Abstract

We study the (quantum) Schur algebras of type B/C corresponding to the Hecke algebras with unequal parameters. We prove that the Schur algebras afford a stabilization construction in the sense of Beilinson–Lusztig–MacPherson that constructs a multiparameter upgrade of the quantum symmetric pair coideal subalgebras of type AIII/AIV with no black nodes. We further obtain the canonical basis of the Schur/coideal subalgebras, at the specialization associated with any weight function. These bases are the counterparts of Lusztig’s bar-invariant basis for Hecke algebras with unequal parameters. In the appendix we provide an algebraic version of a type D Beilinson–Lusztig–MacPherson construction, which is first introduced by Fan–Li from a geometric viewpoint.

1 Introduction

1.1 Background

The quantum groups introduced by Drinfeld and Jimbo have played a central role in representation theory and many other branches of mathematics. Equally important are Lusztig’s modified (or idempotented) quantum groups (cf. [22]) that admit the canonical bases, which are analogs of the Kazhdan–Lusztig bases for the Hecke algebras. In [4], a geometric construction of the modified quantum group |$\dot{\textbf U} \ (\mathfrak{g}\mathfrak{l}_n)$| is given by Beilinson–Lusztig–MacPherson. Their construction is now referred as the BLM or stabilization construction after a stabilization property of the family of the (quantum) Schur algebras of type A. In this paper, by a (equal-parameter) stabilization construction of type |$X$| we mean a construction of an algebra |$\dot{\textbf{K}}^X_n$| over |${\mathbb{Z}}[v,v^{-1}]$| such that

  • (1) There is a family of quantum Schur algebras |$\textbf{S}^X_{n,d}$|⁠, which are the centralizing algebras to the action of the Hecke algebra |$\textbf{H}^X_d$| of type |$X_d$|⁠, for all |$n,d$|⁠;

  • (2) The family |$\{\textbf{S}^X_{n,d} \ | \ d\in{\mathbb{N}}\}$| admits a stabilization property, namely, the algebra |$\dot{\textbf{K}}^X_n = \mathop{\textrm{Stab}}\limits _{\infty \leftarrow d} \textbf{S}^X_{n,d}$| is well defined. As a consequence, there is a basis of |$\dot{\textbf{K}}^X_n$| that is compatible with the Kazhdan–Lusztig bases for |$\textbf{H}^X_d$|⁠, and the canonical bases of |$\textbf{S}^X_{n,d}$| for all |$d$|⁠.

The stabilization constructions have been developed for classical type and for certain affine type (see Table 1 for the references)—there are geometric approaches using partial flags and counting over finite fields developed, while there also are algebraic approaches in the framework of the Hecke algebras using combinatorics on Coxeter groups.

Table 1

Known BLM/stabilization constructions

TypeFinite AFinite B/CFinite DAffine AAffine C
Geometric[4][3][14][23][15]
Algebraic[10]??[11][16]
TypeFinite AFinite B/CFinite DAffine AAffine C
Geometric[4][3][14][23][15]
Algebraic[10]??[11][16]
Table 1

Known BLM/stabilization constructions

TypeFinite AFinite B/CFinite DAffine AAffine C
Geometric[4][3][14][23][15]
Algebraic[10]??[11][16]
TypeFinite AFinite B/CFinite DAffine AAffine C
Geometric[4][3][14][23][15]
Algebraic[10]??[11][16]

We remark that the algebraic approach for finite type B/C is more or less a special case for affine type C, while the algebraic approach for type D will be given in the appendix of this present paper.

The stabilization construction in general produces not the Drinfeld–Jimbo’s quantum groups but Letzter–Kolb’s quantum symmetric pairs (cf. [18, 21]). For example, the stabilization constructions of type A and B/C lead to the quantum symmetric pairs of type AIII/IV with no black nodes.

1.2 A new direction

A recent work by Bao–Wang–Watanabe brings to the authors’ attention that a multiparameter Schur duality (cf. [7]) plays a governing role among the Schur dualities of classical type. They also introduce a multiparameter upgrade of quantum symmetric pairs of type AIII/AIV with no black nodes.

While it is unclear how to proceed a geometric approach with unequal parameters since dimension counting does not make sense in an obvious way, an algebraic/combinatorial approach seems viable. The goal of this article is to provide a stabilization construction with respect to the Schur duality with unequal parameters in loc. cit. We show that the multiparameter stabilization algebras constructed are the coideal subalgebras appearing in the quantum symmetric pairs of type AIII/AIV with no black nodes. As an application, we construct, for the 1st time, the canonical bases for the type B/C Schur algebras with unequal parameters associated with any weight function, using Lusztig’s bar-invariant basis [24] with unequal parameters.

Table 2

Relation between Schur duality of type B/C at various specializations

|$\mathbb{S}^{\bullet }_{n,d} \curvearrowright \mathbb{V}^{\otimes d}\curvearrowleft \mathbb{H}_d$||$\textrm{over} \ {\mathbb{Z}}[u^{\pm 1}, v^{\pm 1}] $|
|$\Downarrow \ \textrm{specialization at} \ u = \textbf{v}^{\textbf{L}(s_0)}, v = \textbf{v}^{\textbf{L}(s_1)}$|
|$\mathbb{S}^{\bullet ,\textbf{L}}_{n,d} \curvearrowright \mathbb{V}_{\textbf{L}}^{\otimes d}\curvearrowleft \mathbb{H}^{\textbf{L}} _d$||$\textrm{over} \ {\mathbb{Z}}[\textbf{v}^{\pm \textbf{c}}]$|
|$\Downarrow \ \textrm{specialization at} \ u=v=\textbf{v} \ (\textrm{i.e.,} \ \textbf{L} = \ell )$|
|$\textbf{S}^{\bullet }_{n,d} \curvearrowright \textbf{V}^{\otimes d}\curvearrowleft \textbf{H}_d$||$\textrm{over} \ {\mathbb{Z}}[\textbf{v}^{\pm 1}]$|
|$\mathbb{S}^{\bullet }_{n,d} \curvearrowright \mathbb{V}^{\otimes d}\curvearrowleft \mathbb{H}_d$||$\textrm{over} \ {\mathbb{Z}}[u^{\pm 1}, v^{\pm 1}] $|
|$\Downarrow \ \textrm{specialization at} \ u = \textbf{v}^{\textbf{L}(s_0)}, v = \textbf{v}^{\textbf{L}(s_1)}$|
|$\mathbb{S}^{\bullet ,\textbf{L}}_{n,d} \curvearrowright \mathbb{V}_{\textbf{L}}^{\otimes d}\curvearrowleft \mathbb{H}^{\textbf{L}} _d$||$\textrm{over} \ {\mathbb{Z}}[\textbf{v}^{\pm \textbf{c}}]$|
|$\Downarrow \ \textrm{specialization at} \ u=v=\textbf{v} \ (\textrm{i.e.,} \ \textbf{L} = \ell )$|
|$\textbf{S}^{\bullet }_{n,d} \curvearrowright \textbf{V}^{\otimes d}\curvearrowleft \textbf{H}_d$||$\textrm{over} \ {\mathbb{Z}}[\textbf{v}^{\pm 1}]$|
Table 2

Relation between Schur duality of type B/C at various specializations

|$\mathbb{S}^{\bullet }_{n,d} \curvearrowright \mathbb{V}^{\otimes d}\curvearrowleft \mathbb{H}_d$||$\textrm{over} \ {\mathbb{Z}}[u^{\pm 1}, v^{\pm 1}] $|
|$\Downarrow \ \textrm{specialization at} \ u = \textbf{v}^{\textbf{L}(s_0)}, v = \textbf{v}^{\textbf{L}(s_1)}$|
|$\mathbb{S}^{\bullet ,\textbf{L}}_{n,d} \curvearrowright \mathbb{V}_{\textbf{L}}^{\otimes d}\curvearrowleft \mathbb{H}^{\textbf{L}} _d$||$\textrm{over} \ {\mathbb{Z}}[\textbf{v}^{\pm \textbf{c}}]$|
|$\Downarrow \ \textrm{specialization at} \ u=v=\textbf{v} \ (\textrm{i.e.,} \ \textbf{L} = \ell )$|
|$\textbf{S}^{\bullet }_{n,d} \curvearrowright \textbf{V}^{\otimes d}\curvearrowleft \textbf{H}_d$||$\textrm{over} \ {\mathbb{Z}}[\textbf{v}^{\pm 1}]$|
|$\mathbb{S}^{\bullet }_{n,d} \curvearrowright \mathbb{V}^{\otimes d}\curvearrowleft \mathbb{H}_d$||$\textrm{over} \ {\mathbb{Z}}[u^{\pm 1}, v^{\pm 1}] $|
|$\Downarrow \ \textrm{specialization at} \ u = \textbf{v}^{\textbf{L}(s_0)}, v = \textbf{v}^{\textbf{L}(s_1)}$|
|$\mathbb{S}^{\bullet ,\textbf{L}}_{n,d} \curvearrowright \mathbb{V}_{\textbf{L}}^{\otimes d}\curvearrowleft \mathbb{H}^{\textbf{L}} _d$||$\textrm{over} \ {\mathbb{Z}}[\textbf{v}^{\pm \textbf{c}}]$|
|$\Downarrow \ \textrm{specialization at} \ u=v=\textbf{v} \ (\textrm{i.e.,} \ \textbf{L} = \ell )$|
|$\textbf{S}^{\bullet }_{n,d} \curvearrowright \textbf{V}^{\otimes d}\curvearrowleft \textbf{H}_d$||$\textrm{over} \ {\mathbb{Z}}[\textbf{v}^{\pm 1}]$|

The following diagram explains briefly the connection between the stabilization construction of type B/C for equal and unequal parameters (here |$\textbf{c} = \gcd (\textbf{L} (s_0), \textbf{L} (s_1))$|⁠, and there are two distinct cases where |$\bullet $| can be replaced by |$\imath $| or |$\jmath $|⁠).

At the specialization |$u=1$|⁠, the Hecke algebra contains the type D Hecke algebra over |${\mathbb{Z}}(v^{\pm 1})$| as a proper subalgebra. Hence, the multiparameter Schur duality yields a weak Schur duality of type D that is used in [1] to formulate the Kazhdan–Lusztig theory for classical and super type D. The very duality also appears in [13] as a piece of a larger skew Howe duality of the quantum symmetric pair coideal subalgebra with itself.

1.3 Unequal parameters

While the organization of this paper follows closely to the (equal-parameter) affine-type C construction [16], the technical lemmas therein do not generalize naively. Below we mention some notable difficulties working with unequal parameters.

The 1st difficulty comes to dealing with the combinatorics of (type B/C) quantum numbers with two parameters. The key observation here is that the (equal-parameter) quantum numbers/factorials used in the BLM-type constructions arise from the (equal-parameter) Poincare polynomials corresponding to the Weyl groups. Hence, we compute the multiparameter upgrade for the type B/C Poincare polynomials (cf. Lemma 2.7), and then extract from it a type B/C quantum factorial (2.3.3) with two parameters.

The 2nd difficulty arises in constructing a standard basis of |${\mathbb{S}}^{\jmath }_{n,d}$|⁠. For the equal-parameter case such a basis element |$[A]$| is obtained by multiplying a |$\textbf{v}$|-power to the evident basis |$e_A$|⁠, while for unequal parameters, it is not obvious how to define a multiplier |$u^{\bullet } v^{\bullet }$| that specializes to the original |$\textbf{v}$|-power. We solve this problem by reducing it to getting an explicit formula (cf. Lemma 4.2) for the leading coefficient under the bar map. For the equal-parameter case the formula is obtained using certain identities on the dual Kazhdan–Lusztig basis due to Curtis. However, there are no multiparameter Kazhdan–Lusztig basis known to us (yet). Hence, we take a detour via Lusztig’s bar-invariant basis with unequal parameters and have successfully define a standard basis that affords the entire stabilization process.

Finally, we remark that there is an unexpected behavior for our multiparameter monomial bases—the basis elements are not bar-invariant, unlike the (equal-parameter) monomial basis elements. As a result, we can only show the existence of canonical bases for Schur algebras at certain specialization (see Section 4.4).

1.4 Organization and main results

Throughout the article the algebras are over the ground ring
(⁠|$u,v$| are independent indeterminants) and its specializations.

We first start with the case |$\bullet = \jmath $|⁠. In Section 2 we recall combinatorial properties of Weyl groups of type B/C in terms of permutation matrices. We characterize a matrix set |$\Xi _{n,d}$| (see (2.2.2)) associated with certain double coset representatives. We also introduce the multiparameter quantum numbers of type B/C corresponding to the Poincare polynomials. In Section 3 we introduce the Schur algebra |${\mathbb{S}}^{\jmath }_{n,d}$| (see (3.1.6)) with an evident basis |$\{e_A\mid A\in \Xi _{n,d}\}$|⁠. In Section 4 we introduce a standard basis |$\{[A] \mid A\in \Xi _{n,d}\}$| (see (4.2.3)), and we show that, using Lusztig’s basis for the Hecke algebras with unequal parameters, it satisfies a unitriangular condition under the bar involution. The 1st main result is the following multiparameter upgrade of the multiplication formulas in [3]:

 

Theorem A
(Theorem 4.5).

Let |$A, B \in \Xi _{n,d}$| and |$B - b(E_{h,h+1} +E_{-h,-h-1})$| is diagonal. Let |$\gamma _{B,A}^C \in{\mathbb{A}}$| be such that |$[B] [A] =\sum _C \gamma _{B,A}^C [C] \in{\mathbb{S}}^{\jmath }_{n,d}$|⁠. The explicit formula and the vanishing criterion for |$\gamma _{B,A}^C$| are computed.

The multiplication formula plays an essential step towards constructing a monomial basis in the sense that a stabilization property (4.3.3) holds.

 

Theorem B
(Proposition 4.3.1, Theorem 4.4.1).

There exists a monomial basis |$\{m_A\}$| for the Schur algebra |${\mathbb{S}}^{\jmath }_{n,d}$| over |${\mathbb{A}}$|⁠. Consequently, at a specialization associated with a weight function |$\textbf{L}$|⁠, there exists a canonical basis |$\{\{A\}^{\textbf{L}} \}$| for |$\mathbb{S}^{\jmath , \textbf{L}}_{n,d}$|⁠.

In Section 5 we show that the stabilization procedure along the line of Beilinson–Lusztig–MacPherson applies to the family of Schur algebras |$\{{\mathbb{S}}^{\jmath }_{n,d}\mid d \geqslant 1\}$| with a fixed |$n$|⁠, which leads to the construction of stabilization algebra |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| (cf. Corollary 5.1.3) together with its canonical basis.

 

Theorem C
(Theorem 5.2.2).

There exists a monomial basis |$\{m_A\}$| for the stabilization algebra |$\dot{{\mathbb{K}}}^{\jmath }_{n}$|⁠. As a corollary, there exists a canonical basis |$\{\{A\}^{\textbf{L}} \}$| for |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| at a specialization associated with a weight function |$\textbf{L}$|⁠.

Section 6 is dedicated to the counterparts of Theorems B and C for the case |$\bullet = \imath $| (see Theorems 6.2.2 and 6.3.8). In Section 7 we show that the stabilization algebras coincide with the |$\mathfrak{g}\mathfrak{l}$|-variants |${\mathbb{U}}^{\jmath }, {\mathbb{U}}^{\imath }$| of the multiparameter quantum symmetric pair coideal subalgebras studied by Bao–Wang–Watanabe in [7] (referred as |$\textbf{U}^{\jmath }, \textbf{U}^{\imath }$| therein). The argument is made by passing the idempotented (or modified) quantum algebras.

 

Theorem D
(Theorems 7.2.1 and 7.3.2).

There are |${\mathbb{Q}}(u,v)$|-algebra isomorphisms |$_{\mathbb{Q}}\dot{{\mathbb{K}}}^{\jmath }_{n} \simeq \dot{{\mathbb{U}}}^{\jmath }, {}_{\mathbb{Q}}\dot{{\mathbb{K}}}^{\imath }_{n} \simeq \dot{{\mathbb{U}}}^{\imath } $|⁠, where |${}_{\mathbb{Q}} \dot{\mathbb{K}}_n^{\bullet }={\mathbb{Q}}(u,v)\otimes _{{\mathbb{A}}}\dot{\mathbb{K}}_n^{\bullet }$|⁠.

