opdam–Rouquier in [Reference Ginzburg, Guay, Opdam and RouquierGGOR03]. This problem entails constructing a quasi-hereditary 1-cover for Hecke algebras of type $D_{n}$ via the Schur functor for a module category of the Schur algebra. When n is odd, the solution is very similar to the type B approach developed in [Reference Lai, Nakano and XiangLNX22], in light of a Morita equivalence theorem in [Pa94]; for type $D_{2m} = G(2,2,2m)$ . Our approach relies on constructing a 1-cover for the Hu algebra $\mathcal {A}(m)$ via the corresponding Schur algebra. We expect that a similar theory holds for type $G(d,d,dm)$ , in light of Hu–Mathas’ generalization of $\mathcal {A}(m)$ to a more general complex reflection group in [Reference Hu and MathasHM12].
1.7
The paper is organized as follows. In Section 2, we recall the preliminaries, including important examples of the wreath product of groups. In Section 3, we introduce the definition of our quantum wreath product; while in Section 4, we identify various Hecke-like algebras as quantum wreath products (see the Appendix A for definitions of these algebras). Section 5 is devoted to the basis theorem. In particular, we determine the necessary and sufficient conditions for a quantum wreath product to admit a basis of the ‘right’ size. See Appendix B for details of the proof. In Section 6, we obtain a new explicit formula for the special generator of the Hu algebra, which leads to a bar-invariant basis, as well as a Hecke subalgebra which quantizes the wreath product between symmetric groups. In Section 7, we develop a Schur–Weyl duality for quantum wreath product $A = B \wr \mathcal {H}(d)$ from those for the (symmetric) algebra B.
2 Preliminaries
2.1 Wreath products
Let $\Sigma _d$ be the symmetric group on the set $[d] := \{1, 2, \ldots , d\}$ generated by $S = \{s_1, \ldots , s_{d-1}\}$ consisting of simple transpositions $s_i = (i, i+1)$ . Define
Recall that for a group G, the wreath product $G \wr \Sigma _d$ is the semidirect product $G^d \rtimes \Sigma _d$ whose multiplication rule is determined by
When G is finite, say $G \subseteq \Sigma _m$ for some m, then the direct product $G^d \subseteq \Sigma _m^d \subseteq \Sigma _{md}$ . Thus, $G \wr \Sigma _d$ can be regarded as a subgroup of $\Sigma _{md}$ by identifying each $w \in \Sigma _d$ with the following permutation:
Example 2.1.1 (Generalized symmetric groups).
Let $C_m$ be the cyclic group of order m. The generalized symmetric group $C_m \wr \Sigma _d$ identifies with the complex reflection group of type $G(m,1,d)$ . For a fixed generator $x \in C_m$ , $C_m \wr \Sigma _d$ is generated by x and S subject to the relations in $\Sigma _d$ , together with that $x^m = 1$ and $s_{1} x s_{1} x = x s_{1} x s_{1}$ .
In particular, the hyperoctahedral group $C_2 \wr \Sigma _d$ can be identified either with the signed permutation group on $[\pm d] := [d] \sqcup -[d]$ or with the type B Weyl group $W(\mathrm {B}_d)$ generated by $S \cup \{s_0\}$ with extra braid relations $(s_0 s_1)^2 = (s_1s_0)^2$ and $s_0 s_i = s_i s_0$ for $i>1$ .
The complex reflection group $G(m,1,d)$ affords two nonisomorphic deformations – the Ariki–Koike algebras (or cyclotomic Hecke algebras, see Section A.4), and the Yokonuma–Hecke algebras.
Example 2.1.2 ( $\Sigma _m \wr \Sigma _2$ and Weyl groups of type D).
The wreath product $\Sigma _m \wr \Sigma _2$ is now identified with the subgroup of $\Sigma _{2m}$ generated by the Young subgroup $\Sigma _m^2 := \langle s_{(i-1)m+1}, \ldots , s_{im-1}~|~1\leq i\leq 2\rangle \subseteq \Sigma _{2m}$ , as well as the element $t_1=w_{m,m} \in \Sigma _{2m}$ (which corresponds to the generator of $\Sigma _2$ via Equation (2.1.3)), where $w_{a,b}\in \Sigma _{a+b}$ is now the element given by, in two-line notation,
It is well known that $w_{a,b} = s_{a\to 1} s_{a+1 \to 2} \dots s_{a+b-1 \to b}$ is a reduced expression and that
Consequently, we can treat both $\Sigma _m \wr \Sigma _2$ and $\Sigma _{2m}$ as subgroups of a larger wreath product, that is, the signed permutation group on $[\pm 2m]$ , and then obtain the following tower of groups:
Finally, we denote by $W(\mathrm {D}_r)$ the subgroup of $W(\mathrm {B}_r)$ generated by $s_1, \ldots , s_{r-1}$ and $s_u := s_0s_1s_0$ . Indeed, $W(\mathrm {D}_r)$ is a Weyl group of type $\mathrm {D}_r$ with the following Dynkin diagram:
2.2 Affine symmetric groups
Within this section, we allow G to be an infinite monoid in a wreath product $G \wr \Sigma _d$ . Now, we consider the weight lattice $P = \sum _{i=1}^d \mathbb {Z} \epsilon _i$ for $\mathrm {GL}_d$ . The lattice P contains the root lattice $Q = \sum _{i=1}^{d-1} \mathbb {Z} \alpha _i$ , where $\alpha _i = \epsilon _{i} - \epsilon _{i+1}$ . We further let $P^+ := \sum _{i=1}^d \mathbb {N} \epsilon _i$ . Denote the corresponding translation monoids by $t^L =\{ t^\lambda ~|~ \lambda \in L\}$ for $L\in \{P, P^+, Q\}$ , with $t^\lambda t^\mu = t^{\lambda +\mu }$ for all $\lambda , \mu \in L$ .
The extended and affine symmetric groups are the semidirect products $\Sigma ^{\mathrm {aff}}_d := t^Q \rtimes \Sigma _d$ and $\Sigma ^{\mathrm {ext}}_d := t^P \rtimes \Sigma _d$ , respectively, whose multiplications are uniquely determined by $ t^\lambda s_i = s_i t^{s_i(\lambda )}. $ Next, under the identification $P \equiv \mathbb {Z}^d, \lambda = \sum \lambda _i \epsilon _i \mapsto (\lambda _1, \dots , \lambda _d)$ , we see that the simple reflection $s_i$ on P agrees with the place permutation $s_i$ on $\mathbb {Z}^d$ :
Therefore, $\Sigma ^{\mathrm {ext}}_d$ can be realized as the wreath product $\mathbb {Z} \wr \Sigma _d$ whose multiplication is determined uniquely by $ (\lambda _1, \dots , \lambda _d)s_i = s_i(\lambda _1, \dots , \lambda _{i-1},\lambda _{i+1}, \lambda _i,\lambda _{i+2}, \dots , \lambda _d)$ ; while $\Sigma ^{\mathrm {aff}}_d$ is a subgroup given by
Note that we can thus regard the extended affine Hecke algebra (or degenerate affine Hecke algebra, resp.) of type A as a quantization of the wreath product $\mathbb {Z} \wr \Sigma _d = \Sigma ^{\mathrm {ext}}_d$ (or the submonoid $\mathbb {N} \wr \Sigma _d \subseteq \mathbb {Z} \wr \Sigma _d$ , resp.) as in Table 1.
Remark 2.2.1. $\Sigma ^{\mathrm {aff}}_d$ is a Coxeter group with generators $s_0, \dots , s_d$ and the following Dynkin diagram:
The extra generator $s_0$ is obtained as follows: Let $\theta := \alpha _1 + \dots + \alpha _{d-1} = \epsilon _1 - \epsilon _d$ be the highest root. Then
2.3 Iwahori–Hecke algebras
Let $q\in K^\times $ . For any Coxeter system $(W,S)$ , the Hecke algebra $\mathcal {H}_q(W)$ is the associative K-algebra generated by $T_s (s\in S)$ , subject to the braid relations for $(W,S)$ as well as the quadratic relations $T_s^2 = (q-1)T_s + q$ for $s\in S$ . For the Weyl group $W(B_n)$ of type B, we will also consider its multiparameter variant $\mathcal {H}_{(Q,q)}(W(\mathrm {B}_{2m}))$ subject to the same braid relations but with quadratic relations $T_i^2 = (q-1) T_i + q$ (for $i>0$ ) and $T_0^2 = (Q-1)T_0 + Q$ , where $Q\in K^\times $ .
2.4 Augmented algebras
Following [Reference Ginzburg and KumarGK93], by an augmented algebra we mean a K-algebra B equipped with a counit $\epsilon : B\to K$ . Denote by $B^+ := \ker \epsilon \unlhd B$ the augmentation ideal of B. Assume that $B \leq A$ is a normal subalgebra (i.e., $AB^+ = B^+A$ ). Then one can form the quotient algebra (which is also augmented), [Reference Ginzburg and KumarGK93, 5.2],
3 Quantum wreath products
Given a K-algebra B, and an integer $d \geq 2$ and a choice of parameter $Q=(R,S,\rho ,\sigma )$ , we introduce a construction called the quantum wreath product that produces an K-algebra $B \wr \mathcal {H}(d) = B \wr _Q \mathcal {H}(d)$ . Recall that a typical element in a wreath product group $G \wr \Sigma _d$ is of the form $(g_1, \dots , g_d)w$ for some $w\in \Sigma _d$ , $g_i \in G$ . We define an analogous algebra whose typical elements are linear combinations of elements the form below:
3.1 The definition
For an element $Z \in B\otimes B$ and for $1\leq i \leq d-1$ , set
Moreover, for an endomorphism $\phi \in \operatorname {End}_K(B\otimes B)$ , set
Definition 3.1.1 (Quantum wreath product).
Let B be an associative K-algebra, and let $d\in \mathbb {Z}_{\geq 2}$ . Let $Q = (R,S,\rho ,\sigma )$ be a choice of parameters with $R, S \in B \otimes B$ , and $\rho , \sigma \in \operatorname {End}_K(B\otimes B)$ such that $\sigma $ is an automorphism. The quantum wreath product is the associative K-algebra generated by the algebra $B^{\otimes d}$ and $H_1, \dots , H_{d-1}$ such that the following conditions hold, for $1\leq k \leq d-2, 1\leq i \leq d-1, |j-i|\geq 2$ :
Oftentimes, we refer this algebra as $B \wr \mathcal {H}(d)$ , or $B \wr _Q \mathcal {H}(d)$ whenever it is convenient.
Here, ${R_i}$ is an element in $B^{\otimes d}$ , in contrast to those endomorphisms $\text {R}_{i,i+1}$ in the literature acting on the ith and $(i+1)$ th factors, in the context of the Yang–Baxter equations.
The quadratic and wreath relations determine the following local relations in $B\otimes B$ , in the sense that one can recover Equations (3.1.4)–(3.1.5) by embedding the local relation into the ith and $(i+1)$ th positions of $B^{\otimes d}$ as well as replacing the symbol H by $H_i$ :
3.2 Necessary conditions
We first describe certain conditions on Q:
where ${l_X}, r_X$ for $X\in B\otimes B$ are K-endomorphisms defined by left and right multiplication in $B\otimes B$ by X, respectively. The next five conditions will only be necessary when $d\geq 3$ :
where $\{i,j\}= \{1,2\}$ , $r_X$ for $X\in B^{\otimes 3}$ is understood as right multiplication in $B^{\otimes 3}$ by X.
where $\{i,j\}= \{1,2\}$ .
Let $\{b_i \}_{i\in I}$ be a basis of B for some index set I, and let $H_w := H_{i_1} \dots H_{i_N} \in B\wr \mathcal {H}(d)$ for a reduced expression $w = s_{i_1} \dots s_{i_N}$ . Such an element $H_w$ is well defined due to the braid relations (3.1.3).
Proposition 3.2.1. Suppose that $B\wr \mathcal {H}(d)$ has a basis of the form $\{ (b_{\lambda _1} \otimes \dots \otimes b_{\lambda _d}) H_w ~|~ \lambda _j \in I, w\in \Sigma _d\}$ . Then, Conditions (3.2.1)–(3.2.4) hold. Moreover, if $d\geq 3$ , Conditions (3.2.5)–(3.2.9) hold.
Proof. The proposition follows from a direct calculation. See Appendix B.
The conditions can be simplified when we put reasonable assumptions on Q. For example, if we assume that $\sigma $ is the flip map $a\otimes b \mapsto b\otimes a$ and that $\rho $ is the zero map, then Conditions (3.2.1)–(3.2.9) are equivalent to the following conditions:
Note that the more involved conditions such as Equations (3.2.6)–(3.2.9) do not simplify much when $\rho \neq 0$ even if $\sigma $ is the flip map.
Remark 3.2.2. Equation (3.2.2) is equivalent to the statement that $\sigma $ is an algebra automorphism and that $\rho $ is a $\sigma $ -derivation. Then the wreath relation (3.1.5) appears in the definition of an Ore extension $B^{\otimes 2}[H_1; \sigma , \rho ]$ as the new multiplication rule for the polynomial ring $B^{\otimes 2}[H_1]$ . In other words, $B \wr \mathcal {H}(2)$ is the quotient of this skew polynomial ring $B^{\otimes 2}[H_1; \sigma , \rho ]$ by the quadratic relation $H_1^2 = SH_1 + R$ .
