Hostname: page-component-7bb8b95d7b-l4ctd Total loading time: 0 Render date: 2024-09-27T02:45:23.314Z Has data issue: false hasContentIssue false

Laurent family of simple modules over quiver Hecke algebras

Published online by Cambridge University Press:  11 September 2024

Masaki Kashiwara
Affiliation:
Kyoto University Institute for Advanced Study, Research Institute for Mathematical Sciences, Kyoto University, Kyoto 606-8502, Japan and Korea Institute for Advanced Study, Seoul 02455, South Korea masaki@kurims.kyoto-u.ac.jp
Myungho Kim
Affiliation:
Department of Mathematics, Kyung Hee University, Seoul 02447, South Korea mkim@khu.ac.kr
Se-jin Oh
Affiliation:
Department of Mathematics, Sungkyunkwan University, Suwon 16419, South Korea sejin092@gmail.com
Euiyong Park
Affiliation:
Department of Mathematics, University of Seoul, Seoul 02504, South Korea epark@uos.ac.kr
Rights & Permissions [Opens in a new window]

Abstract

We introduce the notions of quasi-Laurent and Laurent families of simple modules over quiver Hecke algebras of arbitrary symmetrizable types. We prove that such a family plays a similar role of a cluster in quantum cluster algebra theory and exhibits a quantum Laurent positivity phenomenon similar to the basis of the quantum unipotent coordinate ring $\mathcal {A}_q(\mathfrak {n}(w))$, coming from the categorification. Then we show that the families of simple modules categorifying Geiß–Leclerc–Schröer (GLS) clusters are Laurent families by using the Poincaré–Birkhoff–Witt (PBW) decomposition vector of a simple module $X$ and categorical interpretation of (co)degree of $[X]$. As applications of such $\mathbb {Z}\mspace {1mu}$-vectors, we define several skew-symmetric pairings on arbitrary pairs of simple modules, and investigate the relationships among the pairings and $\Lambda$-invariants of $R$-matrices in the quiver Hecke algebra theory.

Type
Research Article
Copyright
© 2024 The Author(s). The publishing rights in this article are licensed to Foundation Compositio Mathematica under an exclusive licence

1. Introduction

A cluster algebra and its non-commutative version quantum cluster algebra, were introduced by Berenstein, Fomin and Zelevinsky [Reference Fomin and ZelevinskyFZ02, Reference Berenstein and ZelevinskyBZ05] in an attempt to provide an algebraic and combinatorial framework for investigating the upper global basis of the quantum group.

The quantum cluster algebra $\mathscr {A}_q$ is a non-commutative $\mathbb {Z}\mspace {1mu}[q^{\pm 1/2}]$-subalgebra in the skew field $\mathbb {Q}(q^{1/2})(X_i)_{i \in \mathsf {K}}$ generated by the cluster variables, which are obtained from the initial cluster $\{ X_{i} \}_{i \in \mathsf {K}}$ via the sequences of procedures, called mutations. Even though mutations involve non-trivial fractions, $\mathscr {A}_q$ is still contained in $\mathbb {Z}\mspace {1mu}[q^{\pm 1/2}][X_i^{\pm 1}]_{i \in \mathsf {K}}$ with amazing reductions of fractions which is referred to as the quantum Laurent phenomenon [Reference Berenstein and ZelevinskyBZ05]. The famous conjecture, which is not completely proved yet at this moment, is the quantum Laurent positivity conjecture which asserts that every cluster variable is an element in $\mathbb {Z}\mspace {1mu}_{\geqslant 0}[q^{\pm 1/2}][\mathsf {X}_i^{\pm 1}]_{i \in \mathsf {K}}$ for any cluster $\{\mathsf {X}_i\}_{i \in \mathsf {K}}$. Note that the conjecture is proved in [Reference DavisonDav18] (see also [Reference Lee and SchifflerLS15, Reference Gross, Hacking, Keel and KontsevichGHKK18]) when $\mathscr {A}_q$ is of skew-symmetric type and is widely open when it is of non-skew-symmetric type.

The notion of monoidal categorification of (quantum) cluster algebra was introduced by Hernandez and Leclerc in [Reference Hernandez and LeclercHL10] (see also [Reference Kang, Kashiwara, Kim and OhKKKO18]) as the categorical framework for proving the conjecture as follows: a monoidal category $\mathcal {C}$ with an autofunctor $q$ is a monoidal categorification of $\mathscr {A}_q$, if (a) $\mathbb {A} \otimes _{\mathbb {Z}\mspace {1mu}[q^{\pm 1}]} K(\mathcal {C})$ $(\mathbb {A} \mathbin {:=}\mathbb {Z}\mspace {1mu}[q^{\pm 1/2}])$ is isomorphic to $\mathscr {A}_q$ and (b) the cluster monomials of $\mathscr {A}_q$ are the classes of real simple objects of $\mathcal {C}$. Once $\mathcal {C}$ is a monoidal categorification of $\mathscr {A}_q$, then the conjecture for $\mathscr {A}_q$ follows since it can be interpreted as the existence of a Jordan–Hölder series of an object. In [Reference Kang, Kashiwara, Kim and OhKKKO18], it is proved that the category $\mathscr {C}_w$ over symmetric quiver Hecke algebra $\mathsf {R}$ is a monoidal categorification of the quantum unipotent coordinate ring $\mathcal {A}_\mathbb {A}(\mathfrak {n}(w))$ associated with an element $w$ of the Weyl group $\mathsf {W}$ by using the $\mathbb {Z}\mspace {1mu}$-invariant $\Lambda (M,N)$ of a pair of simple objects $M,N \in \mathscr {C}_w$.

For non-symmetric cases, the monoidal categorification is still out of reach. We know that $\mathscr {C}_w$ categorifies $\mathcal {A}_\mathbb {A}(\mathfrak {n}(w))$ as an algebra [Reference Khovanov and LaudaKL09, Reference Khovanov and LaudaKL11, Reference RouquierRou08, Reference KimuraKim12] and $\mathcal {A}_\mathbb {A}(\mathfrak {n}(w))$ has a quantum cluster algebra structure [Reference Geiss, Leclerc and SchröerGLS13a, Reference Goodearl and YakimovGY17] in every symmetrizable case. The quantum cluster algebra structure is skew-symmetric if the corresponding generalized Cartan matrix is symmetric. However, we cannot prove that $\mathscr {C}_w$ is a monoidal categorification of $\mathcal {A}_\mathbb {A}(\mathfrak {n}(w))$ in non-symmetric cases due to the obstacle that we do not know whether every simple module $M \in \mathscr {C}_w$ admits an affinization [Reference Kashiwara and ParkKP18] or not. Note that the existence of affinizations guarantees that one can define $R$-matrices and the $\mathbb {Z}\mspace {1mu}$-invariant $\Lambda (M,N)$.

In this paper, we study the quantum Laurent positivity for $\mathcal {A}_\mathbb {A}(\mathfrak {n}(w))$ of not necessarily symmetric type in the view point of the monoidal categorification. More precisely, we show that the basis of $\mathcal {A}_\mathbb {A}(\mathfrak {n}(w))$ corresponding to the simple modules in $\mathscr {C}_w$ exhibits a quantum Laurent positivity phenomenon with respect to any quasi-Laurent family, which is a central notion we introduce in this paper and plays the similar role of a cluster in the quantum cluster algebra theory.

The quasi-Laurent family (respectively, Laurent family) $\mathcal {M}=\{ M_j \}_{j \in J}$ consists of mutually commuting affreal simple modules in $\mathscr {C}_w$ satisfying additional conditions (Definition 3.2). Among others, the most important condition is that if a simple module $X$ commutes with all $M_j$, then there are monomials (i.e. convolution products) $\mathcal {M}(\boldsymbol a)$ and $\mathcal {M}(\boldsymbol b)$ in $\{M_j\}_{j\in J}$ such that $X \circ \mathcal {M}(\boldsymbol a)$ is isomorphic to $\mathcal {M}(\boldsymbol b)$. We say the family $\mathcal {M}$ is Laurent if $\mathcal {M}$ is maximal in the sense that, if a simple module $X$ commutes with all $M_j$, then $X$ is isomorphic to a monomial $\mathcal {M}(\boldsymbol b)$ in $\{M_j\}_{j\in J}$.

The main results of this paper are the following.

  1. (A) We show that if $\mathcal {M}$ is a quasi-Laurent family in $\mathscr {C}_w$, then the class $[X]$ in the Grothendieck ring $K(\mathscr {C}_w)$ of any simple object $X$ in $\mathscr {C}_w$ can be written as a Laurent polynomial in $\{ [M_j] \}_{j \in J}$ whose coefficients belong to $\mathbb {Z}\mspace {1mu}_{\ge 0}[q^{\pm 1}]$ (Proposition 3.6).

  2. (B) If $\mathcal {M}$ is a monoidal seed in $\mathscr {C}_w$, then $\mathcal {M}$ is a Laurent family.