In the appendix we provide an algebraic version of a type D Beilinson–Lusztig–MacPherson construction that is first introduced by Fan–Li from a geometric viewpoint.

2 Combinatorics on Weyl Groups

2.1 Weyl groups as permutation groups

Let |${\mathbb{N}} = \{0,1,2,\ldots \}$|⁠. Fix |$N,n,D,d\in{\mathbb{N}}$| such that
(2.1.1)
Let |$\textrm{Perm}(X)$| be the group of permutations on a set |$X$|⁠. Let |$(W,S)$| be the Coxeter system of type B/C by
(2.1.2)
where
(2.1.3)
In particular, |$g(0)=0$| for any |$g\in W$|⁠. The corresponding Coxeter diagram is as below:
Since that any |$g \in W$| is uniquely determined by |$(g(1), \ldots , g(d))$|⁠, we use the two-line/one-line notations (referred as the window notation in [2])
(2.1.4)

Let |$\ell :W\to{\mathbb{N}}$| be the length function on |$W$|⁠. We introduce a truncated length function |$\ell _{\mathfrak{c}}:W \to{\mathbb{N}}$| such that |$\ell _{\mathfrak{c}}(g)$| equals to the total number of |$s_0$|’s in a reduced expression of |$g$|⁠. The function |$\ell _{\mathfrak{c}}$| is well defined since it is the weight function (cf. [24]) determined by |$\ell _{\mathfrak{c}}(s_0) = 1, \ell _{\mathfrak{c}}(s_i) = 0$| for |$i\geqslant 1$|⁠. We set |$\ell _{\mathfrak{a}} = \ell - \ell _{\mathfrak{c}}$|⁠.

 

Lemma 2.1.1
For |$g \in W$|⁠, we have
(2.1.5)
 
(2.1.6)
 
(2.1.7)

 

Proof.

It follows by an easy induction that |$\ell _{\mathfrak{c}}(g) = {}^{\sharp }\{ i\in [1, d] | g(i) < 0\}$|⁠, which yields to (2.1.5) by a direct calculation. The formula (2.1.7) for |$\ell (g)$| is equivalent to the formula [2, (8.2)]. Then there comes the formula (2.1.6) by |$\ell _{\mathfrak{a}}(g)=\ell (g)-\ell _{\mathfrak{c}}(g)$|⁠.

 

Remark 2.1.2

The expressions in Lemma 2.1.1 are not the most straightforward. There are simpler ones, for example, |$\ell _{\mathfrak{a}} = \textrm{inv} +\textrm{neg}$| and |$\ell _{\mathfrak{c}} = \textrm{neg}$| following the convention in [2]. We will see in Lemma 2.2.2 the advantage of choosing such symmetrized expressions. See also [16, Appendix A] for similar symmetrized length formulas for finite and affine classical types.

Denote the set of weak compositions of |$d$| of |$n+1$| parts by
(2.1.8)
For any |$\lambda \in \Lambda _{n,d}$| and integer |$i\in [-n,n]$|⁠, we define integer intervals |$R^{\lambda }_i$| by
(2.1.9)
For any subset |$X \subset [-d,d]$|⁠, let |$\textrm{Stab}(X)$| be the stabilizer of |$X$| in |$W$|⁠. A parabolic subgroup of |$W$| must be of the form
(2.1.10)
Precisely, |$W_\lambda $| is the parabolic subgroup of |$W$| generated by |$S- \{s_{\lambda _0}, s_{\lambda _0+\lambda _1}, \ldots , s_{d-\lambda _n}\}$|⁠. Denote the set of shortest right coset representatives for |$W_\lambda \setminus W$| by
(2.1.11)
 
(2.1.12)
Denote the set of minimal length double coset representatives for |$W_\lambda \backslash W /W_\mu $| by
(2.1.13)

In the following we collect some standard results for Coxeter groups from [10, Proposition 4.16, Lemma 4.17, and Theorem 4.18].

 

Lemma 2.1.3.

Let |$\lambda ,\mu \in \Lambda _{n,d}$| and |$g \in \mathscr{D}_{\lambda \mu }$|⁠.

  • (a) There exists |$\delta \in \Lambda _{n^{\prime },d}$| for some |$n^{\prime }$| such that |$W_{\delta } = g^{-1} W_\lambda g \cap W_\mu .$|

  • (b) The map |$W_\lambda \times (\mathscr{D}_\delta \cap W_\mu ) \rightarrow W_\lambda g W_\mu $| sending |$(x,y)$| to |$xgy$| is a bijection; moreover, we have |$\ell (xgy) = \ell (x) + \ell (g) + \ell (y)$|⁠.

  • (c) The map |$W_\delta \times (\mathscr{D}_\delta \cap W_\mu ) \rightarrow W_\mu $| sending |$(x,y)$| to |$xy$| is a bijection; moreover, we have |$\ell (x) + \ell (y) = \ell (xy)$|⁠.

An essential step in deriving the multiplication formula is to understand the set |$\mathscr{D}_\delta \cap W_\mu $|⁠, which we will see in Section 3.2.

2.2 Set-valued matrices

Let
(2.2.1)
Note that the columns/rows of such a matrix are indexed by |$[-n, n]$| instead of |$[1, N]$|⁠. Let
(2.2.2)

For |$A=(a_{ij}) \in \Xi _{n,d}$| we define a matrix |$A^{\mathcal{P}} = (A^{\mathcal{P}}_{ij})$| to be the unique set-valued matrix satisfying the following:

  • (P0) The sets |$(A^{\mathcal{P}}_{ij})_{ij}$| partition |$[-d,d]$|⁠;

  • (P1) |$|A^{\mathcal{P}}_{ij}| = a_{ij}$| for all |$i,j$|⁠;

  • (P2) Every element in |$A^{\mathcal{P}}_{ij}$| is smaller than any element in |$A^{\mathcal{P}}_{xy}$| if |$(i,j) < (x,y)$| in the lexicographical order (i.e., |$(i,j) < (x,y)$| if and only if |$i<x$| or |$(i=x, j < y)$|⁠).

In words, the set-valued matrix |$A^{\mathcal{P}}$| is obtained by filling integers from |$-d$| to |$d$| into the entries |$A^{\mathcal{P}}_{ij}$| row-by-row, top-to-bottom. For |$T \in \Theta _N$|⁠, we define its row sum vector |$\textrm{row}(T)= (\textrm{row}(T)_k)_{k=-n}^n$| and column sum vector |$\textrm{col}(T)= (\textrm{col}(T)_k)_{k=-n}^n$| by
(2.2.3)

 

Lemma 2.2.1.
The following map is bijective:
(2.2.4)
Moreover, the inverse is given by |$\kappa ^{-1}(A) = (\textrm{row}(A), g_A, \textrm{col}(A))$|⁠, where |$g_A$| is the permutation sending |$k$| to the |$k$|-th number in the column-reading of |$A^{\mathcal{P}}$| (see Example 2.2.3 below).

 
Proof.

The surjectivity follows from |$\kappa (\textrm{row}(A), g_A, \textrm{col}(A))=A\ (\forall A\in \Xi _{n,d})$| by a direct calculation.

For injectivity, we assume |$\kappa (\lambda ,g,\mu ) = A = \kappa (\lambda ^{\prime },g^{\prime },\mu ^{\prime })$|⁠. Then |$\lambda =\lambda ^{\prime } = \textrm{row}(A)$| and |$\mu = \mu ^{\prime } = \textrm{col}(A)$| and hence |$g,g^{\prime }\in \mathscr{D}_{\lambda \mu }$|⁠. It follows from |$|R^{\lambda }_i \cap g R^{\mu }_j|=|R^{\lambda }_i \cap g^{\prime } R^{\mu }_j|\ (\forall i,j\in [-n,n])$| that |$g = w_{(\lambda )} g^{\prime } w_{(\mu )}$| for some |$w_{(\lambda )} \in W_{\lambda }, w_{(\mu )} \in W_{\mu }$|⁠. Therefore, |$g = g^{\prime }$| since they are both minimal double coset representatives in |$W_\lambda \backslash W / W_\mu $|⁠.

Thanks to Lemma 2.2.1, we define length functions |$\ell , \ell _{\mathfrak{c}}, \ell _{\mathfrak{a}}$| on |$\Xi _{n,d}$| by
(2.2.5)
We define index subsets of type A/C by the following:
(2.2.6)
For |$(i,j) \in I_{\mathfrak{c}}$|⁠, we set
(2.2.7)
There is an alternative length formula in terms of products of matrix entries as below.

 

Lemma 2.2.2.
Recall |$a_{ij}^{\natural }$| from (2.2.7). The (truncated) length functions of |$A$| are given by
(2.2.8)
 
(2.2.9)
where |$a_{00}^{\natural \natural } =a_{00}^{\natural }-1= \frac{1}{2}(a_{00}-3)$| and |$a_{ij}^{\natural \natural } =a_{ij}$| if |$(i,j) \in I_{\mathfrak{a}}$|⁠.

 

Proof.

These three formulas are paraphrases of those in Lemma 2.1.1.

Let |$A=\kappa (\lambda ,g,\mu )\in \Xi _{n,d}$|⁠. We define a signed weak composition as below:
(2.2.10)

A direct computation shows that |$\delta (A)$| is indeed a weak composition |$\delta $| in Lemma 2.1.3(a).

 

Example 2.2.3.
Let
$$A = \left [\begin{smallmatrix} 1&3&1\\1&1&1\\1&3&1 \end{smallmatrix}\right ]$$
. We have
Column-reading of |$A^{\mathcal{P}}$| gives us a sequence |$-6,-1,2,-5,-4,-3,0,3,4,5,-2,1,6$|⁠, and hence |$g_A$| is the permutation
Indeed, we have
Furthermore, |$\delta (A)=(1,1,1,3,0,3,1,1,1)$|⁠.

2.3 Quantum combinatorics

We denote the quantum |$v$|-number by
(2.3.1)
We denote the type A quantum |$v$|-factorials by, for |$t \in{\mathbb{N}}$|  |$, A = (a_{ij}) \in \Theta _N$|⁠,
(2.3.2)
The type B/C analogues are defined by, for |$t\in{\mathbb{N}}, A = (a_{ij}), B = (b_{ij})\in \Xi _n$|⁠,
(2.3.3)
In particular, the specialization of |$[2t]_{\mathfrak{c}}$| at |$u=v$| is |$[t](1+v^{2t}) = [2t]$|⁠. Furthermore, we set, for any |$a\in \mathbb{Z}$| and |$b\in \mathbb{N}$|⁠,

 

Lemma 2.3.1.

Let |$A = \kappa (\mu ,g,\nu )$|⁠, and let |$\delta = \delta (A)$|⁠. Then |$\sum \limits _{w \in W_{\delta }} u^{2\ell _{\mathfrak{c}}(w)}v^{2\ell _{\mathfrak{a}}(w)} = [A]_{\mathfrak{c}}^{!}$|⁠.

 

Proof.

Let |$W^{\mathfrak{c}}_d$| be the Weyl group of type C|$_d$|⁠.

Recall |$\delta $| in (2.2.10). We have |$W_\delta \simeq W^{\mathfrak{c}}_{a^{\natural }_{00}}\times \prod _{(i,j)\in I_{\mathfrak{a}}} \ {\mathfrak{S}}_{a_{ij}}$|⁠. For each |$w\in{\mathfrak{S}}_{a_{ij}}$| we have |$\ell _{\mathfrak{c}}(w) = 0, \ell _{\mathfrak{a}}(w) = \ell (w)$|⁠, and hence
(2.3.4)
Thus,
It suffices to show that
(2.3.5)
Let |$\lambda = (0, \ldots , 0, 1, 2d-1,1,0,\ldots ,0) \in \Lambda _{n,d}$|⁠. We have |$W_\lambda \simeq W^{\mathfrak{c}}_{d-1}$|⁠, and hence
(2.3.6)
By (2.1.11), |$g \in \mathscr{D}_\lambda $| if and only if |$g^{-1}$| is order-preserving on |$[-d+1, d-1]$|⁠. Hence,
(2.3.7)
Consequently, we have
(2.3.8)
Therefore, (2.3.5) follows from a downward iteration. The lemma is proved.

3 Schur Algebras

3.1 Schur algebras

The Hecke algebra |${\mathbb{H}} = {\mathbb{H}}(W)$| over |${\mathbb{A}}$| is an algebra with a basis |$\{T_g | g\in W\}$| satisfying
(3.1.1)
 
(3.1.2)
 
(3.1.3)
For any subset |$X \subset W$| and for |$\lambda \in \Lambda _{n,d}$| (2.1.8), set
(3.1.4)
where |$\mathbb{1}$| is the identity element of |$W$|⁠.

 

Lemma 3.1.1.

If |$w \in W_\lambda $|⁠, then |$T_w x_\lambda = u^{2\ell _{\mathfrak{c}}(w)}v^{2\ell _{\mathfrak{a}}(w)} x_\lambda = x_\lambda T_w$|⁠.

 
Proof.

This reduces to the case when |$w = s \in S$|⁠. It then follows from the Hecke relation (3.1.1).

For |$\lambda ,\mu \in \Lambda _{n,d}$|⁠, and |$g\in \mathscr{D}_{\lambda \mu }$|⁠, we consider a right |${\mathbb{H}}$|-linear map |$\phi _{\lambda \mu }^g \in \textrm{Hom}_{\mathbb{H}}(x_\mu{\mathbb{H}}, {\mathbb{H}})$|⁠, sending |$x_\mu $| to |$T^g_{\lambda \mu }.$| Thanks to Lemma 2.1.3(b), we have |$T^g_{\lambda \mu } = x_\lambda T_g T_{\mathscr{D}_\delta \cap W_\mu }$| for some |$\delta \in \Lambda _{n^{\prime},d}$|⁠, and hence we have constructed a right |${\mathbb{H}}$|-linear map
(3.1.5)
The Schur algebra  |${\mathbb{S}}^{\jmath }_{n,d}$| is defined as the following |${\mathbb{A}}$|-algebra:
(3.1.6)
Thanks to Lemma 2.2.1, for |$A = \kappa (\lambda ,g,\mu )$| we define
(3.1.7)
A formal argument as in [12, 17] is applicable to our setting and gives us the following.

 

Lemma 3.2.

The set |$\{ e_A | A\in \Xi _{n,d} \}$| forms an |${\mathbb{A}}$|-basis of |${\mathbb{S}}^{\jmath }_{n,d}$|⁠.

For |$T = (t_{ij})\in \Theta _N$|⁠, let |$\textrm{diag}(T) = (\delta _{ij} t_{ij}) \in \Theta _N$| and denote its centro-symmetrizer by
(3.1.8)
We remark that |$T^{\theta } \not \in \Xi _{n}$| since |$t^{\theta }_{00}$| is even. A matrix |$B \in \Xi _{n,d}$| is called a Chevalley matrix if
(3.1.9)
An easy consequence of Lemma 2.2.2 is that |$g_B =\mathbb{1}$| if |$B$| is Chevalley. We assume from now on that |$B$| is a Chevalley matrix, and we fix |$B = \kappa (\lambda , \mathbb{1}, \mu )$|⁠, |$A = \kappa (\mu , g, \nu )$|⁠. Recall |$[A]^{!}_{\mathfrak{c}}$| from (2.3.3). We have the following identity.

 

Lemma 3.1.3.

|$x_\mu T_{g} x_\nu = [A]^!_{\mathfrak{c}} \, e_A(x_\nu ).$|

 

Proof.
Let |$\delta = \delta (A)$|⁠. By Lemma 2.1.3(c), we have |$x_\nu = x_\delta T_{\mathscr{D}_\delta \cap W_\nu }$|⁠, and hence
(3.1.10)
By Lemma 2.1.3(a), |$w \in g^{-1} W_\mu g \cap W_\nu \subset W_\nu ,$| and hence |$T_gT_w = T_{gw}$| since |$g\in \mathscr{D}_{\mu \nu } \subset \mathscr{D}_\nu ^{-1}$|⁠. Moreover, we have |$gw = w^{\prime } g$| for some |$w^{\prime } \in W_\mu $|⁠. Since |$g\in \mathscr{D}_{\mu \nu } \subset \mathscr{D}_\mu $|⁠, we have
(3.1.11)
and therefore |$\ell (w^{\prime }) = \ell (w)$|⁠. Moreover, note that |$\ell _{\mathfrak{c}}$| is a well-defined weight function (cf. [24]) determined by |$\ell (s_0)=1$| and |$\ell (s_i)=0 \ (i\geqslant 1)$|⁠. Counting the number of |$s_0$| appeared in a reduced form of |$gw=w^{\prime }g$|⁠, we have |$\ell _{\mathfrak{c}}(gw)=\ell _{\mathfrak{c}}(g)+\ell _{\mathfrak{c}}(w)$| and |$\ell _{\mathfrak{c}}(w^{\prime }g) = \ell _{\mathfrak{c}}(w^{\prime }) + \ell _{\mathfrak{c}}(g)$| by (3.1.11). Thus, |$\ell _{\mathfrak{c}}(w)=\ell _{\mathfrak{c}}(w^{\prime })$| (and hence |$\ell _{\mathfrak{a}}(w)=\ell _{\mathfrak{a}}(w^{\prime })$|⁠). Finally, we have
(3.1.12)
where the 2nd equality follows from Lemma 3.1.11, while the 3rd equality follows from Lemma 2.3.1. The rest follows by the definition |$e_A(x_\nu )=x_\mu T_g T_{\mathscr{D}_\delta \cap W_\nu }$|⁠.