3.3 Structure theory
Now, we state the basis theorem regarding the sufficient and necessary conditions on Q such that $B \wr \mathcal {H}(d)$ affords desirable bases. This theorem will be proved in Section 5.
Theorem 3.3.1. Let $\{b_i\}_{i\in I}$ be a basis of B for some index set I. The following are equivalent:
-
(a) Conditions (3.2.1)–(3.2.4) hold and Equations (3.2.5)–(3.2.9) hold additionally if $d\geq 3$ ,
-
(b) $\{ (b_{\lambda _1} \otimes \dots \otimes b_{\lambda _d}) H_w ~|~ \lambda _j \in I, w\in \Sigma _d\}$ forms a basis of $B \wr \mathcal {H}(d)$ ,
-
(c) $\{H_w (b_{\lambda _1} \otimes \dots \otimes b_{\lambda _d}) ~|~ \lambda _j \in I, w\in \Sigma _d\}$ forms a basis of $B \wr \mathcal {H}(d)$ .
3.4
Consider a quantum wreath product $B \wr \mathcal {H}(d)$ . Assume additionally that B is an augmented algebra with counit $\epsilon : B \to K$ , R is invertible and that there is an $h_i \in K$ such that $h_i^2 = \epsilon (S) h_i + \epsilon (R)$ for each i. The space $B^{\otimes d}$ is also an augmented algebra, whose counit is also denoted by $\epsilon $ . It is natural to find the conditions on $\sigma $ and $\rho $ such that $B\wr \mathcal {H}(d)$ is an augmented algebra whose counit is given by
The counit is well defined if and only if $\epsilon (H_i H_{i+1} H_i) = \epsilon (H_i H_{i+1} H_i)$ and $\epsilon (H_i b) = \epsilon (\sigma _i(b) H_{i} + \rho _i(b))$ for all i and $b\in B^{\otimes d}$ . Equivalently, $h_i$ ’s are all the same since R (and hence $h_i$ ’s) are invertible. We can write $h := h_i$ . Moreover,
Hence, $B \wr \mathcal {H}(d)$ is augmented if $\epsilon \sigma = \epsilon $ and $\epsilon \rho = 0$ .
Note that Equation (3.4.2) can hold for more general $\sigma $ and $\rho $ . For examples, those satisfying $\sigma (b_1 \otimes b_2) = k(b_2\otimes b_1)$ for some unit $k\in K^{\times }$ , and $h(1-k) \epsilon = \epsilon \rho \in \mathrm {Hom}(B\otimes B, K)$ . Under the assumption (3.4.2), we can provide more information about the quantum wreath products and their quotients.
Proposition 3.4.1. Let $A = B \wr \mathcal {H}(d)$ be a quantum wreath product with two bases as in Theorem 3.3.1.
-
(a) Assume that both A and B are augmented algebras with counits both denoted by $\epsilon $ . If $\sigma (b_1 \otimes b_2) \in Kb_2\otimes b_1$ , then $B^{\otimes d}$ is a normal subalgebra of A.
As a consequence, $A // B^{\otimes d}$ is isomorphic to the associative K-algebra generated by $T_1, \dots , T_{d-1}$ subject to the usual braid relations of type A together with modified quadratic relations $T_i^2 = \epsilon (S)T_i + \epsilon (R)$ . Such an algebra is denoted by $\mathcal {H}_{\epsilon }(\Sigma _d)$ since it recovers the Hecke algebra $\mathcal {H}_q(\Sigma _d)$ when $\epsilon (S) = (q-1), \epsilon (R) =q$ for some $q\in K^\times $ .
-
(b) Assume that B is a symmetric algebra with trace map $\mathop {\text {tr}}: B\to K$ that induces to $\mathop {\text {tr}}: B^{\otimes d}\to K$ , $b_1\otimes \cdots \otimes b_d \mapsto \prod _i\mathop {\text {tr}}(b_i)$ . If $\mathop {\text {tr}} (\sigma (b)) = \mathop {\text {tr}}(b)$ for $b \in B\otimes B$ , R is invertible and $\rho = 0$ , then A is symmetric.
Proof. (a) Since $(B^{\otimes d})^+ \lhd B^{\otimes d}$ is an ideal, an arbitrary element in $A(B^{\otimes d})^+ = \langle H_i \rangle (B^{\otimes d})^+$ is of the form $\sum _{w}H_wb_w$ , where $\epsilon (b_w) =0$ for all w. Thus, it suffices to show that for any $b \in B^{\otimes d}$ with $\epsilon (b) =0$ , $H_ib \in (B^{\otimes d})^+A = (B^{\otimes d})^+ \langle H_j\rangle $ for all i.
Since $\sigma (b_1 \otimes b_2) \in Kb_2\otimes b_1$ , $\epsilon (\sigma _i(b))$ is a multiple of $\epsilon (b)=0$ . Thus,
where $\rho _i(b) \in (B^{\otimes d})^+$ since $\epsilon (\rho (b)) \in K \epsilon (b) =0$ for all $b\in (B^{\otimes d})^+$ , thanks to Equation (3.4.2).
(b) Since B is symmetric, $\beta : B\times B \to K, (a,b) \mapsto \mathop {\text {tr}}(ab)$ is a nondegenerate associative symmetric bilinear form. We claim that such a form for A is $\beta _A:A \times A \to K$ , given by
This trace map is well defined since A has the bases as in Theorem 3.3.1. It is not hard to verify that $\beta _A$ is nondegenerate and associative (for this to be true, it is crucial that R is invertible). It suffices to check that $\mathop {\text {tr}}(aH_x bH_y) = \mathop {\text {tr}}(bH_y aH_x )$ for all $a, b\in B^{\otimes d}, x,y \in \Sigma _d$ . By applying the wreath relation iteratively, we obtain $H_ya = \sigma _{y}(a) H_y$ , where $\sigma _{y}:= \sigma _{i_1} \cdots \sigma _{i_N}$ for a reduced expression $y = s_{i_1} \cdots s_{i_N} \in \Sigma _d$ , which is well defined as a consequence of Equation (3.2.5). Hence, for all $a, b \in B^{\otimes d}$ ,
Both traces are zero unless $x=y^{-1}$ . In the case $x = y^{-1}$ , assume that $x = s_{i_1} \cdots s_{i_N}$ is a reduced expression. If we write $H_y H_x = \sum _{z\in \Sigma _d} c_z H_z$ , then
On the other hand, if we write $H_x H_y = \sum _{z\in \Sigma _d} c^{\prime }_z H_z$ , then
Note that when $\rho = 0$ and R is invertible, Equation (3.2.3) become $\sigma (S) = S$ and $\sigma (R) = R$ . Since $\sigma _x = \sigma _{i_1} \cdots \sigma _{i_N}$ preserves the trace,
and so the two traces agree, that is, A is a symmetric algebra.
4 Quantum wreath product algebras
In the following, a list of algebras is provided which are special cases of our quantum wreath products. For each case, the reader is referred to Appendix A for a presentation of the algebra A in question by generators and relations. We choose suitably our parameters to realize these algebras as a quantum wreath product $B \wr \mathcal {H}(d)$ . In order to show that $A \simeq B \wr \mathcal {H}(d)$ , one can either show that the relations for A are equivalent to those for $B\wr \mathcal {H}(d)$ , under a certain identification, or use a dimension argument when A is finite-dimensional.
4.1 Hu algebras
The reader is referred to Section 6 for the definition of the Hu algebra $ \mathcal {A}(m)$ . The algebra $\mathcal {A}(m)$ is essentially different from all other examples in Section 4. It is not covered by the theory of Rosso–Savage nor of Kleshchev–Muth due to its intriguing quadratic relation.
Assume that K is a field of characteristic not equal to 2, $f_{2m}(q)\neq 0$ , $B=\mathcal {H}_q(\Sigma _m)$ and the parameter $Q = (R,S,\rho ,\sigma )$ is given by
For this choice, almost all conditions (3.2.1)–(3.2.9) follow immediately (see Equation (3.2.10)). The only nontrivial condition, $\sigma (R) = R$ , follows the construction of $z_{m,m}$ in Section 6.2.
The algebra $\mathcal {H}_q(\Sigma _m) \wr \mathcal {H}(2)$ is generated by $T_x \otimes T_y\ (x, y \in \Sigma _m)$ and $H_1$ . Note that there are no braid relations for a single generator $H_1$ . The quadratic relation translates to that $H_1^2 = z_{m,m}$ , and the wreath relation becomes
A dimension argument (by Proposition 6.2.2) shows that the Hu algebra $ \mathcal {A}(m)$ is canonically isomorphic to $\mathcal {H}_q(\Sigma _m) \wr \mathcal {H}(2)$ via the assignment below:
4.2 (Degenerate) Affine Hecke algebras
See Appendix A.1 for definitions of variants of affine Hecke algebras. Now, we consider the case when the base algebra B is generated by X and that $\sigma : a\otimes b \mapsto b\otimes a$ is the flip map. Verifying relations for all these variants can be made easier via the Demazure operator $\partial $ , given by
In particular, $\partial (X\otimes 1) = 1 = - \partial (1\otimes X)$ .
Proposition 4.2.1. Let $X \in B$ be an element not in the base ring K. Let $\sigma \in \operatorname {End}(B\otimes B)$ be the flip map, and let $\rho \in \operatorname {End}(B\otimes B)$ be such that Equation (3.2.2) holds. If $\beta := \rho (X\otimes 1)$ commutes with both $X\otimes 1$ and $1\otimes X$ , then the following holds for all $i,j \geq 0$ :
Moreover, if $X^{-1} \in B$ , then Equation (4.2.2) also holds for $i,j \in \mathbb {Z}$ .
Proof. First, it follows from Equation (3.2.2) that
On the other hand, $\rho (X \otimes X)$ is also equal to $\rho ((1\otimes X)(X\otimes 1)) = (X\otimes 1) (\beta + \rho (1\otimes X))$ , and hence $\rho (1 \otimes X) = -\beta $ . An inductive argument shows that, for all $i \geq 1$ ,
and hence,
This first part is done. Assume from now on $X^{-1} \in B$ . We apply Equation (3.2.2) to both $\rho ((X\otimes 1)(1\otimes X^{-1}))$ and $\rho ((1\otimes X^{-1})(X\otimes 1))$ and then equate the two. We then obtain (4.2.2) for ${i=0, j<0}$ . Similarly, by applying Equation (3.2.2) to both $\rho ((X^{-1}\otimes 1)(1\otimes X))$ and $\rho ((1\otimes X)(X^{-1}\otimes 1))$ and then equating the two, we obtain Equation (4.2.2) for ${i<0, j=0}$ . The most general case is then obtained by applying Equation (3.2.2) to $\rho ((X^i \otimes 1)(1\otimes X^j))$ .
In other words, for the verification of the conditions appearing in Theorem 3.3.1 involving $\rho $ , one only need to check them on the generator $X\otimes 1$ as long as $\beta := \rho (X\otimes 1)$ commutes with both $X\otimes 1$ and $1\otimes X$ .
Now, for any $\mu = \sum _{i=1}^d \mu _i \epsilon _i \in \sum _i \mathbb {Z} \epsilon _i$ , write $X^\mu \equiv X^{\mu _1}\otimes \dots \otimes X^{\mu _d}$ , and let $s_j$ act on $\sum _i \mathbb {Z} \epsilon _i$ by place permutation on the jth and $(j+1)$ th positions. For an element $X\in B$ , for $1\leq j \leq d$ , set
Corollary 4.2.2 (Bernstein–Lusztig relations).
Retain the assumptions in Proposition 4.2.1. The following local equalities in $B \wr \mathcal {H}(d)$ hold in the sense of Equation (3.1.6):
In particular, the following relations in $B \wr \mathcal {H}(d)$ hold, for $1\leq i \leq d-1$ , and for all $\mu $ ,
Proof. Equations (4.2.7) and (4.2.8) follow from direct computations using Equations (3.1.4)–(3.1.5); while Equation (4.2.9) is a paraphrase of Equation (4.2.8).
4.2.1 Affine Hecke algebras of type A
Let $B = K[X^{\pm 1}]$ be the Laurent polynomials in X over a field K, $q \in K^{\times }$ , and let
The extended affine Hecke algebra $\mathcal {H}^{\mathrm {ext}}_q(\Sigma _d)$ can be realized via the quantum wreath product $K[X^{\pm 1}] \wr \mathcal {H}(d)$ under the identification, for $\lambda = (\lambda _1, \dots , \lambda _d)$ :
Since $\beta = -(q-1) 1\otimes X$ commutes with either $X\otimes 1$ or $1\otimes X$ , Corollary 4.2.2 follows. Therefore, the wreath relations (3.1.5) implies the Bernstein–Lusztig relations (A.1.1) since
It is easy to see that the Bernstein–Lusztig relation implies the wreath relation, and hence, they are equivalent.