  3. (C) In particular, for any reduced sequence ${\boldsymbol{i}}$ of $w$, we show that the family $\mathcal {M}^{\boldsymbol{i}} \mathbin {:=} \{ M(w_{\leqslant k}^{\boldsymbol {i}} \varpi _{i_k},\varpi _{i_k}) \}$ is a Laurent family and, hence, any class $[X]$ of a module $X$ in $\mathscr {C}_w$ can be written as a Laurent polynomial in the unipotent quantum minors $D(w_{\leqslant k}^{\boldsymbol {i}} \varpi _{i_k},\varpi _{i_k})$ with coefficients in $\mathbb {Z}\mspace {1mu}_{\ge 0}[q^{\pm 1}]$. Note that $D(w_{\leqslant k}^{\boldsymbol {i}} \varpi _{i_k},\varpi _{i_k})=[M(w_{\leqslant k}^{\boldsymbol {i}} \varpi _{i_k},\varpi _{i_k}) ]$ and we call $\{D(w_{\leqslant k}^{\boldsymbol {i}} \varpi _{i_k},\varpi _{i_k})\}$ the GLS seed associated with ${\boldsymbol {i}}$ (Proposition 4.5).

  4. (D) We show that if $\mathcal {M}$ is a quasi-Laurent family, then the class $[X]$ of a simple module $X$ in $\mathscr {C}_w$ is pointed and copointed with respect to the partial order $\preccurlyeq _\mathcal {M}$. That is, the set of degrees of the monomials appearing in the Laurent expansion of $[X]$ with respect to $\mathcal {M}$ has a unique maximal element and a unique minimal element with respect to $\preccurlyeq _\mathcal {M}$. We define vectors $\mathbf {g}_\mathcal {M}^R(X)$ and $\mathbf {g}_\mathcal {M}^L(X) \in \mathbb {Z}\mspace {1mu}^{\oplus J}$ as the maximal and the minimal element, respectively.

  5. (E) Each quasi-Laurent family $\mathcal {M}$ also induces new $\mathbb {Z}\mspace {1mu}$-values ${\mathrm G}_\mathcal {M}^R(X,Y)$ and ${\mathrm G}_\mathcal {M}^L(X,Y)$ for any pair of simple modules $X$ and $Y$ which coincides with $\Lambda (X,Y)$ provided $X,Y$ commutes and one of them is affreal.

To the best of the authors’ knowledge, the positivity result in part (C) is new. We can understand result (A) that a quasi-Laurent family is a generalization of a cluster in the categorical view point, and that the positivity conjecture can be extended to elements corresponding to simple modules in all skew-symmetrizable types.

In [Reference Fomin and ZelevinskyFZ07, Reference QinQin17], Fomin-Zelevinsky and Qin defined a pointed (respectively, copointed) element $\mathbf {x}$ in a cluster algebra and its degree $\mathbf {deg}_{\mathcal {S}}(\mathbf {x}) \in \mathbb {Z}\mspace {1mu}^{\oplus \mathsf {K}}$ (respectively, codegree $\mathbf {codeg}_{\mathcal {S}}(\mathbf {x}) \in \mathbb {Z}\mspace {1mu}^{\oplus \mathsf {K}}$) depending on the choice of a seed $\mathcal {S}$ (see also [Reference QinQin20] for codegree and [Reference TranTra11] for degree elements in a quantum cluster algebra). With a fixed choice of a seed, it is proved in [Reference TranTra11] that every cluster monomial is pointed, and in [Reference Derksen, Weyman and ZelevinskyDWZ10, Reference Gross, Hacking, Keel and KontsevichGHKK18] that cluster monomials are determined by their degrees.

For a given quasi-Laurent family $\mathcal {M}$ and a simple module $X \in \mathscr {C}_w$, we define vectors $\mathbf {g}_\mathcal {M}^R(X),\mathbf {g}_\mathcal {M}^L(X) \in \mathbb {Z}\mspace {1mu}^{\oplus J }$ in Definition 3.7 by using the $\mathbb {Z}\mspace {1mu}^{\oplus J}_{\geqslant 0}$-vectors in Lemma 3.3 and guaranteeing its well-definedness in Lemma 3.1. We then prove that, for every simple module $X \in \mathscr {C}_w$, the element $[X]$ in $\mathcal {A}_\mathbb {A}(\mathfrak {n}(w))$ is (co)pointed with respect to the GLS seed $\mathcal {S}^{\boldsymbol {i}}$ and that $\mathbf {g}_{\mathcal {M}^{{\boldsymbol {i}}}}^R(X)$ and $\mathbf {g}_{\mathcal {M}^{{\boldsymbol {i}}}}^L(X)$ coincide with $\mathbf {deg}_{\mathcal {S}^{\boldsymbol {i}}}([X])$ and $\mathbf {codeg}_{\mathcal {S}^{\boldsymbol {i}}}([X])$, respectively.

Utilizing the vectors $\mathbf {g}_\mathcal {M}^R(X)$ and $\mathbf {g}_\mathcal {M}^L(X)$, we define skew-symmetric $\mathbb {Z}\mspace {1mu}$-valued forms ${\mathrm G}_\mathcal {M}^R(-,-)$ and ${\mathrm G}_\mathcal {M}^L(-,-)$ on the pairs $(X,Y)$ of simple modules. Then we compare ${\mathrm G}_\mathcal {M}^R(X,Y)$ and ${\mathrm G}_\mathcal {M}^L(X,Y)$ with the $\mathbb {Z}\mspace {1mu}$-invariant $\Lambda (X,Y)$ when the pair of simple module $(X,Y)$ admits the $\mathbb {Z}\mspace {1mu}$-invariant $\Lambda (X,Y)$. It is proved in Proposition 5.3 that ${\mathrm G}_\mathcal {M}^R(X,Y)$ and ${\mathrm G}_\mathcal {M}^L(X,Y)$ give lower bounds of $\Lambda (X,Y)$, and in Proposition 5.4 that ${\mathrm G}_\mathcal {M}^R(X,Y)={\mathrm G}_\mathcal {M}^L(X,Y)=\Lambda (X,Y)$ when $(X,Y)$ is a commuting pair. Here we would like to emphasize that (1) ${\mathrm G}_\mathcal {M}^R(X,Y)$ and ${\mathrm G}_\mathcal {M}^L(X,Y)$ are defined even for pairs $(X,Y)$ we do not know whether they admit $\Lambda (X,Y)$ or not, and (2) the $\mathbb {Z}\mspace {1mu}$-values ${\mathrm G}_\mathcal {M}^R(X,Y)$ and ${\mathrm G}_\mathcal {M}^L(X,Y)$ do depend on the choice of $\mathcal {M}$ as (co)degree does on the one of seeds (Remark 5.5).

This paper is organized as follows. In § 2, we give preliminaries. In § 3, we define the notions of quasi-Laurent and Laurent families, and investigate their properties. Then we define $\mathbf {g}_{\mathcal {M}}^R(X)$ and $\mathbf {g}_{\mathcal {M}}^L(X)$, and prove that $\mathbf {g}_{\mathcal {M}}^R(X)$ and $\mathbf {g}_{\mathcal {M}}^L(X)$ determine the isomorphism class of $X$. In § 4, we prove that $\mathcal {M}^{\boldsymbol {i}}$ is Laurent by studying PBW decomposition vectors of simple modules. In § 5, we define the skew-symmetric pairings on pairs of simple modules and investigate the relationships among the pairings and $\Lambda$-invariants.

Convention Throughout this paper, we use the following convention.

  1. (i) For a statement $\mathtt {P}$, we set $\delta (\mathtt {P})$ to be $1$ or $0$ depending on whether $\mathtt {P}$ is true or not. As a special case, we use the notation $\delta _{i,j} \mathbin {:=} \delta (i=j)$ (Kronecker's delta).

  2. (ii) For integers $a,b \in \mathbb {Z}\mspace {1mu}$, we set

    \[ [a,b] \mathbin{:=} \{ x\in \mathbb{Z}\mspace{1mu} \; | \; a \leqslant x \leqslant b\}. \]
    We refer to the subset as an interval and understand it as an empty set if $a>b$.
  3. (iii) Let $\mathbf {x}=(x_j )_{ j \in J }$ be a family parameterized by an index set $J$. Then for any $j \in J$, we set

    \[ (\mathbf{x})_j \mathbin{:=} x_j. \]

2. Preliminaries

In this preliminary section, we briefly review the basic material of this paper. We refer the reader to [Reference Berenstein and ZelevinskyBZ05, Reference Fomin and ZelevinskyFZ07, Reference Khovanov and LaudaKL09, Reference RouquierRou08, Reference KimuraKim12, Reference Geiss, Leclerc and SchröerGLS13a, Reference Kang, Kashiwara, Kim and OhKKKO18, Reference Goodearl and YakimovGY17, Reference Kimura and OyaKiOy21, Reference Kashiwara and ParkKP18, Reference Kashiwara, Kim, Oh and ParkKKOP18, Reference Kashiwara and KimKK19, Reference Gross, Hacking, Keel and KontsevichGHKK18] for more details.

2.1 Quantum cluster algebras

Fix a finite index set $\mathsf {K} = {\mathsf {K}_\mathrm {ex}} \sqcup {\mathsf {K}_\mathrm {fr}}$ with a decomposition into the set ${\mathsf {K}_\mathrm {ex}}$ of exchangeable indices and the set ${\mathsf {K}_\mathrm {fr}}$ of frozen indices. Let $L = (l_{ij})_{i,j \in \mathsf {K}}$ be a skew-symmetric integer-valued matrix and let $q$ be an indeterminate. We set $\mathbb {A} \mathbin {:=} \mathbb {Z}\mspace {1mu}[q^{\pm 1/2}]$ where $q^{1/2}$ denotes the formal square root of $q$.