3.2 Multiplication formulas |$\mathscr{D}_\delta \cap W_\mu $|

 

Lemma 3.2.1.
Fix |$B = \kappa (\lambda , \mathbb{1}, \mu )$|⁠, |$A = \kappa (\mu , g, \nu ),$| and let |$\delta = \delta (B)$|⁠. Let |$y^w$| be the shortest double coset representative for |$W_\lambda wg W_\nu $|⁠, and set |$A^w = \kappa (\lambda ,y^w,\nu )$|⁠. Then
(3.2.1)

 

Proof.
By Lemma 3.1.3 and (3.1.5) (which implies |$e_B(x_\mu )=x_\lambda T_{\mathscr{D}_\delta \cap W_\mu }$|⁠) we see that
(3.2.2)
Since |$g \in \mathscr{D}_{\mu \nu } \subset \mathscr{D}_\mu $|⁠, so |$T_w T_g = T_{wg}$| for all |$w \in \mathscr{D}_\delta \cap W_\mu \subset W_\mu $|⁠. For |$w \in \mathscr{D}_\delta \cap W_\mu $|⁠, there exists |$w_\lambda \in W_\lambda , w_\nu \in W_\nu $| such that |$wg = w_\lambda y^w w_\nu $|⁠. Moreover, we have
(3.2.3)
Thus, we have
(3.2.4)
Combining the (3.2.2), (3.2.4) and applying Lemma 3.1.3 on |$x_\lambda T_{y^w} x_\nu $|⁠, we have
(3.2.5)

The lemma follows from (3.2.3).

 

Proposition 3.2.2.

Suppose that |$A, B, C\in \Xi _{n,d}$| and |$h\in [1,n]$|⁠.

  • (1) If |$B-bE_{h,h-1}^{\theta }$| is diagonal, |$\textrm{col}(B)=\textrm{row}(A)$|⁠, then
    (3.2.6)
    where |$t=(t_i)_{-n\leqslant i\leqslant n}\in{\mathbb{N}}^N$| with |$\sum _{i=-n}^n t_i=b$| such that
    (3.2.7)
    and
  • (2) Suppose |$C-ce:{h-1,h}^{\theta }$| is diagonal and |$\textrm{col}(C)=\textrm{row}(A)$|⁠. If |$h\neq 1$|⁠, then
    (3.2.8)
    where |$t=(t_i)_{-n\leqslant i\leqslant n}\in{\mathbb{N}}^N$| with |$\sum _{i=-n}^n t_i=c$| such that |$t_i\leqslant a_{h,i}$|⁠, and
    If |$h=1$|⁠, then
    (3.2.9)
    where |$t=(t_i)_{-n\leqslant i\leqslant n}\in{\mathbb{N}}^N$| with |$\sum _{i=-n}^n t_i=c$| such that |$t_i\leqslant a_{1,i}$|⁠.

 

Proof.
For Part (1), we only present the proof for the most complicated case |$h=1$|⁠. Let |$\delta =\delta (B)$| and take any |$t=(t_i)_{-n\leqslant i\leqslant n}\in{\mathbb{N}}^N$| as in the assumptions. Among those |$w\in \mathscr{D}_\delta \cap W_\mu $| such that |$A^{w}={A}_{t,1}$|⁠, there is a unique shortest element |$w_t$| with
(3.2.10)
In particular, we have
(3.2.11)
By a combinatorial argument, we calculate that
Note that
where (⁠|$\diamondsuit $|⁠) is due to the quantum binomial theorem |$\sum _{x=0}^m {m \brack x} v^{x(x-1)}z^x = \prod _{i=0}^{m-1}(1+v^{2i}z).$| Therefore,
(3.2.12)
Furthermore, it follows from Lemma 2.2.2 that
(3.2.13)
 
(3.2.14)
Part (1) then follows from combining (3.2.1), (3.2.11)–(3.2.14). For Part (2), we only present a proof for the most complicated case that |$h=1$|⁠. Let |$\delta =\delta (C)$| and take any |$t=(t_i)_{-n\leqslant i\leqslant n}\in{\mathbb{N}}^N$| as in the assumptions. Among those |$w\in \mathscr{D}_{\delta } \cap W_{\mu }$| such that |$A^{w}=\widehat{A}_{t,1}$|⁠, there is a shortest element |$w_t$| with
(3.2.15)
Direct computation yields to the following identities:
(3.2.16)
 
(3.2.17)
 
(3.2.18)
Part (2) then follows from combining (3.2.1), (3.2.15)–(3.2.18).

 

Remark 3.2.3.

These explicit formulas match the ones in [3] (resp. the unsigned ones in [14]) if we specialize |$u=v$| (resp. |$u=1$|⁠).

4 Canonical Bases

4.1 The bar involution

There is an |${\mathbb{A}}$|-algebra involution |$\bar{ }:{\mathbb{H}} \rightarrow{\mathbb{H}}$|⁠, which sends |$u\mapsto u^{-1}, v \mapsto v^{-1}, T_w \mapsto T_{w^{-1}}^{-1}$|⁠, for all |$w\in W$|⁠. In particular, we have, for |$s \in S - \{s_0\}$|⁠,
(4.1.1)

For |$\lambda ,\mu \in \Lambda _{n,d}$| (see (2.1.8)), let |$g^+_{\lambda \mu }$| be the longest element in the double coset |$W_\lambda g W_\mu $| for |$g \in \mathscr{D}_{\lambda \mu }$|⁠, and let |$w_\circ ^\mu = \mathbb{1}_{\mu \mu }^+$| be the longest element in the parabolic subgroup |$W_\mu = W_\mu \mathbb{1} W_\mu $|⁠. The lemma below is standard (cf. [10, Corollary 4.19]).

 

Lemma 4.1.1.

Let |$A = \kappa (\lambda ,g,\mu )$|⁠, |$\delta = \delta (A)$|⁠. Then

  • (a) |$g_{\lambda \mu }^+ = w_\circ ^{\lambda } g w_\circ ^{\delta } w_\circ ^{\mu }$|⁠, and |$\ell (g_{\lambda \mu }^+) = \ell (w_\circ ^{\lambda }) + \ell (g) - \ell (w_\circ ^{\delta }) + \ell (w_\circ ^{\mu }).$|

  • (b) |$W_\lambda g W_\mu = \{w \in W | g \leqslant w \leqslant g^+_{\lambda \mu }\}$|⁠.

Following [19], denote by |$\{C^{\prime }_w\}$| the Kazhdan–Lusztig |${\mathbb{Z}}[v, v^{-1}]$|-basis of the Hecke algebra |${\mathbb{H}}|_{u=v}$| characterized by Conditions (C1)–(C2) below:

  • (C1) |$C^{\prime }_w$| is bar-invariant;

  • (C2) |$C^{\prime }_w = v^{-\ell (w)}\sum _{y \leqslant w} P_{yw}(v) T_y$|⁠.

Here |$\leqslant $| is the (strong) Bruhat order, and |$P_{yw}$| is the Kazhdan–Lusztig polynomial satisfying that |$P_{ww}=1$| and |$P_{yw}\in{\mathbb{Z}}[v^2]$| with |$ \deg _v P_{yw} \leqslant \ell (w)-\ell (y)-1$| for |$y<w$|⁠. Recall |$T^g_{\lambda \mu }$| from (3.1.4) and denote
(4.1.2)
Following [9], let |$\textbf{H}_{\lambda \mu }$| be the |${\mathbb{Z}}[v, v^{-1}]$|-submodule of |$\left .{\mathbb{H}}\right |{}_{u=v}$| with basis |$\{ T_{\lambda \mu }^g\}_{g \in \mathscr{D}_{\lambda \mu }}$|⁠. It is shown in loc. cit. that |$\{C^g_{\lambda \mu }\}_{g\in \mathscr{D}_{\lambda \mu }}$| also forms a bar-invariant basis of |$\textbf{H}_{\lambda \mu }$|⁠.
It is shown in [24, §5] that, for any weight function |$\textbf{L}:W \to{\mathbb{N}}$|⁠, there exists a bar-invariant basis |$\{C^{\textbf{L}} \ _w\}$| (referred as |$c_w$| therein) at the specialization |$u = \textbf{v}^{\textbf{L} (s_0)}, v = \textbf{v}^{\textbf{L} (s_1)}$|⁠, given by
(4.1.3)
where |$p_{y,w}(\textbf{v})$| is an analogue of Kazhdan–Lusztig polynomial. For |$\lambda ,\mu \in \Lambda _{n,d}$|⁠, let |${\mathbb{H}}_{\lambda \mu }$| be the |$ {\mathbb{Z}}[u^{\pm 2},v^{\pm 2}] $|-submodule of |${\mathbb{H}}$| with basis |$\{ T_{\lambda \mu }^g\}_{g \in \mathscr{D}_{\lambda \mu }}$|⁠. It follows from [8, Lemma 2.10] and Lemma 3.1.1 that |${\mathbb{H}}_{\lambda \mu }$| can be characterized as below:
(4.1.4)
Below we show that the bar involution is closed on |${\mathbb{H}}_{\lambda \mu }$| although lacking of bar-invariant basis.

 

Lemma 4.1.2.
Let |$A = \kappa (\lambda ,g,\mu )$|⁠. Then |$\overline{T^g_{\lambda \mu }} \in{\mathbb{H}}_{\lambda \mu }$|⁠. In particular,
(4.1.5)
Moreover, |$u^{-\ell _{\mathfrak{c}}(w_\circ ^{\mu })} v^{-\ell _{\mathfrak{a}}(w_\circ ^\mu )} x_\mu $| is bar-invariant.

 

Proof.
First, we show that |$\overline{x_\nu } \in{\mathbb{A}} x_\nu $| for all |$\nu \in \Lambda _{n,d}$| via bar-invariant basis |$C^{\textbf{L}} \ _w$|⁠. Let |${\mathbb{H}}^{\textbf{L}} \ _{\lambda \mu }$| be the specialization of |${\mathbb{H}}_{\lambda \mu }$| at |$u = \textbf{v}^{\textbf{L} (s_0)}, v = \textbf{v}^{\textbf{L} (s_1)}$|⁠. From (4.1.4), a direct calculation shows that |$C^{\textbf{L}} \ _{w_\circ ^{\nu }} \in{\mathbb{H}}^{\textbf{L}} \ _{\nu \nu }$|⁠, and hence
(4.1.6)
Upon comparing coefficients, we obtain
(4.1.7)
Note that |$x_\nu = T_{\nu \nu }^{\mathbb{1}}$|⁠. Hence, for any weight function |$\textbf{L} $|⁠, we have
(4.1.8)
Therefore, |$\overline{x_\nu } = u^{-2\ell _{\mathfrak{c}}(w_\circ ^{\nu })} v^{-2\ell _{\mathfrak{a}}(w_\circ ^{\nu })} x_\nu $|⁠. We now show that |$\overline{T_{\lambda \mu }^g} \in{\mathbb{H}}_{\lambda \mu }$|⁠. By Lemma 3.1.3, we have |$T_{\lambda \mu }^g \in{\mathbb{Z}}[u^{\pm 2},v^{\pm 2}] x_\lambda T_{g} x_\mu ,$| and hence
(4.1.9)
Similar to (3.2.3), we have |$x_\lambda T_z x_\mu \in{\mathbb{Z}}[u^{\pm 2},v^{\pm 2}] x_\lambda T_y x_\mu $| for some |$y \in \mathscr{D}_{\lambda \mu }$| such that |$y\leqslant z$|⁠. Finally, we have |$\overline{T_{\lambda \mu }^g} \in \sum _{y \in \mathscr{D}_{\lambda \mu }} {\mathbb{Z}}[u^{\pm 2},v^{\pm 2}] x_\lambda T_y x_\mu \subseteq{\mathbb{H}}_{\lambda \mu }$|⁠. The leading coefficient is obtained by a lengthy calculation, which we omit.
The bar involution |$\bar{ }$| on |${\mathbb{S}}^{\jmath }_{n,d}$| is defined as follows: for each |$f \in \textrm{Hom}_{{\mathbb{H}}}(x_\mu{\mathbb{H}}, x_\lambda{\mathbb{H}})$|⁠, let |$\overline{f}\in \textrm{Hom}_{{\mathbb{H}}}(x_\mu{\mathbb{H}}, x_\lambda{\mathbb{H}})$| be the |${\mathbb{H}}$|-linear map that sends |$x_\mu $| to |$\overline{f(\overline{x_\mu })}$|⁠.

4.2 A standard basis in |${\mathbb{S}}^{\jmath }_{n,d}$|

We define, for |$ A \in \Xi _{n,d}$|⁠, the (truncated) generalized length functions of |$A$| by
(4.2.1)
 
(4.2.2)
where |$a_{00}^{\natural \natural } = \frac{1}{2}(a_{00}-3)$| and |$a_{ij}^{\natural \natural } =a_{ij}$| if |$(i,j) \in I_{\mathfrak{a}}$|⁠. We shall see in Proposition 4.4 that |$\widehat{\ell }_{\mathfrak{a}}(A), \widehat{\ell }_{\mathfrak{c}}(A) \in{\mathbb{N}}$|⁠.
 
Remark 4.2.1

The function |$\widehat{\ell }$| counts the dimension of the generalized Schubert variety associated with the matrix |$A$| (cf. [16, Appendix A]) and is equal to the length of |$A$| when |$A$| is a permutation matrix (that is when the associated variety is a genuine Schubert variety).

Set
(4.2.3)
The set |$\{[A] | A \in \Xi _{n,d}\}$| forms an |${\mathbb{A}}$|-basis of |${\mathbb{S}}^{\jmath }_{n,d}$|⁠, which we call the standard basis.
For |$A\in \Xi _n$|⁠, we let
(4.2.4)
Now we define a partial order |$\leqslant _{\textrm{alg}}$| on |$\Xi _n$| by letting, for |$A, B \in \Xi _n$|⁠,
(4.2.5)
We denote |$A <_{\textrm{alg}} B$| if |$A \leqslant _{\textrm{alg}} B$| and |$A\neq B$|⁠.

 

Proposition 4.2.2

Let |$A =\kappa (\lambda ,g,\mu )\in \Xi _{n,d}$|⁠. Then we have |$\overline{[A]} \in [A] +\sum _{B <_{\textrm{alg}} A} {\mathbb{A}} [B].$|

 

Proof.
By the finite-type analogue of [16, Proposition 5.3], we have
(4.2.6)
Hence,
(4.2.7)
Thus, by Lemma 4.1.2, the map |$\overline{[A]}$| is determined by
(4.2.8)
We note that |$[\kappa (\lambda ,y,\mu )](x_\mu ) \in{\mathbb{A}} T^y_{\lambda \mu }$|⁠. An induction on |$\ell (g)$| shows that
(4.2.9)
A finite-type analogue of [16, Corollary 5.5] shows that |$\kappa (\lambda ,y,\mu ) <_{\textrm{alg}} A$| if |$y < g$|⁠. We conclude the statement.

Let us reformulate the multiplication formula for |${\mathbb{S}}^{\jmath }_{n,d}$| (Proposition 3.2.2) in terms of the standard basis.

 

Theorem 4.2.3

Suppose that |$A, B, C\in \Xi _{n,d}$|⁠, and |$h\in [1,n]$|⁠.