Since it is well known (see [Reference LusztigLu89]) that $\mathcal {H}^{\mathrm {ext}}_q(\Sigma _d)$ admits a basis of the form $\{Y^\mu T_w ~|~ \mu \in P, w\in \Sigma _d\}$ , conditions (3.2.1)–(3.2.9) hold as long as Theorem 3.3.1 is proved. On the other hand, one can give another proof of Lusztig’s basis theorem by a tedious but elementary verification of conditions (3.2.1)–(3.2.9).
Note that by this choice of $R, S$ , we have $H(X\otimes 1)H = q(1\otimes X)$ from Corollary 4.2.2, or equivalently, $H_i X^{(i)} H_i = qX^{(i+1)}$ for all i. If we choose $R = 1\otimes 1, S = (q^{-1} - q)(1\otimes 1)$ instead, we would obtain $H_i X^{(i)} H_i = X^{(i+1)}$ for all i.
The affine Hecke algebra $\mathcal {H}_q(\Sigma _d^{\mathrm {aff}})$ is then realized as the subalgebra
In particular, the affine generator $T_0\in \mathcal {H}_q(\Sigma _d^{\mathrm {aff}})$ can be identified as
4.2.2 Ariki–Koike algebras
Let $q_1, \dots , q_m\in K$ , and let $Q= (R, S, \rho , \sigma )$ be as in (4.2.10). The Ariki–Koike algebra is the following cyclotomic quotient of the quantum wreath product
In particular, the Jucys–Murphy elements $L_i$ ’s are identified as below:
Therefore, the well-known surjection $\mathcal {H}_q^{\mathrm {ext}}(\Sigma _d) \to \mathcal {H}_{q,q_1, \dots , q_m}(C_m\wr \Sigma _d), Y^{\epsilon _i} \mapsto L_i$ is the canonical map
4.2.3 Affine Hecke algebras of type B/C
The (nonextended) affine Hecke algebra $\mathcal {H}(C^{\mathrm {aff}}_d)$ of type C is the Hecke algebra $\mathcal {H}(C^{\mathrm {aff}}_d) = \langle T_0, \dots , T_d\rangle $ for the Coxeter group of type $C^{\mathrm {aff}}_d$ , in which $T_0$ corresponds to the type C node in the Dynkin diagram, while $T_d$ corresponds to the affine node. Let $B = K[X^{\pm 1}, Y^{\pm 1}]$ , $q, \xi , \eta \in K^{\times }$ , and let
Then
under the identification
On the other hand, the extended affine Hecke algebra $\mathcal {H}^{\mathrm {ext}}(B_d)$ of type B is generated by $T_0$ , $T_1$ , $\dots $ , $T_{d-1}$ , $Y^\lambda (\lambda \in P(B^\vee _d))$ , where $T_0$ corresponds to the type B node in the Dynkin diagram. It is possible to identify $\mathcal {H}^{\mathrm {ext}}(B_d)$ as a subalgebra of $K[X^{\pm 1}] \wr \mathcal {H}(2d)$ (with respect to Equation (4.2.10)) generated by
4.2.4 Degenerate affine Hecke algebras
Let $\mathcal {H}^{\mathrm {deg}}(d)$ be the degenerate affine Hecke algebra of type A, generated by $s_1, \dots , s_{d-1}, x_i, \dots , x_d$ (see Appendix A.1). The $s_i$ ’s and $x_j$ ’s interact by the cross relations (A.2.2). Let $B = K[X]$ be the polynomial ring. We choose
Then $\mathcal {H}^{\mathrm {deg}}(d) = K[X] \wr \mathcal {H}(d)$ under the identification
For $\lambda = \sum _i \lambda _i \epsilon \in P^+ := \sum _{i=1}^d \mathbb {Z}_+\epsilon _i$ , we set $x^\lambda := x_1^{\lambda _1} \dots x_d^{\lambda _d}$ . Now, $\beta = - 1$ commutes with either $X\otimes 1$ or $1\otimes X$ , and hence Corollary 4.2.2 follows. Similar to the case for affine Hecke algebras, our wreath relations (3.1.5) are equivalent to the Bernstein–Lusztig–type relations below, for $\lambda \in P^+, 1\leq i \leq d-1$ :
Since it is well known (see [Reference LusztigLu89]) that $\mathcal {H}^{\mathrm {deg}}(d)$ admits a basis of the form $\{x^\lambda w ~|~ \lambda \in P^+, w\in \Sigma _d\}$ , conditions (3.2.1)–(3.2.9) hold as long as Theorem 3.3.1 is proved.
Kostant–Kumar’s nil Hecke rings (see Appendix A.3) can be realized similarly by setting $R=0$ , instead.
4.2.5 Hecke-type categories
It is known that the Nil Hecke rings describe the endomorphism spaces in the Khovanov–Lauda–Rouquier (KLR) algebras. This generalizes to Elias’ Hecke-type category, whose endomorphism spaces (of rank d) can be regarded as an algebra generated by $H_1, \dots , H_{d-1}$ and $B^{\otimes d}$ (where B is a commutative ring; in contrast to our definition in which B is an arbitrary associative algebra). Their algebras satisfy our quadratic relations (3.1.4) as well as wreath relations (3.1.5), while the braid relations are relaxed as follows:
The key point is that our quantum wreath products is a special case of Elias’ construction when the base algebra B is commutative. However, in our main example, the Hu algebra, the base algebra $B= \mathcal {H}_q(\Sigma _m)$ is far from commutative.
4.3 Rosso–Savage algebras
Let $z\in K$ , and let B be a certain Frobenius K-algebra with a Nakayama automorphism $\psi :B\to B$ , a trace map $\text {tr}:B\to K$ , a basis I and its dual basis $\{b^\vee ~|~b\in I\}$ with respect to $\text {tr}$ . Rosso and Savage introduced in [Reference SavageSa20, Reference Rosso and SavageRS20] the Frobenius Hecke algebra $H_d(B,z)$ , its affinization $H^{\mathrm {aff}}_d(B,z)$ and the affine wreath product algebra $\mathcal {A}_d(B)$ which generalize the Hecke algebra, the affine Hecke algebra and the degenerate affine Hecke algebra, respectively (see Appendix A.5 for details).
We have the following identifications as quantum wreath products:
where $\hat {\otimes }$ denotes the twisted tensor product given by ${(b\otimes 1)X^{\pm 1} = X^{\pm 1}(\psi ^{\pm 1}(b) \otimes 1)}$ for $b\in B$ . Here, $R = 1\otimes 1$ and $\sigma (a\otimes b) = b\otimes a$ for all three cases. For $H_d(B,z)$ , the other parameters are
For $H^{\mathrm {aff}}_d(B,z)$ , one has
For $\mathcal {A}_d(B)$ , one has
In particular, these algebras include the Yokonuma–Hecke algebra $Y_{m,d} = KC_m \wr \mathcal {H}(d)$ , its affinization, Wan–Wang’s [Reference Wan and WangWW08] wreath Hecke algebra $\mathcal {H}^{\mathrm {deg}}(G\wr \Sigma _d) = K[X]G \wr \mathcal {H}(d)$ , Evseev–Kleshchev’s super wreath product algebra.
In [Reference Savage and StuartSS21], Savage and Stuart also considered variants of these algebra by replacing the quadratic relation with $H_i^2=0$ . By a similar argument, one can show that their Frobenius nil Coxeter algebras are also special cases of ours. Their Frobenius nil Hecke algebras can also be included as long as the base algebra is purely even.
4.4 Affine zigzag algebras
Readers are referred to Section A.6 for the definition of the zigzag algebra $Z_K(\Gamma )$ , its affinization $Z^{\mathrm {aff}}_d(\Gamma )$ over a simply-laced connected Dynkin diagram $\Gamma $ . Let $B = Z_{K[X]}(\Gamma )$ be the zigzag algebra over the polynomial ring $K[X]$ . Let $R = 1\otimes 1, S = 0, \sigma : a\otimes b \mapsto b\otimes a$ , and let $\rho : B\otimes B \to B \otimes B$ be determined by, for $\gamma _i \in Z_K, i \neq j \in I$ :
In other words, we have the following local relations, for $\gamma _i \in Z_K, i \neq j \in I$ :
Then $Z_d^{\mathrm {aff}}(\Gamma ) = Z_{K[X]} \wr \mathcal {H}(d)$ under the identification with $T_i \mapsto H_i$ , $z_j \mapsto X^{(j)}$ .
5 Basis theorem for quantum wreath products
This section is dedicated to determining the necessary and sufficient conditions on the choice of parameters such that that our quantum wreath product produces a unital associative algebra having a basis as given in Theorem 3.3.1.
5.1
We start with a left $B^{\otimes d}$ -module $V := B^{\otimes d} \otimes K \Sigma _d$ , on which $B^{\otimes d}$ acts by left multiplication. Let $A := B\wr \mathcal {H}(d)$ . Our goal is to show that, under suitable conditions, V can be given a right A-module structure, and the assignment $a\otimes w \mapsto a H_w$ for $a\in B^{\otimes d}, w\in \Sigma _d$ defines a right A-module isomorphism $V \to A$ . Note that V admits a filtration $(V^\ell )_{\ell \in \mathbb {N}}$ of $B^{\otimes d}$ -modules given by
We fix a reduced expression $r(w)$ for all $w\in \Sigma _d$ that is compatible with others, that is, $r(1) = 1$ and for every $w \in \Sigma _d \setminus \{1\}$ , there exists a unique $s_i$ such that $r(w) = r(w s_i) s_i$ as reduced expressions. Existence of such a choice follows from an inductive argument.
In order to eliminate the left-versus-right issue, in Section 5, we use the following postfix notation such that the juxtaposition of maps means the reversed composition of maps, that is, for any two maps $f:M_1 \to M_2, g:M_2 \to M_3$ , it is understood that
where $m\cdot f = f(m)$ is the postfix evaluation of f at m. In particular, for endomorphisms $f, g \in \operatorname {End}(M)$ , by the juxtaposition $fg$ we mean the the multiplication in the opposite ring $\operatorname {End}(M_1)^{\text {op}}$ .
Next, we define (left) $B^{\otimes d}$ -module homomorphisms $f^{(\ell )}_a = f^{(\ell )}_a(r) \in \operatorname {End}(V^\ell )$ , and $T_i^{(\ell )} = T_i^{(\ell )}(r):V^\ell \to V^{\ell +1}$ , for all $a\in B^{\otimes d}, 1\leq i \leq d-1, \ell \geq 0$ in a recursive manner:
for all w with length $\ell (w) \leq \ell $ . Note that the recursion does terminate since $T_i^{(0)}$ and $f_a^{(0)}$ are given by $(b \otimes 1) \cdot T_i^{(0)} = b\otimes s_i$ and $(b\otimes 1) \cdot f_{a}^{(0)} = (ba)\otimes 1$ for any $b\in B^{\otimes d}$ .
Proposition 5.1.1. For a fixed $a \in B^{\otimes d}$ (or a fixed $1\leq i\leq d-1$ , resp.), the maps $\{f_a^{(\ell )} ~|~ \ell \geq 0\}$ (or $\{T_i^{(\ell )} ~|~ \ell \geq 0\}$ , resp.) are compatible, that is, $f_a^{(\ell )}|_{V^j} = f_a^{(j)}$ and $T_i^{(\ell )}|_{V^j} = T_i^{(j)}$ for all $j \leq \ell $ .
Proof. We prove this by an induction on j. The base case $f_a^{(\ell )}|_{V^0} = f_a^{(0)}$ and $T_i^{(\ell )}|_{V^0} = T_i^{(0)}$ follow from Equations (5.1.3) and (5.1.4), respectively. For the inductive case, pick any $b\otimes x \in V^j$ with $x \neq 1$ (i.e., $\ell (x) \leq j \leq \ell $ ). There is an $s_k$ such that $r(x) = r(w)s_k$ , and thus $b\otimes w \in V^{j-1}$ . Then
The proof of that $T_i^{(\ell )}|_{V^j} = T_i^{(j)}$ for all $j \leq \ell , 1\leq i \leq d-1$ splits into two cases. The first case $\ell (x)> \ell (xs_i)$ follows from the fact that
The other case $\ell (x) < \ell (xs_i)$ follows from the fact that $(b\otimes x)\cdot T_i^{(\ell )} = b\otimes xs_i = (b\otimes x) \cdot T_i^{(j)}$ , thanks to Equation (5.1.4).
Let $f_a = f_a(r), T_i = T_i(r) \in \operatorname {End}_{B^{\otimes d}}(V)$ be given by $f_a|_{V^\ell } = f_a^{(\ell )}$ and $T_i|_{V^\ell } = T_i^{(\ell )}$ for all $\ell $ . Thanks to Proposition 5.1.1, these maps are well defined, and it leads to the following proposition from Equations (5.1.3)–(5.1.4).
Proposition 5.1.2. The following statements hold for all $w\in \Sigma _d, a\in B^{\otimes d}, 1\leq i \leq d-1$ :
We remark that one cannot define $f_a(r)$ and $T_i(r)$ directly using Equations (5.1.7)–(5.1.8) since a circular definition will occur. Later, we will show that these maps are actually independent of r.