Definition 2.1 We define the quantum torus $\mathcal {T}(L)$ to be the $\mathbb {A}$-algebra generated by a finite family of elements $\{X^{\pm 1}_k \}_{k \in \mathsf {K}}$ subject to the following defining relations:

\[ X_jX_j^{-1} = X_j^{-1}X_j=1 \quad\text{and}\quad X_iX_j = q^{l_{ij}}X_jX_i \quad \text{for $i,j \in \mathsf{K}$.} \]

For $\mathbf {a} = (\mathbf {a}_i)_{i \in \mathsf {K}} \in \mathbb {Z}\mspace {1mu}^\mathsf {K}$, we define the element $X^\mathbf {a}$ of $\mathcal {T}(L)$ as

\[ X^\mathbf{a} = q^{({1/2}) \sum_{i>j}\mathbf{a}_i\mathbf{a}_j l_{ij}} \prod^{\longrightarrow}_{i \in \mathsf{K}} X_i^{\mathbf{a}_i}. \]

Here $\overset {\longrightarrow }{\prod }_{i \in \mathsf {K}} X_i^{\mathbf {a}_i} \mathbin {:=} X_{i_1}^{\mathbf {a}_{i_1}} \cdots X_{i_r}^{\mathbf {a}_{i_r}}$, where $\mathsf {K}=\{ i_1,\ldots,i_r \}$ with a total order $i_1 < \cdots < i_r$. Note that $X^\mathbf {a}$ does not depend on the choice of a total order < on $\mathsf {K}$. Then $\{ X^\mathbf {a} \mid \mathbf {a} \in \mathbb {Z}\mspace {1mu}^{\mathsf {K}} \}$ forms an $\mathbb {A}$-basis of $\mathcal {T}(L)$. Since $\mathcal {T}(L)$ is an Ore domain, it is embedded into the skew field of fractions $\mathbb {F}(\mathcal {T}(L))$.

Let $\widetilde {B} = (b_{ij})_{i \in \mathsf {K},j\in {\mathsf {K}_\mathrm {ex}}}$ be an integer-valued $\mathsf {K} \times {\mathsf {K}_\mathrm {ex}}$-matrix whose principal part $B = (b_{ij})_{i,j\in {\mathsf {K}_\mathrm {ex}}}$ is skew-symmetrizable, i.e. there exists a diagonal matrix $D$ with a positive integer entries such that $DB$ is skew-symmetric. Such a matrix $\widetilde {B}$ is called an exchange matrix. We say that a pair $(L,B)$ is compatible if

\[ \sum_{k \in \mathsf{K}} b_{ki} l_{kj} = d_i\delta_{i,j}\quad \text{for any $i\in {\mathsf{K}_\mathrm{ex}}$ and $j\in\mathsf{K}$} \]

for some positive integers $\{ d_i \}_{i \in {\mathsf {K}_\mathrm {ex}}}$. We call the triple $\mathcal {S} = (\{ X_k \}_{k \in \mathsf {K}}, L , \widetilde {B} )$ a quantum seed in the quantum torus $\mathcal {T}(L)$ and $\{ X_k \}_{k \in \mathsf {K}}$ a quantum cluster.

For $k \in {\mathsf {K}_\mathrm {ex}}$, the mutation $\mu _k(L,\widetilde {B}) \mathbin {:=} (\mu _k(L),\mu _k(B))$ of a compatible pair $(L,\widetilde {B})$ in a direction $k$ is defined in a combinatorial way (see [Reference Berenstein and ZelevinskyBZ05]). Note that (i) the pair $(\mu _k(L),\mu _k(B))$ is also compatible with the same positive integers $\{d_i\}_{i \in \mathsf {K}}$ and (ii) the operation $\mu _k$ is an involution, i.e. $\mu _k(\mu _k(L,\widetilde {B})) = (L,\widetilde {B})$. We define an isomorphism of $\mathbb {Q}(q^{1/2})$-algebras $\mu _k^* \colon \mathbb {F}(\mathcal {T}(\mu _kL)) {\mathop\longrightarrow\limits^{\sim}} \mathbb {F}(\mathcal {T}(L))$ by

\[ \mu_k^*(X_j) \mathbin{:=} \begin{cases} X^{\mathbf{a}'} + X^{\mathbf{a}''} & \text{if } j =k, \\ X_j & \text{if } j \ne k, \end{cases} \]

where

\[ \mathbf{a}'_i = \begin{cases} -1 & \text{if } i =k, \\ \max(0,b_{ik}) & \text{if } i \ne k, \end{cases} \quad\text{and}\quad \mathbf{a}''_i = \begin{cases} -1 & \text{if } i =k, \\ \max(0,-b_{ik}) & \text{if } i \ne k. \end{cases} \]

Then the mutation $\mu _k(\mathcal {S})$ of the quantum seed $\mathcal {S}$ in a direction $k$ is defined to be the triple $( \{ X_i\}_{i \ne k} \sqcup \{ \mu _k^*(X_k) \}, \mu _k(L), \mu _k(\widetilde {B}))$.

For a quantum seed $\mathcal {S}=( \{X_k\}_{k\in \mathsf {K}},L,\widetilde {B})$, an element in $\mathbb {F}(\mathcal {T}(L))$ is called a quantum cluster variable (respectively, quantum cluster monomial) if it is of the form

\[ \mu_{k_1}^* \cdots \mu_{k_\ell}^*(X_j) \quad \text{$($respectively, } \mu_{k_1}^* \cdots \mu_{k_\ell}^*(X^\mathbf{a}) ) \]

for some finite sequence $(k_1,\ldots,k_\ell ) \in \mathsf {K}_\mathrm {ex}^\ell$ $(\ell \in \mathbb {Z}\mspace {1mu}_{\geqslant 0})$ and $j \in \mathsf {K}$ (respectively, $\mathbf {a} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^\mathsf {K}$). For a quantum seed $\mathcal {S} = (\{ X_k \}_{k \in \mathsf {K}}, L,\widetilde {B} )$, the quantum cluster algebra $\mathscr {A}_q(\mathcal {S})$ is the $\mathbb {A}$-subalgebra of $\mathbb {F}(\mathcal {T}(L))$ generated by all the quantum cluster variables. Note that $\mathscr {A}_q(\mathcal {S})\simeq \mathscr {A}_q(\boldsymbol {\mu }(\mathcal {S}))$ for any sequence $\boldsymbol {\mu }$ of mutations.

The quantum Laurent phenomenon, proved by Berenstein and Zelevinsky in [Reference Berenstein and ZelevinskyBZ05], says that the quantum cluster algebra $\mathscr {A}_q(\mathcal {S})$ is indeed contained in $\mathcal {T}(L)$.

For a quantum seed $\mathcal {S}$ with a compatible pair $(L,\widetilde {B})$, an element $\mathbf {x} \in \mathcal {T}(L)$ is called pointed (respectively, copointed) if it is of the following form:

(2.1)\begin{equation} \mathbf{x} = q^{a} X^{\mathbf{g}^R} + \sum_{ \mathbf{c} \in \mathbb{Z}\mspace{1mu}_{\geqslant 0}^{{\mathsf{K}_\mathrm{ex}}} \setminus \{ 0 \} } p_\mathbf{c} X^{\mathbf{g}^R+\widetilde{B} \mathbf{c}} \quad \bigg(\text{respectively, } \mathbf{x} = q^{a} X^{\mathbf{g}^L} + \sum_{ \mathbf{c} \in \mathbb{Z}\mspace{1mu}_{\leqslant 0}^{{\mathsf{K}_\mathrm{ex}}} \setminus \{ 0 \} } p_\mathbf{c} X^{\mathbf{g}^L+\widetilde{B} \mathbf{c}}\bigg) \end{equation}

for some $a \in \frac {1}{2}\mathbb {Z}\mspace {1mu}$, $\mathbf {g}^R \in \mathbb {Z}\mspace {1mu}^\mathsf {K}$ (respectively, $\mathbf {g}^L \in \mathbb {Z}\mspace {1mu}^\mathsf {K}$) and $p_\mathbf {c} \in \mathbb {A}$. In this case, we call $\mathbf {g}^R$ the degree (respectively, codegree) of the pointed (respectively, copointed) element $\mathbf {x}$ and denote it by $\mathbf {deg}_{\mathcal {S}}(\mathbf {x})$ (respectively, $\mathbf {codeg}_{\mathcal {S}}(\mathbf {x})$). The degree (respectively, codegree) is often the called $g$-vector (respectively, dual $g$-vector) of $\mathbf {x}$ (see [Reference QinQin17, Definition 3.1.4] and [Reference QinQin20, Definition 3.1.3]). It is worth remarking that the notion of $g$-vector (respectively, dual $g$-vector) does depend on the compatible pair $(L,\widetilde {B})$ and, hence, on the seed $\mathcal {S}$. It is proved in [Reference TranTra11, Theorem 5.3] that every quantum cluster monomial in $\mathscr {A}_q(\mathcal {S})$ is pointed.