  • (1) If |$B-bE_{h,h-1}^{\theta }$| is diagonal, |$\textrm{col}(B)=\textrm{row}(A)$|⁠, then
    (4.2.10)
    where |$t$| is summed over as in Proposition 3.2.2 (1), and
    (4.2.11)
  • (2) Suppose |$C-ce:{h-1,h}^{\theta }$| is diagonal and |$\textrm{col}(C)=\textrm{row}(A)$|⁠. If |$h\neq 1$| then
    (4.2.12)
    where |$t$| is summed over as in Proposition 3.2.2 (2), and
    (4.2.13)
    If |$h=1$| then
    (4.2.14)
    where
    (4.2.15)

 

Proof.
For Part (1), by Proposition 3.2.2, we have
Part (1) concludes by combining the following identities via direct computation:
For Part (2), we only present the most complicated case that |$h=1$|⁠. A direct computation shows that
(4.2.16)
Part (2) follows from combining (4.2.16) and the calculation below:

4.3 A monomial basis in |${\mathbb{S}}^{\jmath }_{n,d}$|

Thanks to Remark 3.6, we can use results in [3] freely when we specialize |$u=v$|⁠. For |$A \in \Xi _{n,d}$|⁠, we can use the algorithm in [3, Theorem 3.10] with the fixed order therein to produce a unique family of Chevalley matrices |$\{A^{(1)}, \ldots , A^{(x)}\}$| in |$\Xi _{n,d}$| for some |$x = x(A) \in{\mathbb{N}}$|⁠. At the specialization |$u=v$|⁠, a unitriangular relation is satisfied:
(4.3.1)
Denote the product of the corresponding elements in |${\mathbb{S}}^{\jmath }_{n,d}$| by
(4.3.2)
Let |$I$| be the identity matrix. Since the algorithm in [3, Theorem 3.10] produces matrices |$A^{(1)}, \ldots , A^{(x)}$| according to mainly the off-diagonal matrices of |$A$| and then determine the diagonal entries of these |$A^{(i)}$| by the row and column sums, we have that |$x(A) = x(A+pI)$| and |$(A+pI)^{(i)} = A^{(i)}+ pI$| for all |$p\in 2{\mathbb{N}}$|⁠, that is,
(4.3.3)

 

Proposition 4.3.1
For |$A \in \Xi _{n,d}$| the element |$m_A \in{\mathbb{S}}^{\jmath }_{n,d}$| has the following property:
(4.3.4)
Moreover, |$\{m_A\}_{A\in \Xi _{n,d}}$| form a basis of |${\mathbb{S}}^{\jmath }_{n,d}$|⁠, which we call the monomial basis.

 

Proof.
A direct proof can be pursued using the multiplication formulas (Proposition 4.5), similar to the proofs of [3, Theorem 3.10] and [14, Theorem 4.6.3]. Here we offer a simpler proof by combining [3, Theorem 3.10] and [14, Theorem 4.6.3] as below: now
It follows from [3,Theorem 3.10] (resp. [14,Theorem 4.6.3]) that |$v^{\alpha (A)}v^{\beta (A)}=1$| (resp. |$1^{\alpha (A)}v^{\beta (A)}=1$|⁠), which forces that |$u^{\alpha (A)}v^{\beta (A)}=1$| and hence (4.3.4) holds. Hence, the transition matrix from |$\{m_A | A\in \Xi _{n,d}\}$| to the standard basis |$\{[A] | A \in \Xi _{n,d}\}$| is unital triangular. Therefore, |$\{m_A | A\in \Xi _{n,d}\}$| form a basis of |${\mathbb{S}}^{\jmath }_{n,d}$|⁠.

 

Remark 4.7.

The monomial basis acts as an intermediate step toward constructing canonical basis in the one-parameter case. Moreover, the two-parameter stabilization procedure is made possible thanks to the property (4.3.3) of monomial basis.

4.4 The canonical basis at the specialization

For any weight function |$\textbf{L} $|⁠, let |$\textbf{c} = \gcd (\textbf{L} (s_0), \textbf{L} (s_1))$|⁠. We show that the specialization of |${\mathbb{S}}^{\jmath }_{n,d}$| at |$u=\textbf{v}^{\textbf{L} (s_0)}, v=\textbf{v}^{\textbf{L} (s_1)}$| admits canonical basis with respect to |$\textbf{v}^{\textbf{c}}$|⁠. For |$A \in \Xi _{n,d}$|⁠, let |$[A]^{\textbf{L}} $| (and |$m_A^{\textbf{L}} $|⁠, resp.) be the standard basis (and monomial basis, resp.) of the specialization of |${\mathbb{S}}^{\jmath }_{n,d}$| at |$u=\textbf{v}^{\textbf{L} (s_0)}, v=\textbf{v}^{\textbf{L} (s_1)}$|⁠. It follows from (4.2.9) and (4.3.4) that the following unitriangular relations hold:
(4.4.1)
 
(4.4.2)
If |$A$| is diagonal, set |$\{A\}^{\textbf{L}} = [A]^{\textbf{L}}$|⁠. Arguing inductively on the partial order |$\leqslant _{\textrm{alg}}$| and using a standard argument (cf. [22, 24.2.1]) there exists a unique element |$\{A\}^{\textbf{L}} \in{\mathbb{S}}^{\jmath }_{n,d}$| such that
(4.4.3)

Let |$\mathbb{S}^{\jmath , \textbf{L}} _{n,d}$| be the specialization of |${\mathbb{S}}^{\jmath }_{n,d}$| at |$u=\textbf{v}^{\textbf{L} (s_0)}, v=\textbf{v}^{\textbf{L} (s_1)}$|⁠.

 

Theorem 4.4.1.

There exists a canonical basis |$\{\{A\}^{\textbf{L}} \ |\ A \in \Xi _{n,d}\}$| for |$\mathbb{S}^{\jmath , \textbf{L}} _{n,d}$|⁠, which is characterized by the property (4.4.3).

5 Stabilization Algebra |$\dot{{\mathbb{K}}}^{\jmath }_{n}$|

In this section, we shall establish a stabilization property for the family of Schur algebras |${\mathbb{S}}^{\jmath }_{n,d}$| as |$d$| varies, which leads to a quantum algebra |$\dot{{\mathbb{K}}}^{\jmath }_{n}$|⁠.

5.1 A BLM-type stabilization

Let
(5.1.1)
Extending the partial ordering |$\leqslant _{\textrm{alg}}$| for |$\Xi _{n}$|⁠, we define a partial ordering |$\leqslant _{\textrm{alg}}$| on |$\widetilde{\Xi }_n$| using the same recipe (4.2.5). For each |$A \in \widetilde{\Xi }_n$| and |$p \in 2{\mathbb{N}}$|⁠, we write
(5.1.2)
Then |${_{p}}{A} \in \Xi _n$| for even |$p \gg 0$|⁠. Let |$\pi $| be an indeterminate (independent of |$u,v$|⁠), and |$\mathcal{R}_1$| be the subring of |${\mathbb{Q}}(u,v)[\pi ,\pi ^{-1}]$| generated by, for |$a\in{\mathbb{Z}}, k\in{\mathbb{Z}}_{>0}$|⁠,
(5.1.3)
where
(5.1.4)
 
(5.1.5)
Let |$\mathcal{R}_2$| be the subring of |${\mathbb{Q}}(u,v)[\pi , \pi ^{-1}]$| generated by, for |$a\in{\mathbb{Z}}, k\in{\mathbb{Z}}_{>0}$|⁠,
(5.1.6)
where we extend the bar-involution to act on |$\pi $| by |$\overline{\pi }=\pi ^{-1}$|⁠.

 

Proposition 5.1.1.
Let |$A_1, \ldots , A_f \in \widetilde{\Xi }_n$| be such that |$\textrm{col}(A_i) = \textrm{row}(A_{i+1})$| for all |$i$|⁠. Then there exist matrices |$Z_1, \ldots , Z_m \in \widetilde{\Xi }_n$| and |$\zeta _i(u,v,\pi ) \in \mathcal{R}_1$| such that for even integer |$p\gg 0,$|  
(5.1.7)

 

Proof.
We assume first that |$f=2$| and |$A_1$| is such that |$A_1-bE_{h,h-1}^{\theta }$| is diagonal for some |$h\in [1,n]$| and some |$b\geqslant 0$|⁠. Let |$A_2=A=(a_{ij})$|⁠. For each |$t=(t_i)_{-n\leqslant i\leqslant n}\in{\mathbb{N}}^N$|⁠, we define
where |$\beta (t)$| is defined in (4.2.11). Though |$\beta (t)$| depends on |$A$|⁠, it is invariant if |$A$| is replaced by |${_{p}}{A}$|⁠. Therefore, we have the following formula for large enough even |$p$| by (4.2.10):
The statement holds in this case.
We next assume that |$f=2$| and |$A_1$| is such that |$A_1-ce:{h-1,h}^{\theta }$| is diagonal for some |$h\in [1,n]$| and some |$c\geqslant 0$|⁠. Let |$A_2=A=(a_{ij})$|⁠. Recall |$\beta ^{\prime}(t)$| and |$\beta ^{\prime \prime } (t)$| in (4.2.13) and (4.2.15), respectively. If |$h\neq 1$|⁠, for each |$t=(t_i)_{-n\leqslant i\leqslant n}\in{\mathbb{N}}^N$|⁠, we define
If |$h=1$|⁠, we define
It is clear that both |$\beta ^{\prime }(t)$| and |$\beta ^{\prime \prime }(t)$| are invariant if |$A$| is replaced by |${_{p}}{A}$|⁠. Therefore, the following formula holds for large enough even |$p$| by (4.2.12):
Hence, the proposition is verified in the present case.
Using induction on |$f$|⁠, we know that the proposition holds for general |$f$| in the case where |$A_1,\ldots ,A_f$| are Chevalley matrices (i.e., of one of the two types considered above). It follows from (4.3.2) and (4.3.4) that for any |$A\in \Xi _{n,d}$|⁠, there exists Chevalley matrices |$B_1, B_2,\ldots ,B_M$| such that
Then we can prove the proposition by using induction on |$\Psi (A)=\sum _{i<j}\sigma _{ij}(A)$|⁠. We omit the subsequent argument here since it is totally as the same as those for [4, Proposition 4.2].

By an argument identical with [4, Proposition 4.3], we obtain below the stabilization of bar involution by allowing extra coefficients as seen in (5.1.6).

 

Proposition 5.1.2.
For any |$A \in \widetilde{\Xi }_n$|⁠, there exist matrices |$T_1, \ldots , T_s\in \widetilde{\Xi }_n$| and |$\tau _i(u,v,\pi ) \in \mathcal{R}_2$| such that, for even integer |$p\gg 0,$|  
(5.1.8)

Let |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| be the free |${\mathbb{A}}$|-module with an |${\mathbb{A}}$|-basis given by the symbols |$[A]$| for |$A\in \widetilde{\Xi }_n$| (which will be called a standard basis of |$\dot{{\mathbb{K}}}^{\jmath }_{n}$|⁠). By Propositions 5.1 and -5.2 and applying a specialization at |$\pi =1$| (note that |$\zeta _i(u,v, 1), \tau _i(u,v, 1)\in{\mathbb{A}}$| because of |$r^{(1)}_{a,k}(u,v,1), r^{(2)}_{a,k}(u,v,1)\in{\mathbb{A}}$|⁠), we have the following corollary.

 

Corollary 5.1.3.
There is a unique associative |${\mathbb{A}}$|-algebra structure on |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| with multiplication given by
Moreover, the map |$\bar{ }: \dot{{\mathbb{K}}}^{\jmath }_{n} \rightarrow \dot{{\mathbb{K}}}^{\jmath }_{n}$| given by |$\overline{[A]} = \sum _{i=1}^s \tau _i(u,v, 1) [T_i]$| is an |${\mathbb{A}}$|-linear involution.

The following multiplication formula in |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| follows directly from Theorem 4.5 by the stabilization construction.

 

Proposition 5.1.4.

Let |$A, B, C \in \widetilde{\Xi }_{n}$|⁠, and |$h\in [1,n]$|⁠.

  • (1) If |$B-bE_{h,h-1}^{\theta }$| is diagonal and |$\textrm{col}(B)=\textrm{row}(A)$|⁠, then
    (5.1.9)
    where |$t=(t_i)_{-n\leqslant i\leqslant n}\in{\mathbb{N}}^N$| with |$\sum _{i=-n}^n t_i=b$| such that
  • (2) Suppose |$C-ce:{h-1,h}^{\theta }$| is diagonal and |$\textrm{col}(C)=\textrm{row}(A)$|⁠.

    If |$h\neq 1$| then
    (5.1.10)
    where |$t=(t_i)_{-n\leqslant i\leqslant n}\in{\mathbb{N}}^N$| with |$\sum _{i=-n}^n t_i=c$| such that |$t_i\leqslant a_{h,i}$| if |$i\neq h$|⁠.
    If |$h=1$| then
    (5.1.11)
    where |$t=(t_i)_{-n\leqslant i\leqslant n}\in{\mathbb{N}}^N$| with |$\sum _{i=-n}^n t_i=c$| such that |$t_i\leqslant a_{1,i}$| if |$i\neq 1$|⁠.

5.2 Monomial and canonical bases for |$\dot{{\mathbb{K}}}^{\jmath }_{n}$|

The proposition below follows from Proposition 4.3.1 by the stabilization construction.

 

Proposition 5.2.1.
For any |$A\in \widetilde{\Xi }_{n}$|⁠, there exist Chevalley matrices |$A^{(1)}, \ldots , A^{(x)}$| in |$\widetilde{\Xi }_{n}$| satisfying |$\textrm{row}(A^{(1)})=\textrm{row}(A)$|⁠, |$\textrm{col}(A^{(x)})=\textrm{col}(A)$|⁠, |$\textrm{col}(A^{(i)})=\textrm{row}(A^{(i+1)})$| for |$1\leqslant i\leqslant x-1$|  
(5.2.1)

By abuse of notation, we denote the product in |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| by
(5.2.2)
Hence, |$\{m_A | A\in \widetilde{\Xi }_{n}\}$| forms a basis for |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| (called a monomial basis). Similar to Section 4.4, we define, by abuse of notation, elements |$[A]^{\textbf{L}}, m_A^{\textbf{L}}, \{A\}^{\textbf{L}} $| to be the according basis elements of |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| at the specialization |$u = \textbf{v}^{\textbf{L} (s_0)}, v= \textbf{v}^{\textbf{L} (s_1)}$|⁠.

 

Theorem 5.2.2.

There exists a canonical basis |$\dot{{\mathfrak{B}}} = \{\{A\}^{\textbf{L}} \ |\ A \in \widetilde{\Xi}_{n}\}$| for |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| at the specialization |$u=\textbf{v}^{\textbf{L} (s_0)}, v=\textbf{v}^{\textbf{L} (s_1)}$|⁠, which is characterized by the property (4.4.3).

6 A Different Stabilization Algebra |$\dot{{\mathbb{K}}}^{\imath }_{n}$|

In this section we formulate a variant of Schur algebras and their corresponding stabilization algebras. We construct the distinguished bases of these algebras. Recall |$N=2n+1$|⁠.

6.1 |$\imath $|-Schur algebras

Recall |$\Xi _{n,d}$| from (2.2.2). Let
(6.1.1)
Recall |$\Lambda _{n,d}$| (2.1.8). Let
The lemma below is the |$\imath $|-analog of Lemma 2.2.1, which follows by a similar argument.
 
Lemma 6.1.1.

The map |$\kappa ^{\imath }: \bigsqcup _{\lambda ,\mu \in \Lambda _{n,d}^{\imath }} \{\lambda \} \times \mathscr{D}_{\lambda \mu } \times \{\mu \} \longrightarrow \Xi ^{\imath }$| sending |$(\lambda , g, \mu )$| to |$(|R_i^{\lambda } \cap g R_j^\mu |)$| is a bijection.

Now we define the |$\imath $|-Schur algebra as
(6.1.2)
By definition the algebra |${\mathbb{S}}^{\imath }_{n,d}$| is naturally a subalgebra of |${\mathbb{S}}^{\jmath }_{n,d}$|⁠. Moreover, both |$\{ e_A | A \in \Xi ^{\imath } \}$| and |$\{ [A] | A \in \Xi ^{\imath } \}$| are bases of |${\mathbb{S}}^{\imath }_{n,d}$| as a free |${\mathbb{A}}$|-module.

6.2 Monomial and canonical bases for |${\mathbb{S}}^{\imath }_{n,d}$|

 

Proposition 6.2.2.