Corollary 5.1.3. The map $f: B^{\otimes d} \to \operatorname {End}_{B^{\otimes d}}(V)^{\text {op}}, a \mapsto f_a$ is K-linear. In particular, $f_a+f_b = f_{a+b}$ for all $a, b\in B^{\otimes d}$ .
Proof. It suffices to prove that f is K-linear on $1^{\otimes d} \otimes w$ for any $w \in \Sigma _d$ . From Equation (5.1.7), it is clear that f is K-linear in $1^{\otimes d} \otimes w$ if $w=1$ . If $\ell (w)> 1$ , then there is an $s_i$ such that $r(w) = r(ws_i)s_i$ . By an induction on $\ell (w)$ , f is K-linear on $1^{\otimes d} \otimes w$ by combining Equation (5.1.7) and the fact that both $\sigma _i$ and $\rho _i$ are K-linear.
5.2 Grand loop argument
In the following, we present a key ingredient towards the proof of Theorem 3.3.1 that has a similar flavor as Kashiwara’s grand loop argument, that is, we will prove intermediate statements below based on an induction on $\ell $ :
Note that the Condition $R[\ell ]$ implies that $T_i \in \mathrm {Hom}_{B^{\otimes d}}(V^\ell , V)$ does not depend on r as well.
Proposition 5.2.1. Assume that Equations (3.2.1)–(3.2.4) hold, and Equations (3.2.5)–(3.2.9) hold additionally if $d\geq 3$ .
Proof. The proposition follows from an involved calculation that can be found in the Appendix B.
Lemma 5.2.2. Assume that Equations (3.2.1)–(3.2.4) hold, and Equations (3.2.5)–(3.2.9) hold additionally if $d\geq 3$ . Then, $W[\ell ], M[\ell ], Q[\ell ], B_2[\ell ], B_3[\ell ]$ and $R[\ell ]$ hold for all $\ell $ .
Proof. The base cases can be verified directly. For the inductive step, we assume that $W[i-2], M[i-1], Q[i-1], B_2[i-2], B_3[i-3], R[i-1]$ hold for all $i \leq \ell $ . Therefore, for $n \in \{2,3\}$ ,
The proof concludes by proceeding inductively.
We remark that the proof of Lemma 5.2.2 work perfectly even for the degenerate case when $d=2$ . Since there is a unique choice of r for $\Sigma _2$ , $R[\ell ]$ holds for all $\ell $ . On the other hand, $B_n[\ell ]$ are vacuous statements and hence always are true. The remaining conditions are verified by using the induction above. The grand loop degenerates to a loop that only involves $W[i]$ ’s, $M[i]$ ’s and $Q[i]$ ’s. The implications therein are given by Proposition 5.2.1 (M), (Q) and (W), which in fact only require Equations (3.2.1)–(3.2.4).
5.3 Proof of the basis Theorem 3.3.1
Proof. First, we prove that (a) is equivalent to (b). One direction is already taken care of by Proposition 3.2.1. For the other direction, we will give V a right A-module structure, and prove that the left $B^{\otimes d}$ -module homomorphism $\phi : V \to A$ , $a \otimes w \mapsto a H_w$ is a right A-module isomorphism.
Step 1: We first show that $\phi $ is onto. Note first an arbitrary element $x\in A$ is a sum of elements of the form $a_1 \cdots a_n$ for some $n=n(x)$ and each $a_j$ is either a generator $H_i$ or an element in $B^{\otimes d}$ . By applying the wreath, quadratic and braid relations iteratively, x becomes a sum of elements of the form
and thus $\phi $ is onto.
Step 2: We show that $\phi $ is an isomorphism. Thanks to Lemma 5.2.2, $W[i], M[i], Q[i], B_2[i], B_3[i]$ and $R[i]$ hold for all i. Hence, the assignment $a \mapsto f_a, H_w \mapsto T_w$ induces an algebra homomorphism $\Phi : A \to \operatorname {End}_{B^{\otimes d}}(V)^{\text {op}}$ , where $T_w := T_{i_1} \cdots T_{i_N}$ if $w = s_{i_1} \cdots s_{i_N}$ is reduced. Using this map, we obtain a right A-module structure of V.
Note that any element in V is of the form $v = \sum _{w\in \Sigma _d} b_w \otimes w$ for some $b_w= b_w(v) \in B^{\otimes d}$ . It suffices to show that if $\sum _{w\in \Sigma _d} b_w H_w = 0$ , then $\sum _{w\in \Sigma _d} b_w\otimes w = 0$ . Indeed,
Therefore, (a) and (b) are equivalent.
Next, we show that (b) and (c) are equivalent. For each $w\in \Sigma _d$ , let $\underline {w} = (s_{i_1}, \dots , s_{i_n})$ be a fixed reduced expression of w, and set $\sigma _{\underline {w}} := \sigma _{i_1} \circ \dots \circ \sigma _{i_n}$ .
By applying the wreath, quadratic and braid relations iteratively to a typical element $H_w b \in A$ , one obtains
with respect to the Bruhat order on $\Sigma _d$ . By substituting b for $\sigma _{\underline {w}}^{-1}(b)$ , Equation (5.3.3) becomes
where the latter inclusion follows applying the former part of (5.3.4) to every $B^{\otimes d} H_x$ , iteratively. In other words, the transition matrices are invertible since every $\sigma _{\underline {w}}$ is an automorphism. Thus, (b) and (c) are equivalent.
5.4 Elias’ basis Theorem
In the study of Hecke-type categories (see Section 4.2.5), the base algebra B is always a commutative ring. In our setup, the base algebra B can be an associative algebra. A fundamental question is to determine whether a diagrammatic algebra generated by crossings and dots with relations given by resolving crossings has the right size. In [Reference EliasE22], Elias proved a Bergman diamond lemma for the endomorphism algebras (see Section 4.2.5) regarding sufficient and necessary conditions in terms of (conceptual) resolvable ambiguities, which corresponds to the very first step of our proof of Theorem 3.3.1.
In particular, our conditions (3.2.1)–(3.2.9) describe explicitly the requirements on the choice of parameters in order to resolve the minimal set of ambiguities given in [Reference EliasE22, (5.1), (5.11)], except for a couple of them that do not appear in our setup since we require the braid relations to hold in our case.
6 The Hu algebra $\mathcal {A}(m)$ and its generalization $\mathcal {H}_{m\wr d}$
In this section, Hu’s original construction of $\mathcal {A}(m)$ is presented, in which the extra generator $H_1 \in \mathcal {H}_q(\Sigma _{2m})$ is defined using an implicit procedure via Jucys–Murphy elements $u_m^{\pm }$ of the Hecke algebra of type $B_{2m}$ with unequal parameters $(1,q)$ . Our new result is an explicit formula for $H_1$ , which leads to a construction of a bar-invariant basis for $\mathcal {A}(m)$ . The explicit formula also makes it possible to construct a generalized Hu algebra $\mathcal {H}_{m\wr d} \subseteq \mathcal {H}_q(\Sigma _{md})$ .
6.1 The construction of the element $h_{m}$
Let $d=2m$ . In order to quantize Equation (2.1.6), consider the type B Hecke algebra $\mathcal {H}^{\mathrm {B}} = \mathcal {H}_{(1,q)}(W(\mathrm {B}_{2m}))$ of unequal parameters $(1,q)$ . One can realize the Hecke algebras for $\Sigma _m \times \Sigma _m$ and for $\Sigma _{2m}$ as subalgebras of $\mathcal {H}^{\mathrm {B}}$ via the generating sets $\{T_1, \dots , T_{m-1}, T_{m+1}, \dots T_{d-1}\}$ and $\{T_1, \dots , T_{d-1}\}$ , respectively. It is well known that $T_w = T_{i_1} \cdots T_{i_r} \in \mathcal {H}^{\mathrm {B}}$ is well defined if $w = s_{i_1} \cdots s_{i_r} \in W(B_d)$ is a reduced expression. The tower of groups in Equation (2.1.6) can be recovered by taking the specialization $q=1$ from the following tower of algebras:
where $\mathcal {A}(m)$ is the Hu algebra which we will define shortly. From now on, all the computation are carried out inside $\mathcal {H}^{\mathrm {B}}$ in the sense of Equation (6.1.1).
Definition 6.1.1. Let $T_{a\to b}$ and $T_{a\to b \to a}$ be the element in $\mathcal {H}^{\mathrm {B}}$ corresponding to $s_{a\to b}$ and $s_{a\to b\to a}$ (see (2.1.1)), respectively. Denote the Jucys–Murphy elements by
It is well defined because its factors commute pairwise. Note that an empty product is understood as 1 so $u^{\pm }_0 = 1$ . The following result is a useful variant of [Reference Dipper, James and MurphyDJM95, Lemma 4.9].
Lemma 6.1.2.
-
(a) $u^+_j$ commutes with $T_i$ for $i \in \{ 0,1, \dots , d-1\}\setminus \{j\}$ .
-
(b) For $i\geq 1$ , $u_i^+ T_{i\to 0} = u_{i+1}^+ T_{1\to i}^{-1} - q^i u_i^+ T_{1\to i}^{-1}$ .
-
(c) Let $x\in \Sigma _{2m}$ , $i\geq 1$ , and let $\mathcal {H} := \mathcal {H}_q(\Sigma _{2m})$ . Then
(6.1.3) $$ \begin{align} u_i^+ T_x T_0\in u_i^+ \mathcal{H} +\sum_{j> i} \mathcal{H} u_j^+ \mathcal{H}. \end{align} $$In particular,(6.1.4) $$ \begin{align} u_i^+ T_{i\to 0} \in u_i^+(-q^i T_{1\to i}^{-1}) + \sum_{j> i} \mathcal{H} u_j^+ \mathcal{H}. \end{align} $$
Proof. Parts (a) and (b) are immediate consequences of the definitions. For part (c), write $x = yc \in \Sigma _{2m}$ for $y \in \Sigma _i \times \Sigma _{2m-i}$ and $c \in \Sigma _{2m}$ the shortest representative in the coset $(\Sigma _i \times \Sigma _{2m-i})x$ so that $u_i^+ T_x = u_i^+ T_y T_c = T_y u_i^+ T_c$ by part (a). If c fixes 1, then $T_c$ commutes with $T_0$ , and hence,
since $u_i^+ T_0 = u_i^+$ when $i \geq 1$ . We are done in this case.
If c does not fix $1$ , then $c = s_{i\to 1} z$ is a reduced expression for some $T_z$ that commutes with $T_0$ , and hence
As a consequence of the lemma, the definition below makes sense as one can expand $T_{w_{m,m}} u_m^-$ and then apply Equation (6.1.3) iteratively to define $h_{m}$ .
Definition 6.1.3. Let $h_m$ be the unique element in $\mathcal {H}:= \mathcal {H}_q(\Sigma _{2m})$ such that
6.2 Definition of $\mathcal {A}(m)$
With the construction of $h_m$ , one can now define the Hu algebra.
Definition 6.2.1 (Hu algebra).
Let $\mathcal {A}(m)$ be the subalgebra of $\mathcal {H}_q(\Sigma _{2m})$ generated by
where $H_1= h_m^*$ , and $*$ is the unique antiautomorphism of $\mathcal {H}_q(\Sigma _{2m})$ determined by $T_i^* = T_i$ for all $1\leq i < 2m$ .
Several of the important properties that Hu proved about $\mathcal {A}(m)$ are summarized below.
Proposition 6.2.2 [Reference HuHu02, 1.7, 1.9, 1.10, 2.2, 2.4].
If $d= 2m \geq 2$ , then
-
(a) $\dim \mathcal {A}(m) = |\Sigma _m \wr \Sigma _2| = 2(m!)^2$ .
-
(b) The element $h_m^*$ is invertible if $f_{2m}(q):= \prod _{i=0}^{2m-1}(1+q^i) \neq 0$ .
-
(c) $(h_m^*)^2 = h_m^2$ , and they are equal to the unique element $z_{m,m} \in Z(\mathcal {H}_q(\Sigma _m \times \Sigma _m))$ such that
$$\begin{align*}u_m^+T_{w_{m,m}} u_m^- T_{w_{m,m}} u_m^+T_{w_{m,m}} u_m^- = u_m^+T_{w_{m,m}} u_m^- z_{m,m}. \end{align*}$$ -
(d) $\mathcal {A}(m)$ admits a presentation with generators $T_1, \dots , T_{m-1}, T_{m+1}, \dots , T_{2m}, H_1$ subject to the braid and quadratic relations for $T_i$ ’s as well as the following relations:
$$\begin{align*}H_1^2 = z_{m,m}, \quad T_i H_1 = \begin{cases} H_1 T_{i+m} & \mathrm{if}\ i < m; \\ H_1 T_{i-m} &\mathrm{if}\ i>m. \end{cases} \end{align*}$$ -
(e) At the specialization $q=1$ , $ h_m = 2^m w_{m,m}$ , and $\mathcal {A}(m)$ specializes to the group algebra of $\Sigma _{m} \wr \Sigma _2$ .