We say that an $\mathbb {A}$-algebra $R$ has a quantum cluster algebra structure if there exists a quantum seed $\mathcal {S}$ and an $\mathbb {A}$-algebra isomorphism $\Omega : \mathscr {A}_q(\mathcal {S}) {\mathop\longrightarrow\limits^{\sim}} R$. In the case, a quantum seed of $R$ refers to the image of a quantum seed in $\mathscr {A}_q(\mathcal {S})$, which is obtained by a sequence of mutations.

2.2 Quantum unipotent coordinate rings

Let $I$ be an index set. A Cartan datum $(\mathsf {A},\mathsf {P},\Pi,\mathsf {P}^\vee,\Pi ^\vee )$ consists of:

  1. (a) a symmetrizable Cartan matrix $\mathsf {A}=(a_{i,j})_{i,j \in I}$, i.e. $\mathsf {D} \mathsf {A}$ is symmetric for a diagonal matrix $\mathsf {D} = {\rm diag}(\mathsf {d}_i \mid i \in I)$ with $\mathsf {d}_i\in \mathbb {Z}\mspace {1mu}_{>0}$;

  2. (b) a free abelian group $\mathsf {P}$, called the weight lattice;

  3. (c) $\Pi = \{{\mspace {1mu}\alpha }_i \mid i \in I \} \subset \mathsf {P}$, called the set of simple roots;

  4. (d) $\Pi ^\vee = \{ h_i \mid i \in I \} \subset \mathsf {P}^\vee \mathbin {:=} \operatorname {Hom}(\mathsf {P},\mathbb {Z}\mspace {1mu})$, called the set of simple coroots;

  5. (e) a $\mathbb {Q}$-valued symmetric bilinear form $( \cdot,\cdot )$ on $\mathsf {P}$;

satisfying the standard properties (see [Reference Kang, Kashiwara, Kim and OhKKKO18, § 1.1] for instance). Here we take $\mathsf {D} ={\rm diag}(\mathsf {d}_i \mid i \in I)$ such that $\mathsf {d}_i \mathbin {:=}({\mspace {1mu}\alpha }_i,{\mspace {1mu}\alpha }_i)/2\in \mathbb {Z}\mspace {1mu}_{>0}$ $(i\in I)$ in this paper.

For $i \in I$, we choose $\varpi _i \in \mathsf {P}$ such that $\langle h_i,\varpi _j \rangle =\delta _{ij}$ for any $j \in I$ and call it the $i$th fundamental weight. The free abelian group $\mathsf {Q} \mathbin {:=} \bigoplus _{i \in I} \mathbb {Z}\mspace {1mu} {\mspace {1mu}\alpha }_i$ is called the root lattice and we set $\mathsf {Q}^+ = \sum _{i \in I}\mathbb {Z}\mspace {1mu}_{\geqslant 0} {\mspace {1mu}\alpha }_i \subset \mathsf {Q}$ and $\mathsf {Q}^- = \sum _{i \in I}\mathbb {Z}\mspace {1mu}_{\leqslant 0} {\mspace {1mu}\alpha }_i \subset \mathsf {Q}$. We denote by $\Delta$ the set of roots and by $\Delta ^\pm$ the set of positive roots (respectively, negative roots). For ${\mspace {1mu}\beta } \in \sum _{i \in I} m_i{\mspace {1mu}\alpha }_i \in \mathsf {Q}^+$, we set $|{\mspace {1mu}\beta }| \mathbin {:=} \sum _{i \in I} m_i$, ${\rm supp}({\mspace {1mu}\beta }) \mathbin {:=}\{ i \in I \mid m_i \ne 0 \}$ and $I^{\mspace {1mu}\beta } \mathbin {:=} \{ \nu = (\nu _1, \ldots,\nu _{|{\mspace {1mu}\beta }|}) \in I^{|{\mspace {1mu}\beta }|} \mid {\mspace {1mu}\alpha }_{\nu _1}+ \cdots + {\mspace {1mu}\alpha }_{\nu _{|{\mspace {1mu}\beta }|}} ={\mspace {1mu}\beta } \}$. Note that the symmetric group $\mathfrak {S}_{|{\mspace {1mu}\beta }|} = \langle r_{1},\ldots, r_{|{\mspace {1mu}\beta }|} \rangle$ acts on $I^{\mspace {1mu}\beta }$ by the place permutations.

Let $\mathfrak {g}$ be the Kac–Moody algebra associated with the Cartan datum $(\mathsf {A},\mathsf {P},\Pi,\mathsf {P}^\vee,\Pi ^\vee )$, and $\mathsf {W}$ the Weyl group of $\mathfrak {g}$. It is generated by the simple reflections $s_i\in \operatorname {Aut}(\mathsf {P})$ $(i \in I)$ defined by $s_i(\lambda ) = \lambda - \langle h_i,\lambda \rangle {\mspace {1mu}\alpha }_i$ for $\lambda \in \mathsf {P}$. For a sequence ${\boldsymbol {i}}=(i_1,\ldots,i_r) \in I^r$, we call it a reduced sequence of $w \in \mathsf {W}$ if $s_{i_1} \ldots s_{i_r}$ is a reduced expression of $w$. For $w,v \in \mathsf {W}$, we write $w \geqslant v$ if there is a reduced sequence of $v$ which appears in a reduced sequence of $w$ as a subsequence.

For $\lambda,\mu \in \mathsf {P}$, we write $\lambda \preccurlyeq \mu$ if there exists a sequence of real positive roots ${\mspace {1mu}\beta }_k$ $(1 \leqslant k \leqslant l)$ such that $\lambda = s_{{\mspace {1mu}\beta }_l} \cdots s_{{\mspace {1mu}\beta }_1} \mu$ and $({\mspace {1mu}\beta }_k,s_{{\mspace {1mu}\beta }_{k-1}} \cdots s_{{\mspace {1mu}\beta }_1}\mu )>0$ for $1 \leqslant k \leqslant l$. When $\Lambda \in \mathsf {P}^+$ and $\lambda,\mu \in \mathsf {W}\Lambda$ the relation $\lambda \preccurlyeq \mu$ holds if and only if there exist $w,v \in \mathsf {W}$ such that $\lambda = w\Lambda$, $\mu = v\Lambda$ and $v \leqslant w$.

Let $\mathcal {U}_q(\mathfrak {g})$ be the quantum group of $\mathfrak {g}$ over $\mathbb {Q}(q^{1/2})$, generated by $e_i,f_i$ $(i \in I)$ and $q^h$ $(h \in \mathsf {P}^\vee )$. We denote by $\mathcal {U}_q^+(\mathfrak {g})$ the subalgebra of $\mathcal {U}_q(\mathfrak {g})$ generated by $e_i$ and $\mathcal {U}_{\mathbb {A}}^+(\mathfrak {g})$ the $\mathbb {A}$-subalgebra of $\mathcal {U}_q(\mathfrak {g})^+$ generated by $e_i^n/[n]_i !$ $(i \in I, \; n \in \mathbb {Z}\mspace {1mu}_{>0})$, where

\begin{align*} q_i \mathbin{:=} q^{\mathsf{d}_i}, \quad [k]_i = \dfrac{q_i^k-q_i^{-k}}{q_i - q_i^{-1}} \quad\text{and}\quad [k]_i! = \prod_{s=1}^k [s]_i. \end{align*}

Set

\begin{align*} \mathcal{A}_q(\mathfrak{n}) = \oplus_{\beta \in \mathsf{Q}^-} \mathcal{A}_q(\mathfrak{n})_\beta \quad \text{where } \mathcal{A}_q(\mathfrak{n})_\beta \mathbin{:=} \operatorname{Hom}_{\mathbb{Q}(q^{1/2})}(\mathcal{U}^+_q(\mathfrak{g})_{-\beta}, \mathbb{Q}(q^{1/2})), \end{align*}

where $\mathcal {U}^+_q(\mathfrak {g})_{-\beta }$ denotes the $(-{\mspace {1mu}\beta })$-root space of $\mathcal {U}^+_q(\mathfrak {g})$. Then $\mathcal {A}_q(\mathfrak {n})$ also has an algebra structure and is called the quantum unipotent coordinate ring of $\mathfrak {g}$. We denote by $\mathcal {A}_{\mathbb {A}}(\mathfrak {n})$ the $\mathbb {A}$-submodule of $\mathcal {A}_q(\mathfrak {n})$ generated by $\psi \in \mathcal {A}_q(\mathfrak {n})$ such that $\psi (\mathcal {U}_{\mathbb {A}}^+(\mathfrak {g})) \subset \mathbb {A}$. Then, $\mathcal {A}_q(\mathfrak {n})$ is an $\mathbb {A}$-subalgebra with a $\mathcal {U}_{\mathbb {A}}^+(\mathfrak {g})$-bimodule structure.

For each $\lambda \in \mathsf {P}^+ \mathbin {:=} \sum _{i \in I} \mathbb {Z}\mspace {1mu}_{\geqslant 0} \varpi _i$ and Weyl group elements $w,w' \in \mathsf {W}$, we can define a specific homogeneous element $D(w\lambda,w'\lambda )$ of $\mathcal {A}_\mathbb {A}(\mathfrak {n})$, called a unipotent quantum minor (see, for example, [Reference Kang, Kashiwara, Kim and OhKKKO18, § 9]).