For each |$A \in \Xi ^{\imath }$|⁠, we have |$m_A\in{\mathbb{S}}^{\imath }_{n,d}$|⁠. Hence, the set |$\{m_A | A \in \Xi ^{\imath } \}$| forms an |${\mathbb{A}}$|-basis of |${\mathbb{S}}^{\imath }_{n,d}$|⁠. Furthermore, we have |$m_A \in [A] +\sum _{B\in \Xi ^{\imath }, B<_{\textrm{alg}} A} {\mathbb{A}} [B]$|⁠.

 

Proof.

It follows from [3, Proposition 5.6] thanks to Remark 3.2.6.

 

Theorem 6.2.3.

At the specialization |$u = \textbf{v}^{\textbf{L}} (s_0), v= \textbf{v}^{\textbf{L}} (s_1)$|⁠, there is a canonical basis |$\mathfrak B_{n,d}^{\imath } = \{ \{A\}^{\textbf{L}} | A \in \Xi ^{\imath } \}$| of |${\mathbb{S}}^{\imath }_{n,d}$| such that |$\overline{\{A\}^{\textbf{L}}} \ =\{A\}^{\textbf{L}} $| and |$\{A\}^{\textbf{L}} \in [A]^{\textbf{L}} + \sum _{B\in \Xi ^{\imath }, B<_{\textrm{alg}} A} \textbf{v}^{-\textbf{c}} {\mathbb{Z}}[\textbf{v}^{-\textbf{c}}] [B]^{\textbf{L}} $|⁠. Moreover, we have |$\mathfrak B_{n,d}^{\imath } = \mathfrak B_{n,d}^{\jmath }\cap{\mathbb{S}}^{\imath }_{n,d}$|⁠.

 

Proof.

The 1st half statement on the canonical basis follows by Proposition 6.2 and a standard argument (cf. [22, 24.2.1]). The 2nd half statement follows from the uniqueness characterization of the canonical basis |$\mathfrak B_{n,d}^{\imath }$|⁠.

6.3 Stabilization algebra of type |$\imath $|

We define two subsets of |$\widetilde{\Xi }_n$| (5.1.1) as follows:
(6.3.1)
For any matrix |$A \in \widetilde{\Xi }_n$| and |$p\in 2{\mathbb{N}}$|⁠, we define
(6.3.2)

 

Lemma 6.3.1.
For |$A_1, A_2,\ldots , A_f \in \widetilde{\Xi }_n^>$|⁠, there exist |$\mathcal{Z}_i \in \widetilde{\Xi }_n^>$| and |$\zeta ^{\imath }_i(u,v,\pi ) \in \mathcal{R}_1$| such that for all even integers |$p\gg 0$|⁠, we have an identity in |${\mathbb{S}}^{\jmath }_{n,d}$| of the form

 
Proof.

The proof is similar to the proof of Proposition 5.1 where |${_{p}}{A} = A+pI$| is used instead of |${_{\breve{p}}}{A}$|⁠.

Consequently, the vector space |$\dot{{\mathbb{K}}}^{>}_n $| over |${\mathbb{A}}$| spanned by the symbols |$[A]$|⁠, for |$A \in \widetilde{\Xi }_n^>$|⁠, is a stabilization algebra whose multiplicative structure is given by (with |$f=2$|⁠; associativity follows from |$f=3$|⁠):
(6.3.3)

Precisely, we have the following multiplication formulas for Chevalley generators in |$\dot{{\mathbb{K}}}^{>}_n $|⁠.

 

Proposition 6.3.2

Let |$A, B, C \in \widetilde{\Xi }_n^>$|⁠, and |$h\in [1,n]$|⁠.

  • (1) If |$B-bE_{h,h-1}^{\theta }$| is diagonal and |$\textrm{col}(B)=\textrm{row}(A)$|⁠, then
    (6.3.4)
    where |$t=(t_i)_{-n\leqslant i\leqslant n}\in{\mathbb{N}}^N$| with |$\sum _{i=-n}^n t_i=b$| such that
  • (2) Suppose |$C-ce:{h-1,h}^{\theta }$| is diagonal and |$\textrm{col}(C)=\textrm{row}(A)$|⁠.If |$h\neq 1$| then
    (6.3.5)
    where |$t=(t_i)_{-n\leqslant i\leqslant n}\in{\mathbb{N}}^N$| with |$\sum _{i=-n}^n t_i=c$| such that |$t_i\leqslant a_{h,i}$| if |$i\neq h$|⁠.If |$h=1$| then
    (6.3.6)
    where |$t=(t_i)_{-n\leqslant i\leqslant n}\in{\mathbb{N}}^N$| with |$\sum _{i=-n}^n t_i=c$| such that |$t_i\leqslant a_{1,i}$| if |$i\neq 1$|⁠.

By arguments entirely analogous to those for Corollary 5.3 and Theorem 5.6, |$\dot{{\mathbb{K}}}^{>}_n $| admits a (stabilizing) bar involution, |$\dot{{\mathbb{K}}}^{>}_n $| admits a monomial basis |$\{m_A | A \in \widetilde{\Xi }_n^>\},$| and a canonical basis |$\dot{\mathfrak B}^{\jmath ,>}$|⁠. Let |$\dot{{\mathbb{K}}}^{\imath }_{n}$| be the |${\mathbb{A}}$|-submodule of |$\dot{{\mathbb{K}}}^{>}_n $| generated by |$\{ [A] | A \in \widetilde{\Xi ^{\imath }}\}$|⁠, where
(6.3.7)
The goal of this subsection is to realize |$\dot{{\mathbb{K}}}^{\imath }_{n}$| as a subquotient of |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| with compatible bases by following [3, Appendix A]. It follows from (6.3.7) that |$\dot{{\mathbb{K}}}^{\imath }_{n}$| is a subalgebra of |$\dot{{\mathbb{K}}}^{>}_n $|⁠. Since the bar-involution on |$\dot{{\mathbb{K}}}^{>}_n $| restricts to an involution on |$\dot{{\mathbb{K}}}^{\imath }_{n}$|⁠, we reach the following conclusion.

 

Lemma 6.3.3

The set |$\dot{{\mathbb{K}}}^{\imath }_{n} \cap \dot{\mathfrak B}^{\jmath ,>}$| forms a canonical basis of |$\dot{{\mathbb{K}}}^{\imath }_{n}$|⁠.

The submodule of |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| spanned by |$[A]$| for |$A \in \widetilde{\Xi ^{\imath }}$| is not a subalgebra. This is why we need a somewhat different stabilization above to construct the canonical basis for |$\dot{{\mathbb{K}}}^{\imath }_{n}$|⁠. We shall see below the stabilization above is related to the stabilization used earlier. Define |${\mathbb{J}}$| to be the |${\mathbb{A}}$|-submodule of |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| spanned by |$[A]$| for all |$A \in \widetilde{\Xi }_n^<$|⁠.

 

Lemma 6.3.4

The submodule |${\mathbb{J}}$| is a two-sided ideal of |$\dot{{\mathbb{K}}}^{\jmath }_{n}$|⁠.

 

Proof.
We note that |${\mathbb{J}}$| is clearly invariant under the anti-involution for |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| below:
(6.3.8)
Hence, the claim that |${\mathbb{J}}$| is a left ideal of |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| is equivalent to that |${\mathbb{J}}$| is a right ideal of |$\dot{{\mathbb{K}}}^{\jmath }_{n}$|⁠. We shall show that |${\mathbb{J}}$| is a left ideal of |$\dot{{\mathbb{K}}}^{\jmath }_{n}$|⁠. To that end, it suffices to show that |$[B] [A] \in{\mathbb{J}}$| for arbitrary |$A \in \widetilde{\Xi }_n^<$| and |$B\in \widetilde{\Xi }_n$| such that |$B-bE_{h,h-1}$| or |$B-bE_{h-1,h}$| is diagonal for some |$h\in [1,n]$| and |$b\geqslant 0$|⁠. Thanks to the multiplication formulas in Proposition 5.1.4, unless the case of |$B-bE_{0,1}^{\theta }$| being diagonal, the |$(0,0)$|-entry of the terms arising in |$[B][A]$| never exceeds |$a_{0,0}$|⁠. Thus, |$[B][A]\in{\mathbb{J}}$| in these cases.
Consider the case that |$B-bE_{0,1}^{\theta }$| is diagonal. Recall the formula (6.3.6). If the |$(0,0)$|-entry |$a_{0,0}+2t_0$| of the term |$[\widehat{A}_{t,1}]$| is positive, then the coefficient of this term must be zero since
because of |$a^{\natural }_{00}+1\leqslant 0< a^{\natural }_{00}+t_{0}$|⁠. Therefore, we always have |$[B][A] \in{\mathbb{J}}$|⁠.

 

Lemma 6.3.5

If |$A \in \widetilde{\Xi }_n^<$| then |$m_A \in{\mathbb{J}}$|⁠.

 

Proof.

The proof is as the same as the one of [3, Lemma A.6 (1)].

Recall |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| admits a canonical basis of |$\dot{{\mathfrak{B}}}$| at the specialization |$u=\textbf{v}^{\textbf{L} (s_0)}, v=\textbf{v}^{\textbf{L} (s_1)}$| from Theorem 5.2.2.

 

Theorem 6.3.6

The ideal |${\mathbb{J}}$| admits a monomial basis |$\{m_A | A \in \widetilde{\Xi }_n^<\}$|⁠. Moreover, its specialization at |$u=\textbf{v}^{\textbf{L} (s_0)}, v=\textbf{v}^{\textbf{L} (s_1)}$| (denoted by |${\mathbb{J}}^{\textbf{L}} $|⁠) has a canonical basis |$\dot{{\mathfrak{B}}} \cap{\mathbb{J}}^{\textbf{L}} = \{\{A\}^{\textbf{L}} | A \in \widetilde{\Xi }_n^<\}$|⁠.

 

Proof.

The 1st statement follows from the above lemma directly. Since |$m_A=[A]+ \ \textrm{lower terms}$|⁠, we know that |${\mathbb{J}}^{\textbf{L}} $| is bar invariant. Thus, |${\mathbb{J}}^{\textbf{L}} $| does admit a canonical bases parameterized by |$A \in \widetilde{\Xi }_n^<$|⁠, which should be |$\dot{{\mathfrak{B}}} \cap{\mathbb{J}}^{\textbf{L}} = \{\{A\}^{\textbf{L}} | A \in \widetilde{\Xi }_n^<\}$| by the uniqueness of canonical basis.

 

Proposition 6.3.7

The following statements hold:

  • (a) The quotient algebra |$\dot{{\mathbb{K}}}^{\jmath }_{n}/{\mathbb{J}}$| admits a monomial basis |$\{m_A + {\mathbb{J}} | A\in \widetilde{\Xi }_n^>\}$|⁠.

  • (b) The specialization at |$u = \textbf{v}^{\textbf{L} (s_0)}, v= \textbf{v}^{\textbf{L} (s_1)}$| of the quotient algebra |$\dot{{\mathbb{K}}}^{\jmath }_{n}/{\mathbb{J}}$| admits a canonical basis |$\{\{A\}^{\textbf{L}} + {\mathbb{J}}^{\textbf{L}} | A \in \widetilde{\Xi }_n^>\}$|⁠.

  • (c) The map |$\sharp : \dot{{\mathbb{K}}}^{\jmath }_{n}/{\mathbb{J}} \rightarrow \dot{{\mathbb{K}}}^{>}_n $| sending |$[A] + {\mathbb{J}} \mapsto [A]$| is an isomorphism of |${\mathbb{A}}$|-algebras, which matches the corresponding monomial bases. It also matches the corresponding canonical bases at the specialization |$u = \textbf{v}^{\textbf{L} (s_0)}, v= \textbf{v}^{\textbf{L} (s_1)}$|⁠.

 

Proof.

Parts (a) and (b) follow directly from Theorem 6.3.6. Below we prove the Part (c). Knowing that the map |$\sharp $| is a linear isomorphism, we need to verify it is an algebraic homomorphism. Comparing the multiplication formulas for |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| in Proposition 5.1.4 with the ones for |$\dot{{\mathbb{K}}}^{>}_n $| in Proposition 6.3.2, we can see that the structure constants with respect to the Chevalley generators for |$\dot{{\mathbb{K}}}^{\jmath }_{n}/{\mathbb{J}}$| are as the same as those for |$\dot{{\mathbb{K}}}^{>}_n $|⁠. Therefore, |$\sharp $| is an algebraic homomorphism.

Since |$\sharp $| matches the Chevalley generators, it matches the corresponding monomial bases. We also obtain that |$\sharp $| commutes with the bar involution. Notice that the partial orders |$<_{\textrm{alg}}$| are compatible; hence, |$\sharp $| also matches the corresponding canonical bases at the specialization |$u = \textbf{v}^{\textbf{L} (s_0)}, v= \textbf{v}^{\textbf{L} (s_1)}$|⁠.

We summarize Lemma 6.3.3 and Proposition 6.3.7 above as follows.

 

Theorem 6.3.8

As an |${\mathbb{A}}$|-algebra, |$\dot{{\mathbb{K}}}^{\imath }_{n}$| is isomorphic to a subquotient of |$\dot{{\mathbb{K}}}^{\jmath }_{n}$|⁠, with compatible standard, monomial basis. They have compatible canonical bases at the specialization |$u = \textbf{v}^{\textbf{L} (s_0)}, v= \textbf{v}^{\textbf{L} (s_1)}$|⁠.

Let |$\dot{{\mathbb{K}}}_n^{\jmath ,1}$| be the |${\mathbb{A}}$|-submodule of |$\dot{{\mathbb{K}}}^{\jmath }_{n}$| spanned by |$[A]$| where |$A\in \widetilde{\Xi }_{n}$| with |$\textrm{row}(A)_{0}=\textrm{col}(A)_0=1$|⁠. It is clear that |$\dot{{\mathbb{K}}}_n^{\jmath ,1}$| is a subalgebra of |$\dot{{\mathbb{K}}}^{\jmath }_{n}$|⁠. Let |${\mathbb{J}}^1={\mathbb{J}}\cap \dot{{\mathbb{K}}}_n^{\jmath ,1}$|⁠, that is,
Imitating the argument in [3, §A.3], we have the following.

 

Proposition 6.3.9

  • (a) The monomial basis of |${\dot{{\mathbb{K}}}^{\jmath }_{n}}$| restricts to the monomial basis of |$\dot{{\mathbb{K}}}_n^{\jmath ,1}$|⁠; the monomial basis of |$\dot{{\mathbb{K}}}_n^{\jmath ,1}$| restricts to the monomial basis of |${\mathbb{J}}^1$|⁠. So does the canonical basis at the specialization |$u = \textbf{v}^{\textbf{L} (s_0)}, v= \textbf{v}^{\textbf{L} (s_1)}$|⁠.

  • (b) The quotient |${\mathbb{A}}$|-subalgebra |$\dot{{\mathbb{K}}}_n^{\jmath ,1}/{\mathbb{J}}^1$| admits a monomial basis |$\{m_A + {\mathbb{J}}^1 | A\in \widetilde{\Xi ^{\imath }}\}$|⁠. It also admits a canonical basis |$\{\{A\}^{\textbf{L}} + {\mathbb{J}}^{1,\textbf{L}} \ | A \in \widetilde{\Xi ^{\imath }}\}$| at the specialization |$u = \textbf{v}^{\textbf{L} (s_0)}, v= \textbf{v}^{\textbf{L} (s_1)}$|⁠, where |${\mathbb{J}}^{1,\textbf{L}} \ ={\mathbb{J}}^1|_{u = \textbf{v}^{\textbf{L} (s_0)}, v= \textbf{v}^{\textbf{L} (s_1)}}$|⁠.

  • (c) There is an |${\mathbb{A}}$|-algebra isomorphism |$\dot{{\mathbb{K}}}_n^{\jmath ,1}/{\mathbb{J}}^1\cong \dot{{\mathbb{K}}}^{\imath }_{n}$|⁠, which matches the corresponding monomial bases. It also matches the corresponding canonical basis at the specialization |$u = \textbf{v}^{\textbf{L} (s_0)}, v= \textbf{v}^{\textbf{L} (s_1)}$|⁠.