6.3 A new formula for $H_{1}$
In this section, we assume that there is an invertible element $v\in K^\times $ such that $v^{2} = q$ . This is true when K is an algebraically closed field and when K is the ring of Laurent polynomials in v. The main goal of this section is to give a new and useful interpretation of $H_{1}$ .
For simplicity, we consider a renormalized standard basis $\{I_w := v^{-\ell (w)}T_w~|~ w\in \Sigma _{2m}\}$ of $\mathcal {H}_q(\Sigma _{2m})$ satisfying, for all $I_i := I_{s_i}$ ,
Denote by $\overline {\phantom {a}}: \mathcal {H}_q({\Sigma _{2m}}) \to \mathcal {H}_q({\Sigma _{2m}})$ the bar involution sending v to $v^{-1}$ and $I_w$ to $I_{w^{-1}}^{-1}$ . We use the shorthand notation
Denote by $\{\pm \}^m := \{+, -\}^m$ the set of m-tuples of plus or minus signs. For $\epsilon = (\epsilon _1, \dots , \epsilon _m) \in \{\pm \}^m$ , let
and let $\Sigma _d$ act on $\{\pm \}^m$ by place permutations.
Proposition 6.3.1.
-
(a) Let $\epsilon = (\epsilon _1, \dots , \epsilon _m) \in \{\pm \}^m$ . If $i<m$ , then
$$\begin{align*}I_\epsilon I_i = I_{i+m} I_{s_i(\epsilon)}. \end{align*}$$ -
(b) If $a,b>c$ , then $I_{a\to c} I_{c\to a}$ commutes with $I_{b\to c} I_{c\to b}$ .
-
(c) For any $j\in \mathbb {Z}$ , define an assignment
(6.3.4) $$ \begin{align} \pi_{jm}: T_i \mapsto T_{i+jm}\quad \text{for all }i. \end{align} $$Let $c_{m,0}:=1$ , and let $c_{m,i}:= \pi _{m}(I_{i\to 1} I_{1\to i}) = I_{m+i \to m+1}I_{m+1 \to m+i}$ for $i\geq 1$ . Then $c_{m,i}$ commutes with $c_{m,j}$ for all $i,j \geq 1$ . Moreover, the following product is well defined:(6.3.5) $$ \begin{align} C_{\epsilon} = \prod_{i; \epsilon_i = -} c_{m,i-1}. \end{align} $$
Proof. (a) This is a direct consequence of the braid relations. One obtains a graphical proof by treating $I_i$ and $I_i^{-1}$ as opposite crossings of the ith and the $(i+1)$ th strings.
(b) It suffices to prove the case $a=b+1$ . Note that
it then follows by an inductive argument that $I_b, \dots , I_c$ (and hence $I_{b\to c} I_{c\to b}$ ) commutes with $I_{a\to c} I_{c\to a}$ .
(c) The first assertion follows from Part (b) since $c_{m,i} c_{m,j} = \pi _m (I_{i\to 1}I_{1\to i} I_{j\to 1} I_{1\to j})$ . The second assertion is an immediate consequence of the first one.
Lemma 6.3.2. Let $h_{m,i}$ be the unique element in $\mathcal {H}:=\mathcal {H}_q(\Sigma _{2m})$ such that
Then, the elements $h_{m,i} (1\leq 1 \leq m)$ are given by the formula below recursively:
In particular, $h_m = h_{m,m}$ is given by
Proof. The base case $i=1$ follows immediately from Lemma 6.1.2(c). For the inductive step, by definition of $u_i^-$ , we have
The first term on the right-hand side of Equation (6.3.10) is
and hence its contribution to $h_{m,i+1}$ is $ q^{i} h_{m,i} T_{i+m\to i+1} = v^{m+2i} h_{m,i} I_{i+m\to i+1}$ . The second term on the right-hand side of Equation (6.3.10) is
and hence its contribution to $h_{m,i+1}$ is $v^{m+2i} c_{m,i} h_{m,i} I^-_{i+m \to i+1}$ , and we are done.
Theorem 6.3.3. We have an explicit formula for $H_1 = H_1(m) :=h_m^*$ as follows:
Example 6.3.4. Consider $H_1 \in \mathcal {A}(2) \subseteq \mathcal {H}_q(\Sigma _4)$ . We use abbreviations such as $\overline {T}_2 T_1 \overline {T}_3 \overline {T}_2 := T_{\overline {2}.1.\overline {3.2}}$ . Then $h_{2} = v^6 (I_{2.1.3.2} + I_{2.1.\overline {3.2}} + I_3^2(I_{\overline {2.1}.3.2} + I_{\overline {2.1}.\overline {3.2}}){)}$ , or
While $I_{2.1.3.2}^2$ consists of summands involving $I_2$ in terms of the standard basis, its alternative, $H_1$ , squares into the parabolic subalgebra $\mathcal {H}_q(\Sigma _2 \times \Sigma _2) = \langle I_1, I_3\rangle $ . To be precise, we obtained an explicit formula for the central element $z_{2,2} = H_1^2$ (see Proposition 6.2.2(c)):
We currently have no explanations for these coefficients appearing in $z_{m,m}$ .
6.4 Bar-invariant bases
In this section, we assume $K = \mathbb {Z}[v^{\pm 1}]$ . The element $H_1$ is defined via solving an equality (see Definition 6.1.3) that involves Jucys–Murphy elements in the type B Hecke algebra of unequal parameters. There seems to be no a priori reason for $H_1$ to admit an explicit formula in Hecke algebra of Type A. In this section, we show miraculously that $H_1$ , after being multiplied by an element arising from the longest element, leads to a highly nontrivial bar-invariant element $b_1$ and thus to a bar-invariant basis for the Hu algebra. The element $b_1$ seems to afford a positive expansion in terms of the dual canonical basis of the ambient Hecke algebra and in turn affords a potential theory of geometric realization and/or categorification for such wreath product $\Sigma _m \wr \Sigma _d$ that are not Coxeter groups in general.
Proposition 6.4.1. Recall $\pi _m$ from (6.3.4). Let $C := C_{(-,-,\dots ,-)}$ , and let $\gamma := \pi _m(I_{w_\circ (m)})$ where $w_\circ (m) \in \Sigma _m$ is the longest element. Then:
-
(a) $I_{w_\circ (m)} = I_{w_\circ (m-1)} I_{m-1\to 1}$ .
-
(b) $I_{w_\circ (m)} I_i = I_{m+1-i} I_{w_\circ (m)}$ .
-
(c) $C = \gamma ^2$ .
Proof. Part (a) follows from the fact that $s_1 s_{2\to 1} \dots s_{m-1\to 1} = w_\circ (m)$ is a reduced expression. Part (b) follows from the fact that $w_\circ (m) s_i = s_{m+1-i} w_\circ (m)$ for all i. For part (c), we prove it by an induction on m. The base case $m=2$ follows since $w_\circ (2) = s_1$ and $C= I_3^2$ as computed in Example 6.3.4. For $m \geq 3$ , we have
The extra generator $H_1$ behaves well with respect to the bar involution in the following way:
Proposition 6.4.2. The following element is bar-invariant:
Proof. Note that for any $\epsilon \in \{\pm \}^m$ ,
where $\overline {\epsilon } := -\epsilon \in \{\pm \}^m$ . Recall that $H_1 = v^{m(2m-1)}\sum _\epsilon I_\epsilon C_\epsilon $ , so $b_1 = \sum _\epsilon I_\epsilon C_\epsilon \overline {\gamma },$ and hence
Therefore, $b_1$ is bar-invariant as long as
Thanks to Proposition 6.4.1(c), one has $\gamma ^2 = C = C_{\overline {\epsilon }} C_{\epsilon }$ for any $\epsilon $ , and hence Equation (6.4.5) follows from a right multiplication by $\overline {\gamma }$ and a left multiplication by $\overline {C}_{\overline {\epsilon }}$ .
With the help of this distinguished element $b_1$ , we are now able to construct a bar-invariant basis $\{b_x~|~x\in \Sigma _m\wr \Sigma _2\}$ of the Hu algebra that satisfies a unitriangular relation with respect to a standard basis $\{I_x ~|~ x\in \Sigma _m \wr \Sigma _2\}$ . See Remark 6.4.5 for a discussion on the positivity of $\{b_x\}_x$ in comparison with the canonical bases.
Theorem 6.4.3. Let $I_{wt_1} := I_w b_1$ for $w\in \Sigma _m \times \Sigma _m$ . Then $\{I_x ~|~ x\in \Sigma _m \wr \Sigma _2\}$ forms a standard basis of $\mathcal {A}(m)$ in the sense that $\overline {I_x} \in I_x + \sum _{y<x} \mathbb {Z}[v^{\pm 1}] I_y$ for all $x\in \Sigma _m \wr \Sigma _2$ .
Proof. It follows from the fact that $\{I_w ~|~ \Sigma _m \times \Sigma _m\}$ forms a standard basis for $\mathcal {H}_q(\Sigma _m\times \Sigma _m)$ with respect to the Bruhat order inherited from $\Sigma _{2m}$ . In particular, for $x = wt_1$ , we have $\overline {I_x} = \overline {I_w} b_1 \in I_{x t_1} + \sum _{y<x} \mathbb {Z}[v^{\pm 1}] I_{y t_1}$ .
Theorem 6.4.4. Let $\{b_w ~|~ \Sigma _m \times \Sigma _m\}$ be the canonical basis for $\mathcal {H}_q(\Sigma _m\times \Sigma _m)$ in the sense that $b_w \in I_w + \sum _{y<w} v\mathbb {Z}[v] I_y$ , and let $b_{wt_1} := b_w b_1$ for $w\in \Sigma _m \times \Sigma _m$ . Then $\{b_x ~|~ x\in \Sigma _m \wr \Sigma _2\}$ forms a bar-invariant basis of $\mathcal {A}(m)$ satisfying
with respect to the Bruhat order inherited from $\Sigma _{2m}$ .
Proof. The bar-invariant part is due to Proposition 6.4.2. The unitriangular part follows from the fact that $b_{xt_1} = b_x b_1 \in I_{xt_1} + \sum _{y<x} v\mathbb {N}[v] I_{yt_1}$ .
Remark 6.4.5. While the canonical basis of $\mathcal {H}_q(W)$ of any Weyl group W admits positive structural constants, that is, $b_x b_y \in \sum _z \mathbb {N}[v^{\pm 1}] b_z$ ; our bar-invariant basis does not have such positivity because $b_1^2$ has mixed signs already.
However, $b_1$ seems to admit a negative expansion in the canonical basis, equivalently, a positive expansion in the dual canonical basis $\{c_w ~|~ w\in \Sigma _{2m}\}$ , that is,
where the dual basis element $c_w$ is characterized by $\overline {c}_w = c_w \in I_w + \sum _{y<w} (-v^{-1})\mathbb {N}[-v^{-1}] I_y$ , and it satisfies that $c_x c_y \in \sum _z \mathbb {N}[-v^{\pm 1}] c_z$ .
6.5
We present a concrete calculation of $b_{1}$ in a specific example below.
Example 6.5.1. For $H_1 = H_1(3)$ , write $I_{\epsilon _1 \epsilon _2}:= I_{(\epsilon _1,\epsilon _2,+)} + I_{(\epsilon _1,\epsilon _2,-)}$ for short. Then $H_1 = v^{15}(I_{++} + I_{+-}I_{4.4} + I_{-+}I_{5.4.4.5} + I_{--}I_{4.4.5.4.4.5})$ , $\gamma = I_{4.5.4}$ , and hence
The element $b_1$ is indeed bar-invariant since
Examples for a positive expansion of $b_1=b_1(m)$ in the dual canonical basis are shown below: for $m=1$ ,
For $m=2$ , we abbreviate $c_{s_{i}s_{j}\dots s_{k}}$ by $c_{i.j.\dots .k}$ , and obtain:
6.6 Generalized Hu algebra
With the explicit formula above, we are in a position to define a Hecke subalgebra $\mathcal {H}_{m\wr d}$ of $\mathcal {H}_q(\Sigma _{md})$ for the wreath product $\Sigma _m \wr \Sigma _d$ . Note that Hu’s Hecke subalgebra is a special case $\mathcal {A}(m) = \mathcal {H}_{m \wr 2}$ .
Definition 6.6.1. Let $\mathcal {H}_{m\wr d}$ be the subalgebra of $\mathcal {H}_q(\Sigma _{md})$ generated by
where $H_{j+1} = \pi _{jm}(h^*_m)$ (see Equation (6.3.4)). The algebra $\mathcal {H}_{m \wr d}$ will be called the generalized Hu algebra with $\mathcal {H}_{m \wr 2} = \mathcal {A}(m)$ .
The braid relations fail for the elements $H_i$ ’s. For example, in the special case when $m=1$ , the $H_i$ ’s satisfy the modified braid relations
which appeared in the nonstandard presentation of the type A Hecke algebra (cf. [Reference WangWa07]). In other words, Wang’s nonstandard presentation of $\mathcal {H}_q(\Sigma _d)$ is just the ‘standard’ presentation of the generalized Hu algebra $\mathcal {H}_{1\wr d}$ . This suggests that there could be a general theory for covering algebras for $\mathcal {H}_{m\wr d}$ .