For $w \in \mathsf {W}$, we denote by $\mathcal {A}_\mathbb {A}(\mathfrak {n}(w))$ the $\mathbb {A}$-submodule of $\mathcal {A}_{\mathbb {A}}(\mathfrak {n})$ consisting of elements $\psi$ such that $e_{i_1} \cdots e_{i_{|{\mspace {1mu}\beta }|}}\psi =0$ for any ${\mspace {1mu}\beta } \in \mathsf {Q}^+ \setminus w \mathsf {Q}^-$ and $(\nu _{i_1},\ldots,\nu _{i_{|{\mspace {1mu}\beta }|}}) \in I^{\mspace {1mu}\beta }$. Then it is an $\mathbb {A}$-subalgebra and we call it the quantum unipotent coordinate ring associated with $w$.

For a reduced sequence ${\boldsymbol {i}}=(i_1,\ldots,i_r)$ of $w \in \mathsf {W}$ and $1 \leqslant k \leqslant r$, define $w^{{\boldsymbol {i}}}_{\leqslant k}=s_{i_1} \cdots s_{i_k}$ and $w^{\boldsymbol {i}}_{< k}=s_{i_1}\cdots s_{i_{k-1}}$. Then $\mathcal {A}_\mathbb {A}(\mathfrak {n}(w))$ is generated by the set of unipotent quantum minors $\{ D(w^{\boldsymbol {i}}_{\leqslant k}\varpi _{i_{k}}, w^{\boldsymbol {i}}_{< k}\varpi _{i_{k}}) \mid 1 \leqslant k \leqslant r \}$ as an $\mathbb {A}$-algebra.

It is proved in [Reference Geiss, Leclerc and SchröerGLS13a, Reference Goodearl and YakimovGY17, Reference Kang, Kashiwara, Kim and OhKKKO18, Reference QinQin20] that $\mathcal {A}_\mathbb {A}(\mathfrak {n}(w))$ has a quantum cluster algebra structure, one of whose quantum seeds $\mathcal {S}^{\boldsymbol {i}}$ can be obtained from a reduced sequence ${\boldsymbol {i}}=(i_1,\ldots,i_r)$ of $w$. To introduce $\mathcal {S}^{\boldsymbol {i}}$, we need preparations.

Let ${\boldsymbol {j}}=(j_1,\ldots,j_l)$ be a sequence of indices in $I$. For $1 \leqslant k \leqslant l$ and $j \in I$, we set

\begin{align*} k^{\boldsymbol{j}}_+&\mathbin{:=} \min(\{u\mid k< u \leqslant l,\; j_u=j_k\} \cup \{l+1\}), \\ k^{\boldsymbol{j}}_-&\mathbin{:=} \max(\{u\mid 1\leqslant u < k,\;j_u=j_k\} \cup \{0\}). \end{align*}

We also set

\[ k^{\boldsymbol{j}}_{\min} \mathbin{:=} \min\{u\mid 1\leqslant u\leqslant k ,\ j_u=j_k\} \quad\text{and}\quad k^{\boldsymbol{j}}_{\max} \mathbin{:=} \max\{u\mid k\leqslant u\leqslant l,\ j_u=j_k\}. \]

We sometimes drop $^{\boldsymbol {j}}$ in the above notation if there is no danger of confusion.

Take $\mathsf {K}=[1,r]$ as an index set and decompose $\mathsf {K}$ into

\[ {\mathsf{K}_\mathrm{fr}} = \{ k \mid 1\leqslant k \leqslant r, \; k^{\boldsymbol{i}}_+ = r+1 \} \quad\text{and}\quad {\mathsf{K}_\mathrm{ex}} \mathbin{:=} \mathsf{K} \setminus {\mathsf{K}_\mathrm{fr}}. \]

We define the $\mathbb {Z}\mspace {1mu}$-valued $\mathsf {K}\times {\mathsf {K}_\mathrm {ex}}$ matrix $\widetilde {B}^{\boldsymbol {i}} =(b^{\boldsymbol {i}}_{st})_{s\in \mathsf {K},t\in {\mathsf {K}_\mathrm {ex}}}$ and the $\mathbb {Z}\mspace {1mu}$-valued skew-symmetric $\mathsf {K}\times \mathsf {K}$ matrix $L^{\boldsymbol {i}}=(l^{\boldsymbol {i}}_{st})_{s,t \in \mathsf {K}}$ as follows:

\begin{align*} b^{\boldsymbol{i}}_{st} &= \begin{cases} \pm 1 & \text{if $s=t^{\boldsymbol{i}}_\pm$,} \\ -a_{i_s,i_t} & \text{if $s< t< s^{\boldsymbol{i}}_+< t^{\boldsymbol{i}}_+$,}\\ a_{i_s,i_t} & \text{if $t< s< t^{\boldsymbol{i}}_+< s^{\boldsymbol{i}}_+$,} \\ 0 & \text{otherwise}, \end{cases} \\ l^{\boldsymbol{i}}_{st} & = (\varpi_{i_s}-w^{\boldsymbol{i}}_{\leqslant s}\varpi_{i_s}, \varpi_{i_t}+w^{\boldsymbol{i}}_{\leqslant t}\varpi_{i_t} )\quad\text{for } s< t. \end{align*}

Then the quantum seed $\mathcal {S}^{\boldsymbol {i}}$ of $\mathcal {A}_\mathbb {A}(\mathfrak {n}(w))$ is given as follows:

(2.2)\begin{equation} \mathcal{S}^{\boldsymbol{i}} \mathbin{:=} \Big( \{ q^{c^{\boldsymbol{i}}_k} D(w^{\boldsymbol{i}}_{\leqslant k}\varpi_{i_k},\varpi_{i_k} ) \}_{k \in \mathsf{K}}, L^{\boldsymbol{i}}, \widetilde{B}^{\boldsymbol{i}} \Big), \end{equation}

where $c^{\boldsymbol {i}}_s = \frac {1}{4} (\varpi _{i_s}-w^{\boldsymbol {i}}_{\leqslant s}\varpi _{i_s} ,\varpi _{i_s}-w^{\boldsymbol {i}}_{\leqslant s}\varpi _{i_s} ) \in \mathbb {Z}\mspace {1mu}/2$. Note that $(L^{\boldsymbol {i}}\widetilde {B}^{\boldsymbol {i}})_{ab} = -2 \mathsf {d}_{i_a} \times \delta _{a,b}$ for $(a,b)\in \mathsf {K}\times {\mathsf {K}_\mathrm {ex}}$, ${\rm wt}(D(w^{\boldsymbol {i}}_{\leqslant k}\varpi _{i_k},\varpi _{i_k} )) = - \varpi _{i_s} + w^{\boldsymbol {i}}_{\leqslant s}\varpi _{i_s}$, and

\[ \big\{ q^{c^{\boldsymbol{i}}_k} D(w^{\boldsymbol{i}}_{\leqslant k}\varpi_{i_k},\varpi_{i_k}) = q^{c^{\boldsymbol{i}}_k} D(w\varpi_{i_k},\varpi_{i_k} ) \mid k \in {\mathsf{K}_\mathrm{fr}}\!\big\} \]

forms the set of frozen variables of the quantum cluster algebra $\mathcal {A}_\mathbb {A}(\mathfrak {n}(w))$. We call $\mathcal {S}^{\boldsymbol {i}}$ the GLS seed (associated with ${\boldsymbol {i}}$).

We set $\mathcal {D}(w) \mathbin {:=} \{ q^m D(w\varpi,\varpi ) \mid m \in \mathbb {Z}\mspace {1mu}/2, \; \varpi \in \mathsf {P}^+ \}$. Then it is well-known that $\mathcal {D}(w)$ consists of $q$-central elements of $\mathcal {A}_\mathbb {A}(\mathfrak {n}(w))$ and, hence, forms an Ore set. We denote by $\mathcal {A}_\mathbb {A}(\mathfrak {n}^w)$ the quotient ring of $\mathcal {A}_\mathbb {A}(\mathfrak {n}(w))$ by the Ore set $\mathcal {D}(w)$. Then $\mathcal {A}_\mathbb {A}(\mathfrak {n}^w)$ has also the quantum cluster algebra structure with the invertible frozen variables $\{ q^{c^{\boldsymbol {i}}_k} D(w^{\boldsymbol {i}}_{\leqslant k}\varpi _{i_k},\varpi _{i_k}) \}_{k \in {\mathsf {K}_\mathrm {fr}}}$ in the sense of [Reference Berenstein and ZelevinskyBZ05].