7 Quantum Symmetric Pairs

7.1 The quantum symmetric pair |$({\mathbb{U}},{\mathbb{U}}^{\jmath })$|

We start with the quantum symmetric pairs of type AIII/AIV without fixed points nor black nodes, associated with the following Satake diagram:

Note that we use half integers for the index set following the convention in [5]. Set
(7.1.1)
Let |$\mathbb{U}=\mathbb{U}(\mathfrak{g}\mathfrak{l}_{2n+1})$| be the algebra over |${\mathbb{Q}}(u,v)$| generated by |$E_i, F_i$|⁠, |$(i\in{\mathbb{I}}_{2n})$| and |$D_a$|⁠, |$(a\in [-n,n])$| subject to the following relations, for |$i,j\in{\mathbb{I}}_{2n}, a,b\in [-n,n]$|⁠:
(7.1.2)
 
(7.1.3)
 
(7.1.4)
 
(7.1.5)
 
(7.1.6)
(Here and below |$K_i:=D_{i-\frac{1}{2}}D_{i+\frac{1}{2}}^{-1}$|⁠.)
Let |${\mathbb{U}}^{\jmath } = {\mathbb{U}}^{\jmath }(\mathfrak{g}\mathfrak{l}_{2n+1})$| be the |${\mathbb{Q}}(u,v)$|-algebra with generators ZZ
subject to the following relations, for |$i\in{\mathbb{I}}^{\jmath } _n,a,b\in [0,n]$|⁠:
(7.1.7)
 
(7.1.8)
 
(7.1.9)
 
(7.1.10)
 
(7.1.11)
 
(7.1.12)
 
(7.1.13)
 
(7.1.14)
(Here |$k_i=d_{i-\frac{1}{2}}d_{i+\frac{1}{2}}^{-1}$|⁠, |$(i\not =\frac{1}{2})$|⁠, and |$k_{\frac{1}{2}}=v^{-1}d_0d_1^{-1}$|⁠.)
It is known in [7, §4.1] that there is a |${\mathbb{Q}}(u,v)$|-algebra homomorphism |${\mathbb{U}}^{\jmath } \to{\mathbb{U}}$| given by, for |$i \in{\mathbb{I}}^{\jmath } _n-\{\frac{1}{2}\}$|⁠, and for |$1\leqslant a \leqslant n$|⁠,
(7.1.15)

 

Remark 7.1.1.

The (multiparameter) quantum symmetric pairs |$({\mathbb{U}},{\mathbb{U}}^{\jmath })$| in this paper are the |$\mathfrak{g}\mathfrak{l}$|-variant of the quantum symmetric pairs in [7].

7.2 Isomorphism |$\dot{\mathbb{U}}^{\jmath }\simeq \dot{{\mathbb{K}}}^{\jmath }_{n}$|

Following [22, §23.1], it is routine to define the modified quantum algebra |$\dot{\mathbb{U}}^{\jmath }$| from |$\mathbb{U}^{\jmath }$|⁠. Let |$\widetilde{\Xi }_n^{\textrm{diag}}$| be the set of all diagonal matrices in |$\widetilde{\Xi }_n$|⁠. Denote by |$\lambda =\textrm{diag}(\lambda _{-n},\lambda _{-n+1},\ldots , \lambda _n)$| a diagonal matrix in |$\widetilde{\Xi }_n^{\textrm{diag}}$|⁠. For |$\lambda ,\lambda ^{\prime }\in \widetilde{\Xi }_n^{\textrm{diag}}$|⁠, we set
(7.2.1)
The modified quantum algebra |$\dot{\mathbb{U}}^{\jmath }$| is defined by
(7.2.2)
Let |$1_\lambda =p_{\lambda ,\lambda }(1)$|⁠, where |$p_{\lambda ,\lambda }: \mathbb{U}^{\jmath }\rightarrow{}_\lambda{\mathbb{U}}^{\jmath }_{\lambda }$| is the canonical projection. Thus, the unit of |$\mathbb{U}^{\jmath }$| is replaced by a collection of orthogonal idempotents |$1_\lambda $| in |$\dot{\mathbb{U}}^{\jmath }$|⁠. It is clear that
For |$\lambda \in \widetilde{\Xi }_n^{\textrm{diag}}$| and |$i\in \mathbb{I}_n^{\jmath }$|⁠, we use the following short-hand notations:
(7.2.3)
We also define, for |$r \in{\mathbb{N}}$|⁠,
(7.2.4)
A multiparameter version of [3, Proposition 4.6] gives a presentation of |$\dot{{\mathbb{U}}}^{\jmath } $| as a |${\mathbb{Q}}(u,v)$|-algebra generated by the symbols, for |$i \in{\mathbb{I}}^{\jmath } _n, \lambda \in \widetilde{\Xi }_n^{\textrm{diag}},$|  
subject to the following relations, for |$i,j \in{\mathbb{I}}^{\jmath } _n, \lambda ,\mu \in \widetilde{\Xi }_n^{\textrm{diag}}$|⁠, |$x,y \in \{1, e_i, e_j, f_i, f_j\}$|⁠:
(7.2.5)
 
(7.2.6)
 
(7.2.7)
 
(7.2.8)
 
(7.2.9)
 
(7.2.10)
 
(7.2.11)
 
(7.2.12)
Here and below we always write |$x_1 1_{\lambda ^1} x_2 1_{\lambda ^2}\cdots x_k 1_{\lambda ^k}=x_1x_2\cdots x_k 1_{\lambda ^k}$|⁠, if the product is not zero; in this case such |$\lambda ^1,\lambda ^2,\ldots ,\lambda ^{k-1}$| are all uniquely determined by |$\lambda ^k$|⁠.
|$\forall i\in \mathbb{I}^{\jmath }_n, \lambda \in \widetilde{\Xi }_n^{\textrm{diag}}$|⁠, write

Set |${}_{\mathbb{Q}}\dot{{\mathbb{K}}}^{\jmath }_{n}={\mathbb{Q}}(u,v)\otimes _{\mathbb{A}} \dot{{\mathbb{K}}}^{\jmath }_{n}$|⁠.

 

Theorem 7.2.1
There is an isomorphism of |${\mathbb{Q}}(u,v)$|-algebras |$\aleph :\dot{\mathbb{U}}^{\jmath }\rightarrow{}_{{\mathbb{Q}}}\dot{{\mathbb{K}}}^{\jmath }_{n}$| such that, |$\forall i\in \mathbb{I}^{\jmath }_n, \lambda \in \widetilde{\Xi }_n^{\textrm{diag}}$|⁠,

 

Proof.
A direct computation using Theorem 4.2.3 shows that relations (7.2.5)-(7.2.12) also hold if we replace |$e_i, f_i$|’s by |$\textbf{e}_i, \textbf f_i$|’s. Here we only present details for (7.2.11) regarding |$\textbf{e}_{\frac{1}{2}} 1_\lambda $| and |$\textbf f_{\frac{1}{2}} 1_\lambda $| as follows:
Combining the identities above, we get |$( [\![ 2 ]\!] \textbf{e}_{\frac{1}{2}}\textbf f_{\frac{1}{2}}\textbf{e}_{\frac{1}{2}}-\textbf{e}_{\frac{1}{2}}^2\textbf f_{\frac{1}{2}}-\textbf{f}_{\frac{1}{2}}\textbf{e}_{\frac{1}{2}}^{2})1_\lambda = [\![ 2 ]\!] (uv^{\lambda _0-\lambda _1}+u^{-1}v^{-\lambda _0+\lambda _1})\textbf{e}_{\frac{1}{2}} 1_\lambda $|⁠. That is, |$\aleph $| is indeed an algebra homomorphism.

We also know that |$\aleph $| is a linear isomorphism. The argument is almost as the same as that for the case of specialization at |$u=v$|⁠, which can be found in the proof of [3, Theorem 4.7]. Therefore, |$\aleph $| is an isomorphism of |${\mathbb{Q}}(u,v)$|-algebras.

7.3 The quantum symmetric pair |$({\mathbb{U}},{\mathbb{U}}^{\imath })$|

Below we formulate the counterparts of Sections 7.1 and 7.2. The proofs are very similar and will often be omitted. We now work on quantum symmetric pairs of type AIII with fixed points associated with the Satake diagram below:

Let |$\mathbb{U}=\mathbb{U}(\mathfrak{g}\mathfrak{l}_{2n})$| be the algebra over |${\mathbb{Q}}(u,v)$| generated by |$E_i, F_i$|⁠, |$(i\in [-n+1,n-1])$| and |$D_a$|⁠, |$(a\in [-n+1,n])$| subject to the following relations, for |$i,j\in [-n+1,n-1], a,b\in [-n+1,n]$|⁠:
(7.3.1)
 
(7.3.2)
 
(7.3.3)
 
(7.3.4)
 
(7.3.5)
(Here and below |$K_i:=D_{i}D_{i+1}^{-1}$|⁠.)
Let |${\mathbb{U}}^{\imath } = {\mathbb{U}}^{\imath }(\mathfrak{g}\mathfrak{l}_{2n})$| be the |${\mathbb{Q}}(u,v)$|-algebra with generators
subject to the following relations, for |$i,j\in [1, n-1], a,b\in [1,n]$|⁠:
(7.3.6)
 
(7.3.7)
 
(7.3.8)
 
(7.3.9)
 
(7.3.10)
 
(7.3.11)
 
(7.3.12)
 
(7.3.13)
(Here |$k_i=d_{i}d_{i+1}^{-1}$|⁠.)
It has been known in [7, §2.1] that there is a |${\mathbb{Q}}(u,v)$|-algebra homomorphism |${\mathbb{U}}^{\imath } \to{\mathbb{U}}$| given by, for |$i \in [1,n-1]$|⁠, and for |$a\in [1,n]$|⁠,
(7.3.14)

 

Remark 7.3.1

It was observed in [7, 20] that the parameter |$\omega \in{\mathbb{Q}}(u,v)$| in the embedding |$t=E_0+vF_0K_0^{-1}+\omega K_0^{-1}$| is irrelevant to the presentation of the algebra |${\mathbb{U}}^{\imath }$|⁠.

Let |${}^{\imath }\widetilde{\Xi }_n^{\textrm{diag}}$| be the set of all diagonal matrices in |$\widetilde{\Xi }^{\imath }$|⁠. Denote by |$\lambda =\textrm{diag}(\lambda _{-n},\ldots ,\lambda _{-1},1,\lambda _{1}, \ldots , \lambda _n)$| a diagonal matrix in |${}^{\imath }\widetilde{\Xi }_n^{\textrm{diag}}$|⁠. We define the modified algebra |$\dot{{\mathbb{U}}}^{\imath } $| similarly to the construction of |$\dot{{\mathbb{U}}}^{\jmath } $| as follows:
where |${}_\lambda \mathbb{U}^{\imath }_{\lambda ^{\prime }}= \mathbb{U}^{\imath }/\left (\sum _{a=1}^n(d_a-v^{\lambda _a})\mathbb{U}^{\imath } +\sum _{a=1}^n\mathbb{U}^{\imath }(d_a-v^{\lambda ^{\prime }_a})\right )$| and |$1_\lambda \in{}_\lambda \mathbb{U}^{\imath }_{\lambda }$| is the canonical projection image of the unit of |$\mathbb{U}^{\imath }$|⁠.
For |$\lambda \in{}^{\imath }\widetilde{\Xi }_n^{\textrm{diag}}$| and |$i\in [1,n-1]$|⁠, we use the following short-hand notations:
(7.3.15)
We thus obtain a presentation of |$\dot{{\mathbb{U}}}^{\imath } $| as a |${\mathbb{Q}}(u,v)$|-algebra generated by the symbols, for |$i \in [1,n-1], \lambda \in{}^{\imath }\widetilde{\Xi }_n^{\textrm{diag}},$|  
subject to the following relations, for |$i,j \in [1,n-1], \lambda ,\mu \in{}^{\imath }\widetilde{\Xi }_n^{\textrm{diag}}$|⁠, |$x,y \in \{1, e_i, e_j, f_i, f_j, t\}$|⁠:
(7.3.16)
 
(7.3.17)
 
(7.3.18)
 
(7.3.19)
 
(7.3.20)
 
(7.3.21)
 
(7.3.22)
 
(7.3.23)
 
(7.3.24)
For |$i\in [1,n-1], \lambda \in{}^{\imath }\widetilde{\Xi }_n^{\textrm{diag}}$|⁠, write
(7.3.25)
Set |${}_{\mathbb{Q}}\dot{{\mathbb{K}}}^{\imath }_{n}={\mathbb{Q}}(u,v)\otimes _{\mathbb{A}} \dot{{\mathbb{K}}}^{\imath }_{n}$|⁠.

 

Theorem 7.3.2
There is an isomorphism of |${\mathbb{Q}}(u,v)$|-algebras |$\aleph :\dot{\mathbb{U}}^{\imath }\rightarrow{}_{\mathbb{Q}}\dot{{\mathbb{K}}}^{\imath }_{n}$| such that, for all |$i\in [1,n-1], \lambda \in \widetilde{\Xi }_n^{\textrm{diag}}$|⁠,

 

Proof.
By a direct computation using Theorem 4.2.3 one can show that the relations (7.3.16)–(7.3.24) for |$t, e_i, f_i$|’s also hold for |$\textbf{t}, \textbf{e}_i, \textbf{f}_i$|’s. Hence, |$\aleph $| is a homomorphism of |${\mathbb{Q}}(u,v)$|-algebras. Here we only present the details for the 1st relation in (7.3.24) as follows. Note that as an element in |$\dot{{\mathbb{K}}}^{>}_n $|⁠,
(7.3.26)
where |$\textbf{e}_0 1_\lambda =[\lambda -E_{1,1}^{\theta }+E_{0,1}^{\theta }]\quad \ \textrm{and}\quad \textbf f_0 1_{\lambda +E_{0,0}^{\theta }-E_{1,1}^{\theta }}=[\lambda -E_{1,1}^{\theta }+E_{1,0}^{\theta }]\in \dot{{\mathbb{K}}}^{>}_n.$| Moreover, we have
Hence,
and
Finally, using (7.3.26) again, we compute that
Combining the identities above, we see that indeed |$(\textbf{t}^2 \textbf{e}_1 + \textbf{e}_1 \textbf{t}^2)1_\lambda = (\left [\![ 2 \right ]\!] \textbf{t} \textbf{e}_1 \textbf{t} + \textbf{e}_1)1_\lambda .$|

An argument similar to the proof of [3, Theorem A.15] also shows |$\aleph $| is a linear isomorphism. Therefore, |$\aleph $| is an isomorphism of |${\mathbb{Q}}(u,v)$|-algebras.

 

Remark 7.3.3
It has been shown in [7, Lemmas 2.1 and 4.1] that there exists a unique |${\mathbb{Q}}$|-linear bar involution on |$\mathbb{U}^{\jmath }$| and |$\mathbb{U}^{\imath }$| such that
This bar involution induces a compatible bar involution on |$\dot{\mathbb{U}}^{\jmath }$| and |$\dot{\mathbb{U}}^{\imath }$|⁠, fixing all the generators |$1_\lambda $|⁠, |$e_i 1_\lambda $|⁠, |$f_i 1_\lambda , t1_\lambda $|⁠. Note that |$\textbf{e}_i 1_\lambda $|⁠, |$\textbf f_i 1_\lambda $|⁠, |$\textbf{t} 1_\lambda $|⁠, |$[\lambda ]$| are bar invariant elements, which implies that the isomorphism |$\aleph $| intertwines the bar involution on |$\dot{\mathbb{U}}^{\jmath }$| (resp. |$\dot{\mathbb{U}}^{\imath }$|⁠) and on |${}_{\mathbb{Q}}\dot{\mathbb{K}}_n^{\jmath }$| (resp. |${}_{\mathbb{Q}}\dot{\mathbb{K}}_n^{\imath }$|⁠) (cf. [3] for the equal parameter case). So Theorem 5.6 (resp. Theorem 6.9) provides a canonical basis for |$\dot{\mathbb{U}}^{\jmath }$| (resp. |$\dot{\mathbb{U}}^{\imath }$|⁠) at the specialization |$u = \textbf{v}^{\textbf{L} (s_0)}, v= \textbf{v}^{\textbf{L} (s_1)}$|⁠. A general theory of canonical bases for quantum symmetric pairs with parameters of arbitrary finite type was developed in [6].