7 The Schur algebras for quantum wreath products
7.1
In this section, we investigate the connection between various properties that are referred as the ‘Schur–Weyl duality’, including the double centralizer property, the existing of a splitting, and the existence of a certain idempotent. A systematical approach is provided to obtain the Schur–Weyl duality for the quantum wreath product $A = B \wr \mathcal {H}(d)$ arising from a Schur–Weyl duality for the base algebra B. As a consequence, the Schur–Weyl duality for the Hu algebra $\mathcal {A}(m) = \mathcal {H}_q(\Sigma _m) \wr \mathcal {H}(2)$ is established for the first time. It should be noted that our approach is quite encompassing and allows us to recover the Schur–Weyl duality for other Hecke-like algebras appearing in Section 4.
Definition 7.1.1 (Wreath Schur algebra).
Assume that T is a right A-module. We call $S^A(T) := \operatorname {End}_A(T)$ the wreath Schur algebra.
7.2
Various constructions have appeared in the literature that have involved endomorphism algebras and the double centralizer property (cf. [Reference Cline, Parshall and ScottCPS96, Chapter 1] [Reference Doty, Erdmann and NakanoDEN04, Sections 2-3]). In familiar cases like commuting actions of the general linear group and the symmetric group, the ‘Schur functor’ is constructed via an explicit idempotent. For more general settings, idempotents in algebras are not easy to locate. For the purposes of this paper, we present an alternative approach via splittings to prove the existence of an idempotent $e\in S$ (where S is an endomorphism algebra) that can be used to construct a functor from $\text {Mod}(S)$ to $\text {Mod}(eSe)$ .
For now, assume that A is an arbitrary associative K-algebra, and T is a right A-module with $S=\operatorname {End}_{A}(T)$ . Then T is a left S-module (under composition), and T is an S-A-bimodule. Suppose one has a splitting, as right A-modules,
There exists a contravariant functor ${\mathcal G}(-):=\text {Hom}_{A}(-,T)$ whose image is a left S-module. Under this functor, ${\mathcal G}(A)=\text {Hom}_{A}(A,T)\cong T$ is a direct summand of S as a left S-module by using the splitting. Consequently, T is a projective left S-module.
Under the condition (7.2.1), there exists an idempotent $e\in S$ such that $T\cong Se$ . Moreover,
One can define a Schur functor ${\mathcal F}:\text {Mod}(S) \rightarrow \text {Mod}(eSe)$ via ${\mathcal F}(M)=eM$ . In this setting, the theory of Schur functors and their inverses has been developed in [Reference Doty, Erdmann and NakanoDEN04].
7.3
The question remains when one can identify A with $eSe$ . When this holds the aforementioned Schur functor will conveniently take objects in $\text {Mod}(S)$ to $\text {Mod}(A)$ . When A is a self-injective algebra the following result is a corollary of a theorem proved by König, Slungard and Xi [Reference König, Slungard and XiKSX01]. The original statement of the corollary first appeared in [Reference Curtis and ReinerCR62, 59.6].
Theorem 7.3.1. Let $S=\operatorname {End}_{A}(T)$ . Assume that
-
(a) A is a finite-dimensional self-injective algebra,
-
(b) T is a faithful A-module.
Then $\operatorname {End}_{S}(T)\cong A$ .
Condition (a) of the theorem will be satisfied when A is a quantum wreath product as long as the base algebra B is a symmetric algebra and the assumptions in Proposition 3.4.1(b) hold. For examples of quantum wreath products, to ensure the existence of a natural Schur functor, one needs to establish a splitting of T as in (7.2.1). This will ensure that condition (b) of the theorem is satisfied.
7.4 A splitting lemma, I
For applications in Schur–Weyl duality, we make the following assumptions:
Within Section 7.4, we also assume that
Let $V_B$ be the vector space over K with basis $\{v_i ~|~ i \in I\}$ , where I is a (possibly infinite) poset that is totally ordered. Assume that $V_B$ is a right B-module, and hence $V_B^{\otimes d}$ is a right module over $B^{\otimes d}$ via the factorwise action. For $\mu = (\mu _j)_j \in I^d$ , set $v_\mu = v_{\mu _1} \otimes \ldots \otimes v_{\mu _d} \in V_B^{\otimes d}$ . Assume that there is $\mu \in I^d$ such that $\mu _1 < \dots < \mu _d$ , and let $W = W(\mu )$ be the subspace of $V_B^{\otimes d}$ spanned by elements of the form $v_{\mu \cdot g}$ , where $\Sigma _d$ acts on $I^d$ by place permutations. Let the Hecke-like generators $H_i$ ’s in $A = B \wr \mathcal {H}(d)$ act on the $B^{\otimes d}$ -submodule W from the right by
Thanks to the assumption (7.4.2) and that I is totally ordered, the action (7.4.3) on W extends to the following action on $V_B^{\otimes d}$ , as long as Equation (7.4.3) is well defined:
We now show that Equation (7.4.3) is well defined in two separate cases. The first case is when B acts on $V_B$ (and hence on the tensor products of $V_B$ ) by a counit $\epsilon :B \to K$ . The second case is the degenerate case when A is just a twisted tensor product of $B^{\otimes d}$ by a group algebra $K\Sigma _d$ .
Proposition 7.4.1. Assume that Equation (7.4.1) holds. Then:
Proof. For (a), the verification of the quadratic relations reduces to the rank two case: for $i < j$ , $(v_{(i,j)} H_1) H_1 = v_{(j,i)} H_1 = \epsilon (S) v_{(j,i)} + \epsilon (R) v_{(i,j)} = v_{(i,j)} (H_1^2)$ . The other case $i> j$ can be checked similarly. The verification of the braid relations reduces to the rank three case. We only present the most complicated case for $i < j < k$ :
When $i=j$ , the braid relations hold due to a case-by-case analysis similar to the ones for $i\neq j$ . The verification of quadratic relations reduces to the rank two case:
The wreath relations hold since $\rho = 0$ and the fact that both $a\otimes b, b\otimes a \in B\otimes B$ act on $V_B^{\otimes 2}$ by the same scalar multiple $\epsilon (a)\epsilon (b)$ .
For (b), the A-action on W is given explicitly by $v_\mu. H_i = v_{\mu \cdot s_i}$ . The braid and quadratic relations are immediate, while the wreath relations follow from the rank two verification:
where $b = \sum _k b^{\prime }_k \otimes b^{\prime \prime }_k \in B\otimes B$ .
Schur–Weyl duality in the literature often assumes a condition $n \geq d$ for the double centralizer property to hold. In the following lemma, we demonstrate that the condition $n\geq d$ is essentially a shadow of the assumption that $V_B$ contains d distinct direct summands which are isomorphic to B as B-modules. To be precise, one needs to assume additionally the subspace W obtained from these direct summands admits an A-action given by Equation (7.4.3) and right multiplications by $B^{\otimes d}$ (e.g., the two cases in Proposition 7.4.1). In such a case, $W \cong A$ as A-modules, and hence the desired splitting is obtained when the A-action on W extends to $V_B^{\otimes d}$ .
Lemma 7.4.2. Assume that Equation (7.4.1) holds, and there are right B-module isomorphisms $\phi _j: U_j \to B, v_{\mu _j} b \mapsto b$ such that $U_j$ ’s are distinct direct summands of $V_B$ .
If Equation (7.4.3) extends to an A-action on $W=W(\mu )$ , then $W \simeq A$ as right A-modules. If Equation (7.4.3) extends further to an A-action on $V_B^{\otimes d}$ , then $V_B^{\otimes d}$ has the following splitting, as right A-modules,
Proof. We may assume that ${\mu _1} < {\mu _2} < \dots < {\mu _d}$ . It suffices to show that the assignment
defines an A-module isomorphism $W \to A$ . We show first that $\psi $ is well defined. Suppose that $v_\mu H_w b = v_\mu H_{w'}b'$ for some $w, w'\in \Sigma _d, b,b' \in B^{\otimes d}$ . By Equation (7.4.3), $v_\mu H_w = v_{\mu \cdot w} \in U_{w(1)} \otimes \cdots U_{w(d)}$ so is $v_\mu H_w b$ because $U_j$ ’s are summands. Since $U_j$ ’s are all distinct, w must agree with $w'$ . Therefore, $b = (\phi _{w(1)} \otimes \dots \otimes \phi _{w(d)})(v_{\mu \cdot w} b) =(\phi _{w(1)} \otimes \dots \otimes \phi _{w(d)})(v_{\mu \cdot w} b') = b'$ .
Next, it is easy to verify that $\psi $ is an epimorphism. For injectivity, thanks to the basis theorem, an arbitrary element in A is of the form $a= \sum _{w\in \Sigma _d} H_w (b_1^{(w)} \otimes \dots \otimes b_d^{(w)})$ for some $b^{(w)}_i \in B$ . Since ${\mu _1} < {\mu _2} < \dots < {\mu _d}$ , by Equation (7.4.3), $H_w$ acts on $v_{({\mu _1}, {\mu _2}, \dots , {\mu _d})}$ by place permutation.
Suppose that $v_{({\mu _1}, {\mu _2}, \dots , {\mu _d})}.\sum _{w\in \Sigma _d} H_w (b_1^{(w)} \otimes \dots \otimes b_d^{(w)}) = 0$ . Each $v_{({\mu _1}, {\mu _2}, \dots , {\mu _d})} H_w (b_1^{(w)} \otimes \dots \otimes b_d^{(w)})$ lies in a different summand $U_{w(\mu _1)} \otimes \dots \otimes U_{w(\mu _d)} \subseteq V_B^{\otimes d}$ , and hence $b_1^{(w)} \otimes \dots \otimes b_d^{(w)}$ must be zero for all w.
Finally, a double centralizer property is obtained in the context of Theorem 7.3.1, where the S-A-bimodule T is $V_B^{\otimes d}$ .
Corollary 7.4.3. Let $A = B \wr \mathcal {H}(d)$ be a quantum wreath product for a finite-dimensional algebra B such that Proposition 3.4.1(b) holds. Assume that Lemma 7.4.2 holds. Then, the Schur duality holds, that is, $\operatorname {End}_{S^A(T)}(T)\cong A$ , where $T = V_B^{\otimes d}$ . Moreover, there exists a Schur functor.
7.5 Examples
7.5.1 Hecke algebras of type A
We retain the setup as in Section 4.2.2 and then set $m=1$ . Therefore, $B=K$ and $A=B \wr \mathcal {H}(d)$ is just the Hecke algebra of type A. Note that in this degenerate case, X is identified with $q_1 \in K^\times $ and $\rho = 0$ . Let $V_B=V(n)$ be the K-space spanned by $v_i$ where $i \in I := \{1, \dots , n\}$ with the usual total order. Thus, $V_B^{\otimes d}$ admits an A-action by Proposition 7.4.1(a). To be precise, for $1\leq i \leq d-1$ ,
The assumptions in Lemma 7.4.2 hold if and only if $V(n)$ contains a d-dimensional subspace, that is, $n\geq d$ , which is the well-known condition to afford the quantum Schur–Weyl duality [Reference JimboJ86, Reference Dipper and JamesDJ89]. Moreover, $S^A(V_B^{\otimes d})$ coincides with the type A q-Schur algebra $S_q(n,d) := \operatorname {End}_{\mathcal {H}_q(\Sigma _d)}(V(n)^{\otimes d})$ .
Corollary 7.5.1. If $n\geq d$ , then the Schur duality holds between the Hecke algebra $\mathcal {H}_q(\Sigma _{d})$ and its corresponding q-Schur algebra $S_q(n,d)$ . Moreover, a Schur functor exists.
7.5.2 Evseev–Kleshchev’s wreath product algebras
Let B be an arbitrary superalgebra, and let $R = 1\otimes 1, S = 0, \sigma : a\otimes b \mapsto b\otimes a$ and $\rho = 0$ . So $A = B \wr \mathcal {H}(d)$ is the super wreath product algebra $W_d^B = B^{\otimes d} \rtimes K \Sigma _d$ in [Reference Evseev and KleshchevEK17]. Note that $V_B^{\otimes d}$ admits an A-action by Proposition 7.4.1(c). Let $n\in \mathbb {N}$ , and let $V_B = \bigoplus _{i=1}^n B$ as a right B-supermodule in the usual way. If $n \geq d$ , then the assumptions in Lemma 7.4.2 hold. In this case, $U_i$ ’s are just those distinct summands in $V_B$ .
Thus, $S^A((B^{\otimes n})^{\otimes d})$ coincides with the Evseev–Kleshchev’s generalized Schur algebra $S^B(n,d) := \operatorname {End}_{W^B_d}((B^{\otimes n})^{\otimes d})$ , and hence we cover the double centralizer property result in [Reference Evseev and KleshchevEK17].