2.3 Quiver Hecke algebras and categorifications

Let $\mathbf {k}$ be a base field. For $i,j \in I$, we choose polynomials $\mathcal {Q}_{i,j}(u,v) \in \mathbf {k}[u,v]$ such that (a) $\mathcal {Q}_{i,j}(u,v) =\mathcal {Q}_{j,i}(v,u)$ and (b) each $\mathcal {Q}_{i,j}(u,v)$ is of the following form:

\[ \mathcal{Q}_{i,j}(u,v) = \delta(i\ne j) \hspace{-6ex} \sum_{p({\mspace{1mu}\alpha}_i,{\mspace{1mu}\alpha}_i)+q({\mspace{1mu}\alpha}_j,{\mspace{1mu}\alpha}_j)=-2({\mspace{1mu}\alpha}_i,{\mspace{1mu}\alpha}_j)} \hspace{-6ex} t_{i,j;p,q} u^p v^q \quad \text{where} \ t_{i,j;-a_{i,j},0} \in \mathbf{k}^\times. \]

For a Cartan datum $(\mathsf {A},\mathsf {P},\Pi,\mathsf {P}^\vee,\Pi ^\vee )$ and ${\mspace {1mu}\beta } \in \mathsf {Q}^+$, the quiver Hecke algebra $R({\mspace {1mu}\beta })$ associated with $(\mathcal {Q}_{i,j})_{i,j \in I}$ is the $\mathbb {Z}\mspace {1mu}$-graded algebra over $\mathbf {k}$ generated by the elements

\[ \{ e(\nu) \}_{\nu \in I^{\mspace{1mu}\beta}} , \quad \{ x_k \}_{1 \leqslant k \leqslant |{\mspace{1mu}\beta}|}, \quad\{ \tau_m \}_{1 \leqslant m <|{\mspace{1mu}\beta}|} \]

subject to certain defining relations (see [Reference Kashiwara, Kim, Oh and ParkKKOP21, Definition 1.1] for instance). Note that the $\mathbb {Z}\mspace {1mu}$-grading of $R({\mspace {1mu}\beta })$ is determined by the degrees of following elements:

\[ \deg(e(\nu))=0, \quad \deg(x_ke(\nu)) = ({\mspace{1mu}\alpha}_{\nu_k},{\mspace{1mu}\alpha}_{\nu_k}), \quad\text{and}\quad \deg(\tau_me(\nu)) = -({\mspace{1mu}\alpha}_{\nu_m},{\mspace{1mu}\alpha}_{\nu_{m+1}}). \]

We say that $R({\mspace {1mu}\beta })$ is symmetric if $\mathcal {Q}_{i,j}(u,v) \in \mathbf {k}[u-v]$ for $i,j \in {\rm supp}({\mspace {1mu}\beta })$.

We denote by $R({\mspace {1mu}\beta })\text {-}\mathrm {gmod}$ the category of finite-dimensional graded $R({\mspace {1mu}\beta })$-modules with homomorphisms of degree $0$. For $M \in R({\mspace {1mu}\beta })\text {-}\mathrm {gmod}$, we set ${\rm wt}(M) \mathbin {:=} -{\mspace {1mu}\beta } \in \mathsf {Q}^-$. Note that there exists the degree shift functor, denoted by $q$, such that $(qM)_n = M_{n-1}$ for $M=\oplus_{k \in \mathbb {Z}\mspace {1mu}} M_k \in R({\mspace {1mu}\beta })\text {-}\mathrm {gmod}$.

Throughout this paper, we usually deal with graded $R({\mspace {1mu}\beta })$-modules $({\mspace {1mu}\beta } \in \mathsf {Q}^+)$ and sometimes skip grading shifts. Thus, we usually say that $M$ is an $R$-module instead of saying that $M$ is a graded $R({\mspace {1mu}\beta })$-module and $f\colon M \to N$ is a homomorphism if $f\colon q^aM \to N$ is a morphism in $R({\mspace {1mu}\beta })\text {-}\mathrm {gmod}$. We set

\[ \mathrm{H{\scriptstyle OM}}_{R({\mspace{1mu}\beta})}(M,N) \mathbin{:=} \oplus_{a \in \mathbb{Z}\mspace{1mu}} \mathrm{H{\scriptstyle OM}}_{R({\mspace{1mu}\beta})}(M,N)_a \]

with $\mathrm {H{\scriptstyle OM}}_{R({\mspace {1mu}\beta })}(M,N)_a\mathbin {:=}\operatorname {Hom}_{R({\mspace {1mu}\beta })\text {-}\mathrm {gmod}}(q^aM,N)_a$. We write $\deg (f) \mathbin {:=} a$ for an $f \in \mathrm {H{\scriptstyle OM}}_{R({\mspace {1mu}\beta })} (M,N)_a$.

For an $R({\mspace {1mu}\beta })$-module $M$ and an $R(\gamma )$-module $N$, we can obtain $R({\mspace {1mu}\beta }+\gamma )$-module

\[ M \circ N \mathbin{:=} R({\mspace{1mu}\beta}+\gamma) e({\mspace{1mu}\beta},\gamma) \mathop\otimes_{R({\mspace{1mu}\beta}) \otimes R(\gamma)} (M \otimes N), \]

where $e({\mspace {1mu}\beta },\gamma ) \mathbin {:=} \sum _{\nu \in I^{\mspace {1mu}\beta }, \nu ' \in I^\gamma } e(\nu * \nu ') \in R({\mspace {1mu}\beta }+\gamma )$. Here $\nu *\nu '$ denotes the concatenation of $\nu$ and $\nu '$, and $\circ$ is called the convolution product. We say that two simple $R$-modules $M$ and $N$ strongly commute if $M \circ N$ is simple. If a simple module $M$ strongly commutes with itself, then $M$ is called real. A simple $R$-module $M$ is said to be prime if there are no non-trivial simple $R$-modules $N_1$ and $N_2$ such that $M \simeq N_1 \circ N_2$.

For an $R({\mspace {1mu}\beta })$-module $M$, the dual space $M^* \mathbin {:=} \operatorname {Hom}_\mathbf {k}(M,\mathbf {k})$ admits an $R({\mspace {1mu}\beta })$-module structure via

\[ (r \cdot f)(u)= f(\psi(r)u) \quad (r \in R({\mspace{1mu}\beta}), \ u \in M, \ f \in M^*). \]

Here $\psi$ denotes the $\mathbf {k}$-algebra anti-involution $R({\mspace {1mu}\beta })$ which fixes the generators of $R({\mspace {1mu}\beta })$. A simple $R({\mspace {1mu}\beta })$-module $M$ is called self-dual if $M^* \simeq M$.

Set $R\text {-}\mathrm {gmod} \mathbin {:=} \oplus_{{\mspace {1mu}\beta } \in \mathsf {Q}^+}R({\mspace {1mu}\beta })\text {-}\mathrm {gmod}$. Then the category $R\text {-}\mathrm {gmod}$ is a monoidal category with the tensor product $\circ$ and the unit object $\mathbf {1} \mathbin {:=} \mathbf {k} \in R(0)\text {-}\mathrm {gmod}$. Hence, the Grothendieck group $K(R\text {-}\mathrm {gmod} )$ has the $\mathbb {Z}\mspace {1mu}[q^{\pm 1}]$-algebra structure derived from $\circ$ and the degree shift functors $q^{\pm 1}$.

For a monoidal abelian subcategory $\mathcal {C}$ of $R\text {-}\mathrm {gmod}$ stable by grading shifts, we set

\[ \mathcal{K}_\mathbb{A}(\mathcal{C}) \mathbin{:=} \mathbb{A} \mathop\otimes_{\mathbb{Z}\mspace{1mu}[q^{\pm1}]} K(\mathcal{C} ), \]

where $K(\mathcal {C} )$ denotes the Grothendieck ring of $\mathcal {C}$. For a subcategory $\mathcal {C}$ of $R\text {-}\mathrm {gmod}$, we denote by ${\rm Irr}(\mathcal {C})$ the set of the isomorphism classes of self-dual modules in $\mathcal {C}$. Note that ${\rm Irr}(R\text {-}\mathrm {gmod})$ forms an $\mathbb {A}$-basis of $\mathcal {K}_\mathbb {A}(R\text {-}\mathrm {gmod})$.

It is proved in [Reference Khovanov and LaudaKL09, Reference Khovanov and LaudaKL11, Reference RouquierRou08] that there exists an $\mathbb {A}$-algebra isomorphism

(2.3)\begin{equation} \Omega \colon \mathcal{K}_{\mathbb{A}}(R\text{-}\mathrm{gmod}) {\mathop\longrightarrow\limits^{\sim}} \mathcal{A}_{\mathbb{A}}(\mathfrak{n}). \end{equation}

Proposition 2.2 [Reference Kashiwara, Kim, Oh and ParkKKOP18, Proposition 4.1]

For $\varpi \in \mathsf {P}^+$ and $\mu,\zeta \in \mathsf {W} \varpi$ with $\mu \preccurlyeq \zeta$, there exists a self-dual real simple $R( \zeta -\mu )$-module $M(\mu,\zeta )$ such that

\[ \Omega ([M(\mu,\zeta)]) = D(\mu,\zeta). \]

Here, $[M(\mu,\zeta )]$ denotes the isomorphism class of $M(\mu,\zeta )$ which is called the determinantal module associated with $D(\mu,\zeta )$.