Appendix A. An Algebraic Approach to Schur Algebras of Type D

As we mentioned in Section 2, at the specialization |$u=1$| the multiparameter Schur duality yields a weak Schur duality of type D that is used in [1] to formulate the Kazhdan–Lusztig theory for classical and super type D. These algebras |$\mathbb{S}^{\bullet }_{n,d}|_{u=1}$| (⁠|$\bullet = \imath $| or |$\jmath $|⁠), however, are not the Schur algebras introduced in [14]. While bases of Schur algebras of finite type A/B/C and affine type A/C can be parametrized by a matrix set (cf. |$\Xi _{n,d}$| in 2.2.2), for finite type D Fan and Li showed that a matrix set is not enough—a notion of signed matrices that indexes a larger algebra is needed. From a geometric point of view, this reflects the fact that there are two connected components for the maximal isotropic Grassmannian associated with |$\textrm{SO}(2d)$|⁠. In this appendix, we provide an algebraic approach to Fan–Li’s construction parallel to our multiparameter results. With our algebraic approach we also clear up some previous misconceptions (e.g., compare (A.4.2) and [14, (22)]). The arguments are very similar to the multiparameter counterpart, so we will omit the easy proofs in this appendix.

A.1 Weyl groups of type |$\textbf{D}$|

Fix |$d\in \mathbb{N}$|⁠, and we set
(A.1.1)
Let |$W_{\textbf{D}}$| be the Weyl group of type |$\textbf{D}_d$|⁠. It is known (c.f. [2]) that |$W_{\textbf{D}}$| can be identified as a permutation subgroup of |$J_d$| that consists of those permutations |$g$| satisfying that
Let |$S_{\textbf{D}}=\{\varsigma _0,\varsigma _1,\ldots ,\varsigma _{d-1}\}$|⁠, where |$\varsigma \in W_{\textbf{D}}$| are given by the following products of transpositions:
It is also known (see [2, (8.18) and (8.19)]) that |$(W_{\textbf{D}}, S_{\textbf{D}})$| is a Coxeter group associated with the length function as below:

 

Lemma A.1.1
The length of |$g\in W_{\textbf{D}}$| is given by

A.2 Signed compositions

Fix |$n\in{\mathbb{N}}$|⁠. Recall that (2.1.8) first |$\Lambda _{n,d}$| is the set of weak compositions of |$d$| into |$n+1$| parts. Set
(A.2.1)
In below we abbreviate |$(\lambda ,\alpha )\in \Lambda ^{\alpha }$| by |$\lambda ^{\alpha }$| where |$\alpha \in \{0,+,-\}$|⁠. We further set
(A.2.2)
Elements in |$\Lambda _{\textbf{D}}$| will be called signed compositions. Recall that |$\lambda _{0,i} =\lambda _0 + \lambda _1 + \cdots + \lambda _i$| for |$i\in [0,n], \lambda \in \Lambda _{n,d}$|⁠. We define positive integer intervals associated with |$\lambda ^{\alpha }$| by
(A.2.3)
 
(A.2.4)
For |$-n\leqslant i \leqslant 1$|⁠, we set |$R_{i}^{\lambda ^{\alpha }}=\{-x|x\in R_{-i}^{\lambda ^{\alpha }}\}$|⁠. We remark that the sets |$\{R_i^{\lambda ^{\alpha }}\}_{i\in [-n,n]}$| partition the set |$J_d$|⁠.
Denote by |$ \ \textrm{Stab}(X)$| the stabilizer of |$J_d$| in |$W_{\textbf{D}}$|⁠, for any |$X\subset J_d$|⁠. For any |$\lambda ^{\alpha }\in \Lambda _{\textbf{D}}$|⁠, let
(A.2.5)

 

Lemma A.2.1.
For any |$\lambda ^{\alpha }\in \Lambda _{\textbf{D}}$|⁠, |$W_{\lambda ^{\alpha }}$| is the parabolic subgroup of |$W_{\textbf{D}}$| generated by

Denote the set of minimal length right coset representatives of |$W_{\lambda ^{\alpha }}$| in |$W_{\textbf{D}}$| by
(A.2.6)
Hence, the set |$\mathscr{D}_{\lambda ^{\alpha }\mu ^{\beta }} = \mathscr{D}_{\lambda ^{\alpha }} \cap \mathscr{D}_{\mu ^{\beta }}^{-1}$| is the set of minimal length double coset representatives for |$W_{\lambda ^{\alpha }} \backslash W_{\textbf{D}} /W_{\mu ^{\beta }}$|⁠.

 

Lemma A.2.2.

Let |$g \in W_{\textbf{D}}$| and |$\lambda ^{\alpha } \in \Lambda _{\textbf{D}}$|⁠.

  • (a) If |$\alpha =\pm $|⁠, then |$g\in \mathscr{D}_{\lambda ^{\alpha }}$| if and only if |$g^{-1}$| is order-preserving on |$R^{\lambda ^{\alpha }}_i$|⁠, for all |$i \in [1,n]$|⁠;

  • (b) If |$\alpha =0$|⁠, then |$g\in \mathscr{D}_{\lambda ^{\alpha }}$| if and only if |$g^{-1}$| is order-preserving on |$R^{\lambda ^{\alpha }}_i$| for all |$i \in [1,n]$| and

By a similar argument for [10, Proposition 4.16, Lemma 4.17, and Theorem 4.18], we have the following facts.

 

Proposition A.2.3.

Let |$\lambda ^{\alpha },\mu ^{\beta } \in \Lambda _{\textbf{D}}$|⁠, and |$g \in \mathscr{D}_{\lambda ^{\alpha }\mu ^{\beta }}$|⁠.

  • (a) There is a weak composition |$\delta = \delta (\lambda ^{\alpha }, g, \mu ^{\beta }) \in \Lambda _{n^{\prime},d}$| for some |$n^{\prime}$| such that |$W_{\delta ^{\beta }} = g^{-1} W_{\lambda ^{\alpha }} g \cap W_{\mu ^{\beta }}$|⁠.

  • (b) The map |$W_{\lambda ^{\alpha }} \times (\mathscr{D}_\delta \cap W_{\mu ^{\beta }}) \rightarrow W_{\lambda ^{\alpha }} g W_{\mu ^{\beta }}$| sending |$(x,y)$| to |$xgy$| is a bijection; moreover, we have |$\ell (xgy) = \ell (x) + \ell (g) + \ell (y)$|⁠.

  • (c) The map |$(\mathscr{D}_\delta \cap W_{\mu ^{\beta }}) \times W_\delta \rightarrow W_{\mu ^{\beta }}$| sending |$(x,y)$| to |$xy$| is a bijection; moreover, we have |$\ell (x) + \ell (y) = \ell (xy)$|⁠.

A.3 Schur algebras

The Hecke algebra |$\textbf{H} = \textbf{H}(W_{\textbf{D}})$| over |$\textbf{A} ={\mathbb{Z}}[v,v^{-1}]$| is an |$\textbf{A}$|-algebra with basis |$\{T_g | g\in W_{\textbf{D}}\}$| satisfying that
For any finite subset |$X \subset W_{\textbf{D}}$| and for |$\lambda ^{\alpha }\in \Lambda _{\textbf{D}}$|⁠, set
(A.3.1)
For |$\lambda ^{\alpha },\mu ^{\beta }\in \Lambda _{\textbf{D}}$|⁠, and |$g\in \mathscr{D}_{\lambda ^{\alpha }\mu ^{\beta }}$|⁠, we consider a right |$\textbf{H}$|-linear map |$\phi _{\lambda ^{\alpha }\mu ^{\beta }}^g \in \textrm{Hom}_{\textbf{H}}(x_{\mu ^{\beta }} \textbf{H}, \textbf{H})$|⁠, sending |$x_{\mu ^{\beta }}$| to |$T_{W_{\lambda ^{\alpha }} g W_{\mu ^{\beta }}}.$| Thanks to Proposition A.2.3(b), we have |$T_{W_{\lambda^\alpha} g W_{\mu^\beta} } = x_{\lambda^\alpha} T_g T_{\mathscr{D}_\delta \cap W_{\mu^\beta} }$| for some |$\delta \in \Lambda _{n^{\prime },d}$|⁠, and hence we have constructed a right |$\textbf{H}$|-linear map
(A.3.2)
We define the Schur algebra |$\textbf{S} _{n,d}$| of type |$\textbf{D}$| as
(A.3.3)
Introduce the following subset of |$\Lambda _{\textbf{D}} \times W_{\textbf{D}} \times \Lambda _{\textbf{D}}$|⁠:
(A.3.4)

 

Lemma A.3.1.

The set |$\{\phi _{\lambda ^{\alpha }\mu ^{\beta }}^g | (\lambda ^{\alpha },g, \mu ^{\beta }) \in \mathscr{D}_{n,d} \}$| forms an |$\textbf{A}$|-basis of |$\textbf{S} _{n,d}$|⁠.

A.4 Signed matrices

From now on, we fix
Notice that |$D$| is even and is different from the convention (2.1.1). Set
(A.4.1)
We set
(A.4.2)
In below we abbreviate |$(A,\alpha )\in \Xi ^{\alpha }$| by |$A^{\alpha }$| where |$\alpha \in \{0,+,-\}$|⁠. We further set
(A.4.3)
whose elements are called signed matrices.
Define a sign map |$\textrm{sgn}:\{0,+,-\}^2 \times W_{\textbf{D}}\rightarrow \{0,+,-\}$| by
(A.4.4)
Define a map |$\kappa :\mathscr{D}_{n,d}\rightarrow \Xi _{\textbf{D}}$| by |$\kappa (\lambda ^{\alpha },g,\mu ^{\beta })=\left (|R_i^{\lambda ^{\alpha }}\cap gR_j^{\mu ^{\beta }}|\right )^{{\textrm{sgn}(\alpha ,\beta ,g)}}.$|

 

Lemma A.4.1.

The map |$\kappa :\mathscr{D}_{n,d}\rightarrow \Xi _{\textbf{D}}$| is a bijection.

For each |${\mathcal{A}}=\kappa (\lambda ^{\alpha },g,\mu ^{\beta })\in \Xi _{\textbf{D}}$|⁠, we write |$e_{\mathcal{A}}=\phi _{\lambda ^{\alpha }\mu ^{\beta }}^g$|⁠, and hence |$\{e_{\mathcal{A}}\mid A \in \Xi \}$| forms a basis of |$\textbf{S} _{n,d}$|⁠. For any |$A=(a_{ij})\in \Xi $|⁠, we set
(A.4.5)
Let |$I^+=(\{0\}\times [0,n])\sqcup ([1,n]\times [-n,n])$| be the index set corresponding to the “positive half part” of matrices in |$\Xi $|⁠.
 
Lemma A.4.2
If |$A^{{\textrm{sgn}(\alpha ,\beta ,g)}}=\kappa (\lambda ^{\alpha },g,\mu ^{\beta })\in \Xi _{\textbf{D}}$|⁠, where |$A=(a_{ij})\in \Xi $|⁠, then the length of |$g\in W_{\textbf{D}}$| is
(A.4.6)
In particular, the length is independent of the sign |${\textrm{sgn}(\alpha ,\beta ,g)}$|⁠. Thus, we write, for |${\mathcal{A}}=A^{{\textrm{sgn}(\alpha ,\beta ,g)}}=\kappa (\lambda ^{\alpha },g,\mu ^{\beta })\in \Xi _{\textbf{D}}$|⁠,
(A.4.7)
For each signed matrix |${\mathcal{A}}=A^{{\textrm{sgn}(\alpha ,\beta ,g)}}=\kappa (\lambda ^{\alpha },g,\mu ^{\beta })\in \Xi _{\textbf{D}}$| with |$A=(a_{ij})\in \Xi $|⁠, we introduce the following notations:
(A.4.8)
Note that |${\mathcal{A}}\pm B$| is a matrix instead of a signed matrix. The following lemmas follow immediately from definition.

 

Lemma A.4.3.

Let |${\mathcal{A}}=\kappa (\lambda ^{\alpha },g,\mu ^{\beta })\in \Xi _{\textbf{D}}$|⁠, then |$p({\mathcal{A}})=+$| (resp. −) if and only if |$g(1)>0$| (resp. |$<0$|⁠).

 

Lemma A.4.4.
For a signed matrix |${\mathcal{A}}\in \Xi _{\textbf{D}}$|⁠, we have
(A.4.9)

Let |${\mathcal{A}}=\kappa (\lambda ^{\alpha },g,\mu ^{\beta })\in \Xi _{\textbf{D}}$|⁠. We define a signed weak composition as below:
(A.4.10)

A direct computation shows that |$\delta ({\mathcal{A}})$| is indeed a weak composition |$\delta $| in Proposition A.2.3(a).

 
Proposition A.4.5.

Let |${\mathcal{A}}=\kappa (\lambda ^{\alpha },g,\mu ^{\beta })\in \Xi _{\textbf{D}}$|⁠. Then |$W_{\delta ({\mathcal{A}})} = g^{-1} W_{\lambda ^{\alpha }} g \cap W_{\mu ^{\beta }}.$|

We define type D quantum factorials by
We further define, for |$A=(a_{ij})\in \Xi $|⁠,
(A.4.11)
We write |$[{\mathcal{A}}]^!_{\mathfrak{d}}=[A]^!_{\mathfrak{d}}$| if |${\mathcal{A}}=A^{\textrm{sgn}{{\mathcal{A}}}}$|⁠. The type D quantum factorials are defined in the sense that the following identity on the Poincare polynomial for |$W_{\delta ({\mathcal{A}})}$| holds:

 

Lemma A.4.6.

For any |${\mathcal{A}}=A^{\alpha }\in \Xi _{\textbf{D}}$| with |$A=(a_{ij})$|⁠, we have |$\sum _{w\in W_{\delta ({\mathcal{A}})}}v^{2\ell (w)}=[A]_{\mathfrak{d}}^{!}.$|

A.5 Multiplication formulas

The proofs of Lemma A.5.1A.5.3 are very similar to their counterparts (Lemma 3.1.3, (3.2.2) and Lemma 3.2.1) so we omit.

 

Lemma A.5.1

Let |${\mathcal{A}}=\kappa (\lambda ^{\alpha },g,\mu ^{\beta })$| for |$\lambda ^{\alpha },\mu ^{\beta } \in \Lambda _{\textbf{D}}, g\in \mathscr{D}_{\lambda ^{\alpha }\mu ^{\beta }}$|⁠. Then |$x_{\lambda ^{\alpha }} T_{g} x_{\mu ^{\beta }} = [A]^!_{\mathfrak{d}} \, e_{\mathcal{A}}(x_{\mu ^{\beta }}).$|

 

Lemma A.5.2

Let |${\mathcal{B}} = \kappa (\lambda ^{\alpha },g_1,\mu ^{\beta })$| and |${\mathcal{A}} = \kappa (\mu ^{\beta },g_2,\nu ^{\gamma })$|⁠, where |$\lambda ^{\alpha },\mu ^{\beta }, \nu ^{\gamma } \in \Lambda _{\textbf{D}}$|⁠, |$g_1 \in \mathscr{D}_{\lambda ^{\alpha }\mu ^{\beta }}$|⁠, and |$g_2 \in \mathscr{D}_{\mu ^{\beta }\nu ^{\gamma }}$|⁠. Write |$\delta = \delta ({\mathcal{A}})$|⁠. Then we have |$e_{\mathcal{B}} e_{\mathcal{A}}(x_{\nu ^{\gamma }}) = \frac{1}{[A]^!_{\mathfrak{d}}} x_{\lambda ^{\alpha }} T_{g_1} T_{(\mathscr{D}_{\delta } \cap W_{\mu ^{\beta }})g_2} x_{\nu ^{\gamma }}.$|

 

Lemma A.5.3.
Let |${\mathcal{B}} = \kappa (\lambda ^{\alpha },1,\mu ^{\beta }), {\mathcal{A}} = \kappa (\mu ^{\beta },g,\nu ^{\gamma })$|⁠. Let |$y^{(w)}$| be the shortest double coset representative for |$W_\lambda wg W_\nu $|⁠, and let |${\mathcal{A}}^{(w)}=\kappa (\lambda ^{\alpha },y^{(w)},\nu ^{\gamma })$|⁠. Then

In the multiplication formulas below, we regard |$e_{\mathcal{A}}=0$| if |${\mathcal{A}}\not \in \Xi _{\textbf{D}}$|⁠.