Corollary 7.5.2. If $n\geq d$ , then the Schur duality holds between the super wreath product algebra $W^B_d$ and its corresponding Schur algebra $S^B(n,d)$ . Moreover, a Schur functor exists.
7.6 A splitting lemma, II
While in Section 7.4 we demonstrate an application of Theorem 7.3.1 via a construction of the splitting over the quantum wreath product $B \wr \mathcal {H}(d)$ from a splitting over the base algebra B, in this section we provide a new application of Theorem 7.3.1 via a different construction of the splitting. To be precise, given an algebra $A'$ , we can obtain the desired splitting over $A'$ as long as we have a splitting over a larger algebra A which contains $A'$ as a self-injective subalgebra.
Lemma 7.6.1. Let A be an associative algebra, and let $A' \subseteq A$ be a self-injective subalgebra. Suppose that one has a splitting of a right A-module T as in Equation (7.2.1). Then, one has a splitting $T \cong A' \oplus Q'$ , as right $A'$ -modules for some $Q'$ .
Proof. Note that $A'$ is a right $A'$ -submodule of A. By the self-injectivity of $A'$ , the right $A'$ -module $A'$ is a direct summand of A. The result follows from the fact that the right A-module A is a direct summand of T.
7.6.1 Hu algebras
Following Section 4.1, let $B = \mathcal {H}_q(\Sigma _m)$ be generated by $T_1, \dots , T_{m-1}$ , and let $\mathcal {A}(m) = B\wr \mathcal {H}(2)$ the Hu algebra generated by $B^{\otimes 2}$ and an Hecke-like generator H. Next, we will apply Lemma 7.6.1 to the case $A' := \mathcal {A}(m) \subseteq A := \mathcal {H}_q(\Sigma _{2m})$ .
Let $V(n)$ be the K-span of $v_1, \dots , v_n$ , and let $T = T(n, 2 m) := V(n)^{\otimes 2m}$ . The Hecke algebra $\mathcal {H}_q(\Sigma _{2m})$ (and hence the Hu algebra $\mathcal {A}(m)$ ) acts on T by Equation (7.5.1). Therefore, we obtain the following diagram regarding the relationship between the q-Schur algebra $S_q(n,2m)$ and the wreath Schur algebra $S^{\mathcal {A}(m)}(T)$ .
Corollary 7.6.2. If $n\geq 2m$ , then the Schur duality holds between the Hu algebra $\mathcal {A}(m)$ and its corresponding Schur algebra $S^{\mathcal {A}(m)}(T(n,2m))$ . Moreover, a Schur functor exists.
Proof. Recall that $A := \mathcal {H}_q(\Sigma _{2m}) \supseteq A' := \mathcal {A}(m) \cong B \wr \mathcal {H}(2)$ , where $B = \mathcal {H}_q(\Sigma _m)$ is symmetric, $\sigma $ is the flip map, $\rho = 0$ and $R = z_m$ is invertible. Thus, the Hu algebra is symmetric thanks to Lemma 3.4.1(b), and hence $A'$ is a self-injective subalgebra of A.
When $n \geq 2 m$ , one has a splitting $T \cong A \oplus Q$ as A-modules. Then, by Lemma 7.6.1, one obtains another splitting $T \cong \mathcal {A}(m) \oplus Q'$ as $\mathcal {A}(m)$ -modules. Therefore, we can apply Theorem 7.3.1 and then obtain the desired Schur duality. The existence of a Schur functor follows from the discussion in Section 7.2.
7.7 Related Schur–Weyl dualities
In this section, we consider Schur–Weyl dualities in which the corresponding base algebras are infinite-dimensional. In this case, the notion of symmetric algebra is not defined. In turns out that we can still construct a natural A-action to afford the splitting (7.2.1), and in turn state the most natural double centralizer property which awaits for a uniform proof.
We also include a proof of Schur duality for the Ariki–Koike algebras using Theorem 7.3.1. Note that our splitting lemma does not apply here.
7.7.1 Affine Hecke algebras
Following Section 4.2.1, we have $B= K[X^{\pm 1}]$ and $A = B\wr \mathcal {H}(d)$ is isomorphic to the extended affine Hecke algebra $\mathcal {H}_q^{\mathrm {ext}}(\Sigma _d)$ of type A. Let $n\in \mathbb {N}$ , and let $V_B = \widehat {V}(n)$ be the space spanned by $\{v_i ~|~ i\in \mathbb {Z}\}$ on which $X^{\pm 1}$ acts by
If $n \geq d$ , then the assumptions in Lemma 7.4.2 hold. In this case, $U_j := v_j.B (1\leq j \leq d)$ are indeed distinct summands.
One can extend the right A-action to the entire tensor space $T = V_B^{\otimes d}$ via Equation (7.5.1). Then $S^A(T)$ coincides with the affine q-Schur algebra of type A. The corresponding Schur–Weyl duality is proved in [Reference Chari and PressleyCP94, Reference GreenGr99] as below:
Proposition 7.7.1 [Reference Chari and PressleyCP94, Reference GreenGr99].
If $n\geq d$ , then the Schur duality holds between the affine Hecke algebra $\mathcal {H}^{\mathrm {ext}}_q(\Sigma _{d})$ and its corresponding affine q-Schur algebra $S^{\mathrm {aff}}_q(n,d) := \operatorname {End}_{\mathcal {H}^{\mathrm {ext}}_q(\Sigma _d)}(\widehat {V}(n)^{\otimes d})$ .
7.7.2 Ariki–Koike algebras
Consider $F = \frac {{K[X]}}{\prod _{i=1}^m (X- q_i)}$ , where $q_i$ ’s are parameters as $v_i$ in Section 4.2.2. Let $n\in \mathbb {N}$ , and let $V_F=V(m,n)$ be the F-module with basis $\{v_1, \dots , v_{mn}\}$ , on which the F-action is given by, for $1\leq j \leq m, 0 \leq k \leq n-1$ ,
where $e_i$ ’s are the elementary symmetric functions that appear in rewriting $\prod _{i=1}^m (X-q_i) =0$ into $X^m = \sum _{i=1}^{m} (-1)^{i-1} e_i(q_1, \dots , q_m) X^{m-i}$ .
Next, we define an A-module T such that $T = V_F^{\otimes d}$ as a vector space. Each $H_i (1\leq i \leq d-1)$ acts by Equation (7.5.1), and each generator $\overline {X^{(i)}} \in A$ acts on the right by $H_{i-1}\dots H_1 X^{(1)} H_1 \dots H_{i-1}$ , where $X^{(1)}$ acts on the first tensor factor by Equation (7.7.2).
If $n \geq d$ , then the splitting as in Equaton (7.2.1) exists since the span of $\{v_\mu ~|~ 1\leq \mu _i \leq md\}$ is a direct summand of T that is isomorphic to A as a right A-module. It is known that A is symmetric (and hence self-injective) by [Reference Malle and MathasMM98, Theorem 5.1], and thus, by Theorem 7.3.1, we cover the results on Schur duality in [Reference GreenGr97, Reference Bao and WangBW18, Reference Bao, Wang and WatanabeBWW18]. Moreover, when $m> 2$ , $S^A(T)$ is a new Schur algebra for the Ariki–Koike algebra that has the double centralizer property. See [Reference Sakamoto and ShojiSS99, Reference James and MathasJM00] for attempts to obtain such a Schur duality.
Corollary 7.7.2. If $n\geq d$ , then the Schur duality holds between the Ariki–Koike algebra $\mathcal {H}_{q, q_1, \dots , q_m}(C_m \wr \Sigma _{d})$ and its corresponding Schur algebra $S^A(V(m,n)^{\otimes d})$ .
7.7.3 Affine Hecke algebras of type C
Following Section 4.2.3, let $F = \mathcal {H}_q(\Sigma ^{\mathrm {aff}}_2)$ be generated by $X,Y$ modulo the quadratic relations $X^2 = (\xi ^{-1} - \xi )X + 1$ and $Y^2 = (\eta ^{-1} -\eta )Y +1$ for parameters $\xi , \eta \in K^{\times }$ . Let $V_F = \widetilde {V}(n)$ be the space spanned by $\{v_i ~|~ i\in \mathbb {Z}\}$ , on which F acts by
For the right A-module structure on $T = {V}_F^{\otimes d}$ , $H_i$ ’s act by Equation (7.5.1), $\overline {X^{(i)}}, \overline {Y^{(d-i+1)}} \in A$ act on the right by $H_{i-1}\dots H_1 X^{(1)} H_1 \dots H_{i-1}$ and $H_{d-i} \dots H_{d-1}Y^{(d)}H_{d-1} \dots H_{d-i}$ , respectively, where $X^{(1)}$ and $Y^{(d)}$ act on the first and last tensor factors, respectively, by Equation (7.7.3). Then $S^A(T)$ coincides with the affine q-Schur algebra of type C. The corresponding Schur–Weyl duality, after a renormalization, is proved in [Reference Fan, Lai, Li, Luo, Wang and WatanabeFL3W220] as below:
Proposition 7.7.3 [Reference Fan, Lai, Li, Luo, Wang and WatanabeFL3W220].
If $n\geq d$ , then the Schur duality holds between the affine Hecke algebra $\mathcal {H}_{q,\xi ,\eta }(C^{\mathrm {aff}}_{d})$ and its corresponding affine q-Schur algebra $\widehat {S}^{\mathfrak {c}}_{q,\xi ,\eta }(n,d) := \operatorname {End}_{\mathcal {H}_{q,\xi ,\eta }(C^{\mathrm {aff}}_{d})}(\widetilde {V}(n)^{\otimes d})$ .
7.7.4 Degenerate Hecke algebras of type A
Following Section 4.2.4, we have $B= K[X]$ and $A = B\wr \mathcal {H}(d)$ is isomorphic to the degenerate affine Hecke algebra $\mathcal {H}^{\mathrm {deg}}(\Sigma _d)$ of type A. Let $n\in \mathbb {N}$ , and let $V_B = \overline {V}(n)$ be the space spanned by $\{v_i ~|~ i\in \mathbb {Z}_{\geq 1} \}$ on which X acts by $v_i.X = v_{i+n}$ . If $n \geq d$ , then the assumptions in Lemma 7.4.2 hold. In this case, $U_j := v_j.B (1\leq j \leq d)$ are indeed distinct summands. One can extend the right A-action to the entire tensor space $T = V_B^{\otimes d}$ via Equation (7.5.1). Then $S^A(T)$ seems to be a new q-Schur algebra for the degenerate affine Hecke algebra.
See [Reference Brundan and KleshchevBK08] for a related cyclotomic version. Note that our approach does not recover the Schur duality functor as in [Reference Arakawa and SuzukiAS98] since their algebra $\mathcal {H}^{\mathrm {deg}}(\Sigma _d)$ acts on a larger module $M \otimes V^{\otimes d}$ .
7.8 Connections to quantum groups
In this article, we only discuss a Schur duality between a quantum wreath product and its corresponding Schur algebra; while in literature, there is usually a third object that surjects onto the Schur algebra. For example, it is the Drinfeld–Jimbo quantum group that surjects onto the q-Schur algebras of type A.
A standard procedure due to Beilinson–Lusztig–MacPherson can be used to produce these higher level algebras from the given families of Schur algebras. See [Reference Beilinson, Lusztig and MacPhersonBLM90] for constructing the quantum group $U_q(\mathfrak {gl_n})$ from $\mathcal {H}_q(\Sigma _d)$ , and [Reference Bao, Kujawa, Li and WangBKLW18, Reference Lai and LuoLL21] for a realization of equal or multiparameter $\imath $ quantum groups $\imath U(\mathfrak {gl}_n)$ from Hecke algebra of type B/C. Furthermore, [Reference Fan, Lai, Li, Luo and WangFL3W23] explains how to obtain the affine $\imath $ quantum groups of type C using affine Hecke algebra of type C. It is still open whether one can generalize this approach to obtain the quantum group analogs for other examples in Section 4, for example, the Ariki–Koike algebras, the Hu algebras or the degenerate affine Hecke algebras.