For an $R({\mspace {1mu}\beta })$-module $M$, we define

\begin{align*} W(M) & \mathbin{:=} \{ \gamma \in \mathsf{Q}^+ \cap ({\mspace{1mu}\beta}-\mathsf{Q}^+) \mid e(\gamma,{\mspace{1mu}\beta}-\gamma)M \ne 0 \}, \\ W^*(M) & \mathbin{:=} \{ \gamma \in \mathsf{Q}^+ \cap ({\mspace{1mu}\beta}-\mathsf{Q}^+) \mid e({\mspace{1mu}\beta}-\gamma,\gamma)M \ne 0 \}. \end{align*}

An ordered pair $(M, N)$ of $R$-modules is called unmixed [Reference Tingley and WebsterTW16, Definition 2.5] if

\[ W^*(M) \cap W(N) \subset \{0\}. \]

For $w \in \mathsf {W}$, we denote by $\mathscr {C}_w$ the full subcategory of $R\text {-}\mathrm {gmod}$ whose objects $M$ satisfy $W(M) \subset \mathsf {Q}^+ \cap w\mathsf {Q}^-$. Then the category $\mathscr {C}_w$ is the smallest monoidal abelian category of $R\text {-}\mathrm {gmod}$ which (i) is stable under taking subquotients, extensions, grading shifts and (ii) contains $\big \{ S^{\boldsymbol {i}}_k \mathbin {:=} M(w^{\boldsymbol {i}}_{\leqslant k}\varpi _{i_k},w^{\boldsymbol {i}}_{< k}\varpi _{i_k}) \mid 1 \leqslant k \leqslant r \big \}$ for any reduced sequence ${\boldsymbol {i}}$ of $w$. We call $S^{\boldsymbol {i}}_k$ the $k$th cuspidal module associated with ${\boldsymbol {i}}$. Defining ${\mspace {1mu}\beta }_k^{\boldsymbol {i}} \mathbin {:=} w^{\boldsymbol {i}}_{< k} {\mspace {1mu}\alpha }_{i_k}$ for $1 \leqslant k \leqslant r$, one can see that $\{{\mspace {1mu}\beta }_k^{\boldsymbol {i}} \mid 1\leqslant k\leqslant r\}=\Delta ^+ \cap w \Delta ^-$, and $-{\rm wt}(S^{\boldsymbol {i}}_k) = {\mspace {1mu}\beta }^{\boldsymbol {i}}_k$. Then we have [Reference KimuraKim12, § 4]

\[ \Omega( \mathcal{K}_\mathbb{A}(\mathscr{C}_w ) ) \simeq \mathcal{A}_\mathbb{A}(\mathfrak{n}(w)). \]

2.4 $R$-matrices and affreal simple modules

For ${\mspace {1mu}\beta } \in \mathsf {Q}^+$ and $i \in I$, let

\[ \mathfrak{p}_{i,{\mspace{1mu}\beta}} = \sum_{ \eta\, \in I^{\mspace{1mu}\beta}} \bigg( \prod_{a \in [1,|{\mspace{1mu}\beta}|]; \ \eta_a = i} x_a \bigg) e(\eta) \in \mathcal{Z}(R({\mspace{1mu}\beta})), \]

where $\mathcal {Z}(R({\mspace {1mu}\beta }))$ denotes the center of $R({\mspace {1mu}\beta })$.

Definition 2.3 [Reference Kashiwara and ParkKP18, Definition 2.2]

For an $R({\mspace {1mu}\beta })$-module $M$, we say that $M$ admits an affinization if there exists an $R({\mspace {1mu}\beta })$-module $\widehat {M}$ satisfying the condition: there exists an endomorphism $z_{\widehat {M}}$ of degree $t \in \mathbb {Z}\mspace {1mu}_{>0}$ such that $\widehat {M}/z_{\widehat {M}}\widehat {M} \simeq M$ and:

  1. (i) $\widehat {M}$ is a finitely generated free module over the polynomial ring $\mathbf {k}[z_{\widehat {M}}]$;

  2. (ii) $\mathfrak {p}_{i,{\mspace {1mu}\beta }} \widehat {M} \ne 0$ for all $i \in I$.

We say that a simple $R({\mspace {1mu}\beta })$-module $M$ is affreal if $M$ is real and admits an affinization.

It is known that any $M \in R({\mspace {1mu}\beta })\text {-}\mathrm {gmod}$ admits an affinization if $R({\mspace {1mu}\beta })$ is symmetric. However, when $R({\mspace {1mu}\beta })$ is not symmetric, it is widely open whether an $R({\mspace {1mu}\beta })$-module $M$ admits an affinization or not.

Theorem 2.4 [Reference Kashiwara, Kim, Oh and ParkKKOP21, Theorem 3.26]

For $\varpi \in \mathsf {P}^+$ and $\mu,\zeta \in \mathsf {W}\varpi$ such that $\mu \preccurlyeq \zeta$, the determinantal module $M(\mu,\zeta )$ is affreal.

Proposition 2.5 [Reference Kang, Kashiwara, Kim and OhKKKO18, Reference Kashiwara, Kim, Oh and ParkKKOP21]

Let $M$ and $N$ be simple modules such that one of them is affreal. Then there exists a unique $R$-module homomorphism $\mathbf {r}_{M,N} \in \mathrm {H{\scriptstyle OM}}_R(M,N)$ satisfying

\[ \mathrm{H{\scriptstyle OM}}_R(M \circ N,N \circ M)=\mathbf{k}\, \mathbf{r}_{M,N}. \]

We call the homomorphism $\mathbf {r}_{M,N}$ the $R$-matrix.

Definition 2.6 For simple $R$-modules $M$ and $N$ such that one of them is affreal, we define

\begin{align*} \Lambda(M,N) &\mathbin{:=} \deg (\mathbf{r}_{M,N}) , \\ \widetilde{\Lambda}(M,N) &\mathbin{:=} \tfrac{1}{2} \big( \Lambda(M,N) + \big({\rm wt}(M), {\rm wt}(N)\big) \big) , \\ \mathfrak{d}(M,N) &\mathbin{:=} \tfrac{1}{2} \big( \Lambda(M,N) + \Lambda(N,M)\big). \end{align*}

It is proved in [Reference Kang, Kashiwara, Kim and OhKKKO18, Reference Kashiwara, Kim, Oh and ParkKKOP21] that the invariants $\widetilde {\Lambda }(M,N)$ and $\mathfrak {d}(M,N)$ in Definition 2.6 belong to $\mathbb {Z}\mspace {1mu}_{\geqslant 0}$.

For simple modules $M$ and $N$, $M \triangledown N$ and $M \triangle N$ denote the head and the socle of $M \circ N$, respectively.

Proposition 2.7 [Reference Kang, Kashiwara, Kim and OhKKKO15, Lemma 3.1.4] and [Reference Kang, Kashiwara, Kim and OhKKKO18, Corollary 4.1.2]

Let $M$ and $N$ be simple $R$-modules such that one of them is affreal.

  1. (i) The image of ${\mathbf r} {M,N}$ is equal to $M\triangledown N$ and $N\triangle M$.

  2. (ii) The head $M \triangledown N$ and socle $M \triangle N$ are simple modules and each of them appears exactly once in the composition series of $M \circ N$ (up to a grading shift).

  3. (iii) Assume that $N$ is affreal.

    1. (a) If a simple subquotient $S$ of $M \circ N$ is not isomorphic to $M\triangledown N$, then $\Lambda (S,N)<\Lambda (M\triangledown N,N)=\Lambda (M,N)$.

    2. (b) If a simple subquotient $S$ of $M \circ N$ is not isomorphic to $M\triangle N$, then $\Lambda (N,S)<\Lambda (N,M\triangle N)=\Lambda (N,M)$.

  4. (iv) If $M$ and $N$ are self-dual, then $q^{\widetilde {\Lambda }(M,N)} M \triangledown N$ is a self-dual simple module.

  5. (v) The following conditions are equivalent:

    1. (a) $M \circ N \simeq N \circ M$ up to a grading shift;

    2. (b) $M \circ N$ is simple;

    3. (c) $\mathfrak {d}(M,N)=0$;

    4. (d) $M \triangledown N \simeq M \triangle N$ up to a grading shift.

Proposition 2.8 [Reference Kashiwara, Kim, Oh and ParkKKOP21, Corollary 3.18]

Let $M$ be an affreal simple module. Let $X$ be an $R$-module in $R\text {-}\mathrm {gmod}$. Let $n \in \mathbb {Z}\mspace {1mu}_{>0}$ and assume that any simple subquotient $S$ of $X$ satisfies $\mathfrak {d}(M, S) \leqslant n$. Then any simple subquotient $N$ of $M \circ X$ satisfies $\mathfrak {d}(M, N) < n$. In particular, any simple subquotient of $M^{\circ n} \circ X$ strongly commutes with $M$.

An ordered sequence of simple modules $\underline {L}=(L_1,\ldots,L_r)$ is called almost affreal if all $L_i$ $(1 \leqslant i \leqslant r)$ are affreal except for at most one.

Definition 2.9 An almost affreal sequence $\underline {L}$ of simple modules is called a normal sequence if the composition of $R$-matrices

\begin{align*} {\mathbf r}{\underline{L}} & \mathbin{:=} \prod_{1 \leqslant i < k \leqslant r} {\mathbf r}{L_i,L_k } = ({\mathbf r}{L_{r-1},L_r}) \circ \cdots ({\mathbf r}{L_2,L_r}\circ \cdots \circ {\mathbf r}{L_2,L_3}) \circ ({\mathbf r}{L_1,L_r} \circ \cdots \circ {\mathbf r}{L_1,L_2}) \\ & \colon \; q^{\sum_{1 \leqslant i < k \leqslant r} \Lambda(L_i,L_k)} L_1 \circ \cdots \circ L_r \rightarrow L_r \circ \cdots \circ L_1 \quad \text{does not vanish.} \end{align*}

Lemma 2.10 [Reference Kashiwara and KimKK19, § 2.3] and [Reference Kashiwara, Kim, Oh and ParkKKOP23, § 2.2]

Let $\underline {L}$ be an almost affreal sequence of simple modules. If $\underline {L}$ is normal, then the image of ${\mathbf r}{\underline {L}}$ is simple and coincides with the head of $L_1 \circ \cdots \circ L_r$ and also with the socle of $L_r \circ \cdots \circ L_1$, up to grading shifts.