 

Proposition A.5.3

Suppose that |${\mathcal{A}}=A^{\textrm{sgn}({\mathcal{A}})}, {\mathcal{B}}, {\mathcal{C}}\in \Xi _{\textbf{D}}$|⁠, and |$h\in [1,n]$|⁠. Let |$\Gamma _r=\{t=(t_i)_{-n\leqslant i\leqslant n}\in{\mathbb{N}}^N | \sum _{i=-n}^nt_i=r\}.$|  

  • (1) If |$h\neq 1$|⁠, |${\mathcal{B}}-rE_{h,h-1}^{\theta }$| is diagonal, |$\textrm{col}({\mathcal{B}})=\textrm{row}({\mathcal{A}})$|⁠, and |$s_r({\mathcal{B}})=s_l({\mathcal{A}})$|⁠, then
    (A.5.1)
    where |$ {{\mathcal{A}}}_{t,h}=(A+t_pE_{h,p}^{\theta }-t_pE_{h-1,p}^{\theta },\textrm{sgn}(s_l({\mathcal{B}}),s_r({\mathcal{A}})))$|⁠, |$s_l( {{\mathcal{A}}}_{t,h})=s_l({\mathcal{B}})$| and |$s_r( {{\mathcal{A}}}_{t,h})=s_r({\mathcal{A}})$|⁠.
  • (2) If |${\mathcal{B}}-rE_{1,0}^{\theta }$| is diagonal, |$\textrm{col}({\mathcal{B}})=\textrm{row}({\mathcal{A}})$|⁠, and |$s_r({\mathcal{B}})=s_l({\mathcal{A}})$|⁠, then
    (A.5.2)
  • (3) If |$h\neq 1$|⁠, |${\mathcal{C}}-rE_{h-1,h}^{\theta }$| is diagonal, |$\textrm{col}({\mathcal{C}})=\textrm{row}({\mathcal{A}})$|⁠, and |$s_r({\mathcal{C}})=s_l({\mathcal{A}})$|⁠, then
    (A.5.3)
    where |$\widehat{{\mathcal{A}}}_{t,h}=(A-t_pE_{h,p}^{\theta }+t_pE_{h-1,p}^{\theta },\textrm{sgn}(s_l({\mathcal{C}}),s_r({\mathcal{A}})))$|⁠, |$s_l(\widehat{{\mathcal{A}}}_{t,h})=s_l({\mathcal{C}})$| and |$s_r(\widehat{{\mathcal{A}}}_{t,h})=s_r({\mathcal{A}})$|⁠.
  • (4) If |${\mathcal{C}}-rE_{0,1}^{\theta }$| is diagonal, |$\textrm{col}({\mathcal{C}})=\textrm{row}({\mathcal{A}})$|⁠, and |$s_r({\mathcal{C}})=s_l({\mathcal{A}})$|⁠, then
    (A.5.4)

 

Proof.
Here we only prove parts (2) and (4) while omitting the easier parts (1) and (3). For part (2), let |${\mathcal{A}}=\kappa (\mu ^{\beta },g_2,\nu ^{\gamma })$|⁠, and let |$\delta =\delta ({\mathcal{B}})$|⁠. Take any |$t\in \Gamma _r$|⁠, we consider two cases: |$r<\frac{1}{2}\textrm{row}({\mathcal{A}})_0$| or |$r=\frac{1}{2}\textrm{row}({\mathcal{A}})_0$|⁠. Case 1: |$r<\frac{1}{2}\textrm{row}({\mathcal{A}})_0$|⁠: Let |$w_t$| be the minimal length element in the set |$\{ w\in \mathscr{D}_{\delta } \cap W_{\mu ^{\beta }} \mid{\mathcal{A}}^{(w)}= {{\mathcal{A}}}_{t,1}\}$|⁠. A direct computation shows that its length is given by
(A.5.5)
By a combinatorial argument, we calculate that
(A.5.6)
Note that
(A.5.7)
 
(A.5.7)
 
(A.5.9)
where (⁠|$\diamondsuit $|⁠) is due to the quantum binomial theorem |$\sum _{r=0}^n {n \brack r} v^{r(r-1)}x^r = \prod _{k=0}^{n-1}(1+v^{2k}x)$|⁠. Hence,
(A.5.8)
Moreover, using (A.4.6), we obtain
(A.5.9)
Combining Lemma A.5.3, (A.5.5), (A.5.8), and (A.5.9), we obtain that, if |$r<\frac{1}{2}\textrm{row}({\mathcal{A}})_0$|⁠,
Case 2: |$r=\frac{1}{2}\textrm{row}({\mathcal{A}})_0$|⁠: In this case, each term |$e_{ {{\mathcal{A}}}_{t,1}}=0$| unless |$a_{0,p}=t_p+t_{-p}$| for all |$p\in [-n, n]$|⁠. (Particularly, |$a^{\prime }_{0,0}=t_0$|⁠.) For the non-vanishing terms, we have
where |$x$| runs over all integers such that |$0\leqslant x\leqslant a^{\prime }_{0,0}$| and |$x+\sum _{p<0}t_p\in 2{\mathbb{N}}$|⁠. Note that
Hence,
(A.5.10)
Combining Lemma A.5.3, (A.5.5), (A.5.9), and (A.5.10), we obtain, if |$r=\frac{1}{2}\textrm{row}({\mathcal{A}})_0$|⁠,
Part (2) concludes.
For part (4), Let |${\mathcal{A}}=\kappa (\mu ^{\beta },g_2,\nu ^{\gamma })$|⁠, |$\delta =\delta ({\mathcal{C}})$| and take any |$t\in \Gamma _r$|⁠. Let |$w_t$| be the shortest element in the set |$\{w\in \mathscr{D}_{\delta } \cap W_{\mu ^{\beta }} \mid{\mathcal{A}}^{(w)}=\widehat{{\mathcal{A}}}_{t,1}\}$|⁠. Its length is given by
(A.5.11)
Moreover, using (A.4.6), we obtain
(A.5.12)
Combining Lemma A.5.3, (A.5.11), and (A.5.12), we finally get that

Take |$r=1$| in Proposition A.5.4, we have the following corollary.

 

Corollary A.5.5

Suppose that |${\mathcal{A}}=A^{\textrm{sgn}({\mathcal{A}})}, {\mathcal{B}}, {\mathcal{C}}\in \Xi _{\textbf{D}}$| and |$h\in [1,n]$|⁠.

  • (1) If |$h\neq 1$|⁠, |${\mathcal{B}}-E_{h,h-1}^{\theta }$| is diagonal, |$\textrm{col}({\mathcal{B}})=\textrm{row}({\mathcal{A}})$|⁠, and |$s_r({\mathcal{B}})=s_l({\mathcal{A}})$|⁠, then
    (A.5.13)
    where |${\mathcal{A}}_p=(A+E_{h,p}^{\theta }-E_{h-1,p}^{\theta },\textrm{sgn}(s_l({\mathcal{B}}),s_r({\mathcal{A}})))$|⁠.
  • (2) If |${\mathcal{B}}-E_{1,0}^{\theta }$| is diagonal, |$\textrm{col}({\mathcal{B}})=\textrm{row}({\mathcal{A}})$|⁠, and |$s_r({\mathcal{B}})=s_l({\mathcal{A}})$|⁠, then
    (A.5.14)
  • (3) If |$h\neq 1$|⁠, |${\mathcal{C}}-E_{h-1,h}^{\theta }$| is diagonal, |$\textrm{col}({\mathcal{C}})=\textrm{row}({\mathcal{A}})$|⁠, and |$s_r({\mathcal{C}})=s_l({\mathcal{A}})$|⁠, then
    (A.5.15)
    where |${\mathcal{A}}(h,p)=(A-E_{h,p}^{\theta }+E_{h-1,p}^{\theta },\textrm{sgn}(s_l({\mathcal{C}}),s_r({\mathcal{A}})))$|⁠.
  • (4) If |${\mathcal{C}}-E_{0,1}^{\theta }$| is diagonal, |$\textrm{col}({\mathcal{C}})=\textrm{row}({\mathcal{A}})$|⁠, and |$s_r({\mathcal{C}})=s_l({\mathcal{A}})$|⁠, then
    (A.5.16)

 

Remark A.5.6.
The multiplication formulas with |$e_{{\mathcal{A}}}$| (Proposition A.5.4 and Corollary A.5.5) match Fan–Li’s multiplication formulas ([14, Proposition 4.3.2 and Corollary 4.3.4].) with |$e^{\textrm{geo}}_{{\mathcal{A}}}$|⁠, via the following correspondence:
(A.5.17)

 

Remark A.5.7.

An immediate application of the multiplication formulas is to demonstrate a stabilization property for |$\{\textbf{S} _{n,d}\mid d\in{\mathbb{N}}\}$| and further construct an algebra |$\mathcal{K}_n$| so that the multiplication rules on |$\mathcal{K}_n$| are compatible with the rules on any |$\textbf{S} _{n,d}$|⁠. The algebras |$\mathcal{K}_n$| have been introduced by Fan and Li in loc. cit.

A.6 Schur duality

Let |$\mathfrak{g}$| be the simple Lie algebra of type |$\textbf{D}_d$|⁠, and let |$\rho $| be the half sum of the positive roots of |$\mathfrak{g}$|⁠. It was mentioned in a framework [25] that |$\Lambda _{\textbf{D}}$| can be viewed as the set of orbits of |$W$| on a (truncated) |$\rho $|-shifted weight lattice of |$\mathfrak{g}$|⁠. Then the |$v$|-tensor space |$\bigoplus _{\lambda ^{\alpha }\in \Lambda _{\textbf{D}}}x_{\lambda ^{\alpha }}\textbf{H}$| can be viewed as the quantum version of the Grothendieck groups of the category |$\mathcal{O}$| of |$\mathfrak{g}$|-modules.

This picture is also valid when |$\Lambda _{\textbf{D}}$| is replaced by its subset. Each subset |$\Lambda _f\subset \Lambda _{\textbf{D}}$| corresponds to a Schur algebra

A Schur duality is also obtained in loc. cit. for each pair |$(\textbf{S} _f, \textbf{H})$| on the tensor space |$\mathop{\oplus }_{\lambda \in \Lambda _f} x_{\lambda } \textbf{H}$|⁠.

 

Remark A.6.1.

If |$\Lambda _f=\Lambda ^+\sqcup \Lambda ^-$|⁠, then |$\textbf{S} _f$| is the algebra |$\mathcal{S}^m$| in [14, §6.1]. The stabilization procedure affords a different quantum algebra |$\mathcal{K}^m$| in loc. cit.

 

Remark A.6.2.

Fan and Li told the authors in private conversations that they have also been aware of the Schur algebra |$\textbf{S} _f$| and the related Schur duality for |$\Lambda _f=\Lambda ^+$| or |$\Lambda ^0\sqcup \Lambda ^+$| although they did not write it down.

Funding

This work was supported by the Science and Technology Commission of Shanghai Municipality (18dz2271000 to L.L.) and the National Nature Science Foundation of China (11671108, 11871214 to L.L.).

Acknowledgments

The authors thank Huanchen Bao and Weiqiang Wang for helpful discussions. We also thank Catherina Stroppel for bringing [13] to our attention. We thank the referees for detailed comments on a previous version of the manuscript.

Communicated by Prof. Weiqiang Wang

References

[1]

Bao
,
H.
Kazhdan–Lusztig theory of super type D and quantum symmetric pairs
.”
Represent. Theory
 
21
, (
2017
):
247
76
.

[2]

Björner
,
A.
, and
Brenti
,
F.
 
Combinatorics of Coxeter Groups
.
Graduate Texts in Mathematics 231
.
New York
:
Springer
,
2005
.

[3]

Bao
,
H.
,
Kujawa
,
J.
,
Li
,
Y.
, and
Wang
,
W.
Geometric Schur duality of classical type
.”
Transform. Groups
 
23
(
2018
):
329
89
.

[4]

A.
 
Beilinson
,
G.
 
Lusztig
and
R.
 
MacPherson
A geometric setting for the quantum deformation of |$G{L}\_n$|
.”
Duke Math. J
.
61
(
1990
):
655
677
.

[5]

Bao
,
H.
and
Wang
,
W.
A new approach to Kazhdan–Lusztig theory of type |$B$| via quantum symmetric pairs
.”
Asterisque
 
402
, (
2018
)
vii+134 pp
.

[6]

Bao
,
H.
and
Wang
,
W.
Canonical bases arising from quantum symmetric pairs
.”
Invent. Math.
 
213
(
2018
):
1099
177
.

[7]

Bao
,
H.
,
Wang
,
W.
, and
Watanabe
,
H.
Multiparameter quantum Schur duality of type B
.”
Proc. Amer. Math. Soc.
 
146
(
2018
):
3203
16
.

[8]

Curtis
,
C.
,
Iwahori
,
N.
, and
Kilmoyer
,
R.
Hecke algebras and characters of parabolic type of finite group with |$\left (B,N\right )$|-pairs
.”
Publ. Math. IHES
 
40
(
1972
):
81
116
.

[9]

Curtis
,
C.
On Lusztig’s isomorphism theorem for Hecke algebras
.”
J. Algebra
 
92
(
1985
):
348
65
.

[10]

Deng
,
B.
,
Du
,
J.
,
Parshall
,
B.
, and
Wang
,
J.
 
Finite Dimensional Algebras and Quantum Groups
.
Mathematical Surveys and Monographs 150
.
Providence, RI
:
American Mathematical Society
,
2008
.

[11]

Du
,
J.
and
Fu
,
Q.
Quantum affine |${\mathfrak{g}\mathfrak{l}}_{\textrm{n}}$| via Hecke algebra
.”
Adv. Math.
 
282
(
2015
):
23
46
.

[12]

Du
,
J.
Kazhdan–Lusztig bases and isomorphism theorems for |$q$|-Schur algebras
.”
Contemp. Math.
 
139
(
1992
):
121
40
.

[13]

Ehrig
,
M.
and
Stroppel
,
C.
Nazarov–Wenzl algebras, coideal subalgebras and categorified skew Howe duality
.”
Adv. Math.
 
331
(
2018
):
58
131
.

[14]

Fan
,
Z.
and
Li
,
Y.
Geometric Schur duality of classical type, II
.”
Trans. Amer. Math. Soc. Ser. B
 
2
(
2015
):
51
92
.

[15]

Fan
,
Z.
,
Lai
,
C.
,
Li
,
Y.
,
Luo
,
L.
, and
Wang
,
W.
Affine flag varieties and quantum symmetric pairs
.”
Mem. Amer. Math. Soc.
(
forthcoming
)
preprint arXiv:1602.04383
.

[16]

Fan
,
Z.
,
Lai
,
C.
,
Li
,
Y.
,
Luo
,
L.
, and
Wang
,
W.
Affine Hecke algebras and quantum symmetric pairs
.”
preprint arXiv:1609.06199
.

[17]

Green
,
R.
Hyperoctahedral Schur algebras
.”
J. Algebra
 
192
(
1997
):
418
38
.

[18]

Kolb
,
S.
Quantum symmetric Kac–Moody pairs
.”
Adv. Math.
 
267
(
2014
):
395
469
.

[19]

Kazhdan
,
D.
and
Lusztig
,
G.
Representations of Coxeter groups and Hecke algebras
.”
Invent. Math.
 
53
(
1979
):
165
84
.

[20]

Letzter
,
G.
Symmetric pairs for quantized enveloping algebras
.”
J. Algebra
 
220
(
1999
):
729
67
.

[21]

Letzter
,
G.
Coideal subalgebras and quantum symmetric pairs
.”
New Directions in Hopf Algebras (Cambridge)
.
MSRI Publications
, Vol.
43
.
117
66
.
Cambridge University Press
,
2002
.

[22]

Lusztig
,
G.
 
Introduction to Quantum Groups
.
Progress in Mathematics 10
.
Boston, MA
:
Birkhäuser Boston, Inc.
,
1993
.

[23]

G.
 
Lusztig
Aperiodicity in quantum affine
|${\mathfrak{g}{\mathfrak{l}}}_n$|⁠.”
Asian J. Math
.
3
(
1999
):
147
177
.

[24]

Lusztig
,
G.
 
Hecke Algebras with Unequal Parameters
.
CRM Monograph Series 18
.
Providence, RI
:
American Mathematical Society
,
2003
.

[25]

Luo
,
L.
and
Wang
,
W.
The |$q$|-Schur algebras and |$q$|-Schur dualities of finite type
.”
preprint arXiv:1710.10375
.

This article is published and distributed under the terms of the Oxford University Press, Standard Journals Publication Model (https://academic.oup.com/journals/pages/open_access/funder_policies/chorus/standard_publication_model)