A Examples of quantum wreath product algebras
A.1 Affine Hecke algebras
We first consider the extended affine Hecke algebra $\mathcal {H}^{\mathrm {ext}}_q(\Sigma _d)$ for $\mathrm {GL}_d$ . Recall that $P = \sum _{i=1}^d \mathbb {Z} \epsilon _i$ is the weight lattice for $\mathrm {GL}_d$ , and $\alpha _i := \epsilon _i - \epsilon _{i+1}$ is our choice of positive simple roots. The algebra $\mathcal {H}^{\mathrm {ext}}_q(\Sigma _d)$ is generated by $T_1, \dots , T_{d-1}$ and $Y^\lambda (\lambda \in P)$ subject to the Bernstein–Lusztig relation, for $\lambda \in P, 1\leq i \leq d-1$ :
as well as the relations below:
We remark that the nonextended affine Hecke algebra $\mathcal {H}^{\mathrm {aff}}_q(\Sigma ) := \mathcal {H}_q(\Sigma ^{\mathrm {aff}}_d)$ of type A is the Hecke algebra for the Coxeter group $\Sigma ^{\mathrm {aff}}_d$ . Furthermore, $\mathcal {H}^{\mathrm {aff}}_q(\Sigma _d)$ can be embedded into $\mathcal {H}^{\mathrm {ext}}_q(\Sigma _d)$ via
A.2 Degenerate affine Hecke algebras
The degenerate affine Hecke algebra of type A is the algebra $\mathcal {H}^{\mathrm {deg}}(d)$ generated by $s_1, \dots , s_{d-1}$ and the polynomial algebra $K[x_1, \dots , x_d]$ subject to the relations below, for $1\leq k \leq d-2, 1\leq i \leq d-1, |j-i|\geq 2$ :
as well as the following cross relations:
A.3 Nil Hecke rings
Kostant–Kumar’s nil Hecke ring [KK86] (or, affine nil Hecke algebra) of type A is the algebra $N\mathcal {H}(d)$ generated by $\partial _1, \dots , \partial _{d-1}$ and the polynomial algebra $K[x_1, \dots , x_d]$ subject to the relations below, for $1\leq k \leq d-2, 1\leq i \leq d-1, |j-i|\geq 2$ :
A.4 Ariki–Koike algebras
The Ariki–Koike algebra (i.e., cyclotomic Hecke algebra), denoted by $\mathcal {H}_{m,d}$ , is a deformation of $G(m,1,d) = C_m \wr \Sigma _d$ of parameters $(q_1, \ldots , q_m, q)$ over a Laurent polynomial $F = K[q_1^{\pm 1} , \ldots , q_m^{\pm 1}, q^{\pm 1}]$ generated by $X, T_1, \ldots T_{d-1}$ subject to the type A braid relations for $T_1, \dots , T_{d-1}$ , as well as the the following relations, for $1\leq i < d$ :
Note that the specialization of $\mathcal {H}_{m,d}$ at $q_i = \xi ^i, q=1$ is the group algebra $KG(m,1,d)$ for $\xi $ a primitive mth root of unity.
A.5 Rosso-Savage algebras
A.5.1 Affine wreath product algebra
Wan–Wang’s wreath Hecke algebras is a special case of Savage’s affine wreath product algebra $\mathcal {A}_d(B)$ . Following [Reference SavageSa20], here B is an $\mathbb {N}$ -graded Frobenius superalgebra with homogeneous parity-preserving linear trace map $\mathop {\text {tr}}:B\to K$ . Denote its Nakayama automorphism by $\psi :B\to B$ , that is,
Assume that B has a basis I consisting of homogeneous elements, and let $\{x^\vee ~|~x \in I\}$ be its dual basis in the sense that $\mathop {\text {tr}}(x^\vee y) = \delta _{x,y}$ for all $x, y \in I$ .
For $x\in B$ , denote by $x^{(i)}$ the element in $B^{\otimes d}$ such that all tensor factors are identities except that the ith factor is x. Next, for $b = b_1 \otimes \dots \otimes b_d \in B^{\otimes d}$ , let $\psi _i(b) = \psi (b_i)^{(i)} \in B^{\otimes d}$ . Furthermore, define $t_{i,j} := \sum _{x \in I} x^{(i)} (x^\vee )^{(j)} \in B^{\otimes d}$ for all $i,j$ .
The algebra $\mathcal {A}_d(B)$ is generated by $b = b_1 \otimes \dots \otimes b_d \in B^{\otimes d}, L_1,\dots , L_d, T_1, \dots , T_{d-1}$ subject to the relations below, for $1\leq i \leq d-1, 1\leq j,k \leq d$ :
A.5.2 Quantum wreath algebra (Frobenius Hecke algebra)
Let $z\in K$ . Following [Reference Rosso and SavageRS20], the Frobenius Hecke algebra (or, quantum wreath algebra) $H_d(B,z)$ is generated by $b = b_1 \otimes \dots \otimes b_d \in B^{\otimes d}, T_1, \dots , T_{d-1}$ subject to the relations below, for $1\leq i, j \leq d-1$ :
A.5.3 Quantum affine wreath algebra (affine Frobenius Hecke algebra)
Let $z\in K$ . Following [Reference Rosso and SavageRS20], the affine Frobenius Hecke algebra (or, quantum affine wreath algebra) $H^{\mathrm {aff}}_d(B,z)$ is generated by $b = b_1 \otimes \dots \otimes b_d \in B^{\otimes d}, X^{\pm }_1,\dots , X^{\pm }_d, T_1, \dots , T_{d-1}$ subject to the relations below, for $1\leq i \leq d-1, 1\leq j,k \leq d$ :
A.6 Affine zigzag algebras
Let $\Gamma $ be a simply-laced connected Dynkin diagram, and let $\overline {\Gamma }$ be its corresponding double quiver with vertex set I and edge set $E = \{ a_{i,j} ~|~ i \neq j \in I\}$ . Let $Z_K = Z_K(\Gamma )$ be the zigzag algebra over K be the quotient of the path algebra $K \overline {\Gamma }$ subject to the following relations:
-
1. Any path of length $\geq 3$ is zero;
-
2. Any path of length $2$ is zero, unless it is a cycle;
-
3. All cycles of length 2 at the same node are equal.
In other words, $Z_K$ has a basis $\{e_i, a_{i,j}, c_i ~|~ j\neq i \in I\}$ such that $e_i$ ’s are the length-0 paths at i, $c_i$ ’s are the length-2 cycles at i (which are identified as long as they have the same base vertex). The identity in $Z_K$ is $1_Z = \sum _{i\in I} e_i$ . We can paraphrase the definition (see [Reference Kleshchev and MuthKM19]) of the affine zigzag algebra $Z^{\mathrm {aff}}_d(\Gamma )$ as follows. Let $Z^{\mathrm {aff}}_d(\Gamma )$ be generated by $Z_K^{\otimes d}, z_1, \dots z_d, T_1, \dots , T_{d-1}$ subject to the relations below, for $1\leq i \leq d-1, 1\leq j,k \leq d, \gamma \in Z_K^{\otimes d}$ :
where $e_\lambda = e_{\lambda _1} \otimes \dots \otimes e_{\lambda _d}$ for $\lambda = (\lambda _i)_i \in I^d$ and $\Sigma _d$ acts on $ Z_K^{\otimes d}$ by place permutations.
B Calculations for the basis theorem
B.1 Proof of Proposition 3.2.1
The proof will entail a term-by-term comparison via the basis $\{(b_1\otimes \dots \otimes b_d) H_w\}$ . For the proofs within this appendix, it is enough to check local relations in either $B\otimes B$ or $B^{\otimes 3}$ . The general proofs follow by using the embedding into $B^{\otimes d}$ for an arbitrary d.
Equation (3.2.1) follows from the fact that $H(1\otimes 1) = \sigma (1\otimes 1) H + \rho (1\otimes 1) = (1\otimes 1) H$ . Equations (3.2.2)–(3.2.3) follow from the fact that $H(ab) = \sigma (ab) H + \rho (ab)$ , while
and the fact that $(H^2)H = (SH+R)H = (S^2+ R)H +SR$ , while
Equation (3.2.4) follows from the fact that $H^2a = (SH+R)a = S(\sigma (a)H + \rho (a)) + Ra$ ; while $H(Ha)$ is equal to
Assume that $d\geq 3$ from now on. Equations (3.2.5)–(3.2.7) follow from combining
with a similar formula for $H_2H_1H_2a$ using symmetry. For Equations (3.2.8)–(3.2.9), one uses $H_1(H_1H_2H_1) = S_1 H_1H_2 H_1 + R_1 H_2 H_1$ , while
B.2 Proof of Proposition 5.2.1
In this section, we use the postfix notation for maps of the form $f_a$ and $T_i$ with $a\in B^{\otimes d}$ . It is the same postfix notation as what we used in Equation (5.1.2). Note that in the subscripts of the $f_a$ ’s, we still use the regular prefix composition and evaluation for $\sigma _i$ and $\rho _i$ .
In the following, we will use the fact that $f_a + f_b = f_{a+b}$ for all $a, b\in B^{\otimes d}$ (see Proposition 5.1.3) without referring to it. For simplicity, we abbreviate $1^{\otimes d}$ by $1$ . When we refer to Condition $R[\ell ]$ , we will use the following equivalent form:
Finally, it suffices to prove each part of Proposition 5.2.1 by considering the postfix evaluation at $1 \otimes w$ for every $w \in \Sigma _d$ with $\ell (w) \leq \ell $ .
Proof of Part (W).
First, consider the case when $\ell (w) < \ell (ws_i)$ . Then,
For the other case when $\ell (w)> \ell (ws_i)$ , note that $(1 \otimes w)\cdot T_i f_a = (1 \otimes ws_i) \cdot T_i^2f_a$ . As elements in $\mathrm {Hom}_{B^{\otimes d}}(V^{\ell -1}, V) $ ,
Therefore, $(1 \otimes w)\cdot T_i f_a = (1 \otimes ws_i) \cdot (T_i f_{\sigma _i(a)} T_i + T_i f_{\rho _i(a)}) = (1 \otimes w) \cdot (f_{\sigma _i(a)} T_i + f_{\rho _i(a)})$ .
Proof of Part (M).
The case when $w=1$ follows immediately from the construction. We may now assume that $w> 1$ . Then, there exists $s_i$ such that $\ell (w)> \ell (ws_i)$ . Note first $(1 \otimes w)\cdot f_{a} f_b = (1 \otimes ws_i) \cdot T_i f_a f_b$ and $(1 \otimes w)\cdot f_{ab} = (1 \otimes ws_i) \cdot T_i f_{ab}$ . We are done since the equalities below hold in $\mathrm {Hom}_{B^{\otimes d}}(V^{\ell -1}, V) $ :
Proof of Part (Q).
The case when $\ell (w) < \ell (w s_i)$ follows immediately from construction. For $\ell (w)> \ell (w s_i)$ , it suffices to show that $T_i^3 = T_i f_{S_i} T_i + T_if_{R_i}$ in $\mathrm {Hom}_{B^{\otimes d}}(V^{\ell -1}, V) $ . Indeed,
Proof of Part (B2).
We first consider w with $\ell (w) < \ell (w s_i)$ and $\ell (w) < \ell (w s_j)$ . Then
By symmetry in i and j, it remains to prove (B2) for the case $\ell (w)> \ell (w s_i)$ . Since $1 \otimes w = (1 \otimes w s_i) \cdot T_i$ , it suffices to prove in $\mathrm {Hom}_{B^{\otimes d}}(V^{\ell -1}, V) $ that $T_i T_j T_i = T_i^2 T_j$ .
Proof of Part (B3).
The proof for the case with $\ell (w) < \ell (ws_i)$ and $\ell (w) < \ell (w s_j)$ is almost identical to the proof of Part (B2), and hence is omitted. By symmetry in i and j, it suffices to prove (B3) for the case $\ell (w)> \ell (w s_i)$ . Since $1 \otimes w = (1 \otimes w s_i) \cdot T_i$ , it suffices to check $T_i T_j T_i T_j = T_i^2 T_j T_i$ in $\mathrm {Hom}_{B^{\otimes d}}(V^{\ell -1}, V) $ .
Proof of Part (R).
Every pair of reduced expressions of w differs by a chain of braid relations. Therefore, it suffices to prove that, for two reduced expressions $r, r'$ of w such that r and $r'$ differ by a single braid relation,
where $s_i$ (resp. $s_j$ ) is the last simple generator of r (resp. $r'$ ). Below, the left-hand side and the right-hand side of Equation (B.2.9) are denoted by left-hand side and right-hand side, respectively.
Equation (B.2.9) holds trivially if $s_i = s_j$ . Now, assume that $s_i \neq s_j$ , hence there exists $x \in \Sigma _d$ such that $w = x \beta _{i, j} = x \beta _{j, i}$ with $\ell (w) = \ell (x) + \ell (\beta _{i, j})$ , where
Note that in following, we must avoid using $W[\ell - 1]$ since otherwise a circular argument will occur in the grand loop argument. Consider first the case that $\ell (\beta _{i, j}) = 2$ .
Finally, consider the case that $\ell (\beta _{i, j}) = 3$ . Note that
and similarly,
The proposition follows by showing the following identity in $\mathrm {Hom}_{B^{\otimes d}}(V^{\ell - 3}, V)$ :
Acknowledgements
We acknowledge Valentin Buciumas, Jun Hu, Hankyung Ko, Eric Marberg, Andrew Mathas, Catharina Stroppel and Weiqiang Wang for insightful discussions regarding various details of the paper. We thank Alistair Savage for making us aware of the connections to the (affine) Frobenius Hecke algebras as well as for providing several suggestions on an early version of the paper. We thank Ben Elias for bringing to our attention connections with the Bergman diamond lemma for Hecke-type categories.
Competing interests
The authors have no competing interest to declare.
Financial support
Research of the first author was supported in part by MSTC grants 113-2628-M-001-011, 112-2628-M-001-003, 111-2628-M-001-007, 109-2115-M-001-011, and the National Center of Theoretical Sciences. Research of the second author was supported in part by NSF grant DMS-2101941. Research of the third author was supported by NSFC (12350710787) and NSFC (12471311).