Lemma 2.11 [Reference Kang, Kashiwara, Kim and OhKKKO18, Proposition 3.2.13]

Let $(A,B,C)$ be an almost affreal sequence. Then we have the following:

  1. (i) $\Lambda (A,B\triangledown C) =\Lambda (A,B)+\Lambda (A,C)$ if $A$ and $B$ commute;

  2. (ii) $\Lambda (A\triangledown B, C) =\Lambda (A,C)+\Lambda (B,C)$ if $B$ and $C$ commute.

For a given almost affreal sequence $\underline {L}$ of $R$-modules, the sufficient conditions for $\underline {L}$ being normal are studied in [Reference Kashiwara and KimKK19, Reference Kashiwara, Kim, Oh and ParkKKOP23]. In this paper, we will use the conditions frequently.

2.5 Commuting families

Let $J$ be an index set. We say that a family of affreal simple modules $\mathcal {M}= \{ M_j \}_{j \in J}$ in $R\text {-}\mathrm {gmod}$ is a commuting family if

\[ M_i \circ M_j \simeq M_j \circ M_i \quad \text{up to a grading shift for any $i,j \in J$.} \]

For a commuting family $\mathcal {M}= \{ M_j \ | \ j \in J\}$ in $R\text {-}\mathrm {gmod}$, let us take ${\mathbf \lambda} \colon \mathbb {Z}\mspace {1mu}^{\oplus J} \times \mathbb {Z}\mspace {1mu}^{\oplus J} \to \mathbb {Z}\mspace {1mu}$ such that

(2.4)\begin{equation} {\mathbf \lambda}(\mathbf{e}_i,\mathbf{e}_j)-{\mathbf \lambda}(\mathbf{e}_j,\mathbf{e}_i) = \Lambda(M_i,M_j) \quad \text{for any $ i,j \in J$.}\end{equation}

Here $\{ \mathbf {e}_j \mid j \in J\}$ is the standard basis of $\mathbb {Z}\mspace {1mu}^{\oplus J}$. Then there exists a family $\{ \mathcal {M}_{\mathbf \lambda} (\mathbf {a}) \mid \mathbf {a} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^{\oplus J}\}$ of simple modules in $\mathscr {C}_w$ such that

(2.5)\begin{equation} \begin{array}{ll} \mathcal{M}_{\mathbf \lambda}(0) = \mathbf{1}, \quad \mathcal{M}_{\mathbf \lambda}(\mathbf{e}_j) =M_j & \text{for any }j\in J, \\ \mathcal{M}_{\mathbf \lambda}(\mathbf{a}) \circ \mathcal{M}_{\mathbf \lambda}(\mathbf{b}) \simeq q^{ -{\mathbf \lambda}(a,b) } \mathcal{M}_{\mathbf \lambda}(\mathbf{a}+\mathbf{b}) & \text{for any }\mathbf{a},\mathbf{b} \in \mathbb{Z}\mspace{1mu}_{\geqslant 0}^{J}. \end{array} \end{equation}

We sometimes omit $_{\mathbf \lambda}$ for notational simplicity.

Remark 2.12 Note that ${\mathbf \lambda} (\mathbf {e}_i,\mathbf {e}_j)=\widetilde {\Lambda }(M_i,M_j)$ satisfies condition (2.4). Moreover, if all the $M_i$ are self-dual, then $\mathcal {M}_{\mathbf \lambda} (\mathbf {a})$ is self-dual for any $\mathbf {a}\in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^{J}$. We usually take this choice of ${\mathbf \lambda}$.

Definition 2.13 A commuting family $\mathcal {M}=\{ M_j \mid j \in J\}$ is called independent if $\mathbf {a},\mathbf {b} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^{\oplus J}$ satisfies $\mathcal {M}(\mathbf {a}) \simeq q^s \mathcal {M}(\mathbf {b})$ for some $s \in \mathbb {Z}\mspace {1mu}$, then we have $\mathbf {a}=\mathbf {b}$.

The following lemma is obvious.

Lemma 2.14 Let $\mathcal {M} = \{ M_j \mid j \in J \}$ be a commuting family. Then it is independent if and only if the set $\big \{ \big [ \mathcal {M}(\mathbf {a})\big ] \mid \mathbf {a} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^{\oplus J} \big \}$ in $K(R\text {-}\mathrm {gmod})$ is linearly independent over $\mathbb {Z}\mspace {1mu}[q^{\pm 1}]$.

2.6 Localization of $\mathscr {C}_w$

Throughout this subsection, we fix $w \in \mathsf {W}$ and set

\[ I_w\mathbin{:=}\{i\in I\mid w\varpi_i\not=\varpi_i\}. \]

For notational simplicity, let us write

\[ C_i \mathbin{:=} M(w \varpi_i,\varpi_i) \in R\text{-}\mathrm{gmod} \quad \text{for $i \in I$.} \]

Then $\{ \Omega ([C_i]) \mid i \in I_w\}$ forms the set of frozen variables of $\mathcal {A}_\mathbb {A}(\mathfrak {n}(w))$. For each $\mu = \sum _{i \in I} \mu _i\varpi _i \in \mathsf {P}^+$, we set $C_\mu = M(w\mu,\mu )$, which is a self-dual convolution product $q^c \circ _{i \in I} C_i^{\circ \mu _i}$ for some $c\in \mathbb {Z}\mspace {1mu}$.

It is proved in [Reference Kashiwara, Kim, Oh and ParkKKOP21, Reference Kashiwara, Kim, Oh and ParkKKOP23, Reference Kashiwara, Kim, Oh and ParkKKOP24a] that there exists a monoidal abelian category $\widetilde {\mathscr {C}}_w = \mathscr {C}_w[C_i^{\circ -1} \mid i \in I]$ with a tensor product $\circ$, a degree shift functor $q$ and a monoidal exact fully faithful functor $\Phi _w\colon \mathscr {C}_w \to \widetilde {\mathscr {C}}_w$ satisfying the following properties.

(2.6)

(For the precise properties, see [Reference Kashiwara, Kim, Oh and ParkKKOP21, Reference Kashiwara, Kim, Oh and ParkKKOP23, Reference Kashiwara, Kim, Oh and ParkKKOP24a].)

Theorem 2.15 ([Reference Kashiwara, Kim, Oh and ParkKKOP21] (see also [Reference Kashiwara, Kim, Oh and ParkKKOP23, Remark 3.6]))

There exists an $\mathbb {A}$-algebra isomorphism

\[ \widetilde{\Omega} \colon \mathcal{K}_\mathbb{A}(\widetilde{\mathscr{C}}_w) \mathbin{:=} \mathbb{A} \mathop\otimes_{\mathbb{Z}\mspace{1mu}[q^{\pm1}]} K(\widetilde{\mathscr{C}}_w) {\mathop\longrightarrow\limits^{\sim}} \mathcal{A}_\mathbb{A}(\mathfrak{n}^w) \quad \text{such that} \quad \widetilde{\Omega}|_{\mathcal{K}_\mathbb{A}(\mathscr{C}_w)} = \Omega. \]

Here $K(\widetilde {\mathscr {C}}_w)$ denotes the Grothendieck ring of $\widetilde {\mathscr {C}}_w$.

A pair $(\varepsilon \colon X \otimes Y \to \mathbf {1}, \; \eta \colon \mathbf {1} \to Y \otimes X)$ of morphisms in a monoidal category with a unit object $\mathbf {1}$ is called an adjunction if the following two conditions hold.

  1. (a) The composition $X \simeq X \otimes \mathbf {1} \mathop\longrightarrow\limits^{[X \otimes \eta]} X \otimes Y \otimes X \mathop\longrightarrow\limits^{[\varepsilon \otimes X]} \mathbf {1} \otimes X \simeq X$ is the identity.

  2. (b) The composition $Y \simeq \mathbf {1} \otimes Y \mathop\longrightarrow\limits^{[\eta \otimes Y]} Y \otimes X \otimes Y \mathop\longrightarrow\limits^{[Y \otimes \varepsilon]} Y \otimes \mathbf {1} \simeq Y$ is the identity.

In the case when $(\varepsilon,\eta )$ is an adjunction, we say that $X$ is a left dual to $Y$, $Y$ is a right dual to $X$ and $(X,Y)$ is a dual pair.

Theorem 2.16 [Reference Kashiwara, Kim, Oh and ParkKKOP21, Reference Kashiwara, Kim, Oh and ParkKKOP23]

The monoidal category $\widetilde {\mathscr {C}}_w$ is rigid; i.e. every object of $\widetilde {\mathscr {C}}_w$ has a right dual and a left dual in $\widetilde {\mathscr {C}}_w$.

2.7 Determinantal modules and monoidal clusters

In this subsection, we denote by $\mathcal {C}$ the category of $\mathscr {C}_w$ or $\widetilde {\mathscr {C}}_w$. Recall that

\begin{align*} \mathscr{A}&=\mathcal{K}_\mathbb{A}(\mathcal{C})\ \text{has a quantum cluster algebra structure via an isomorphism}\\ \Omega &=\Omega\ \text{or}\ \widetilde{\Omega}. \end{align*}

Let ${\boldsymbol {i}}=(i_1,\ldots,i_r)$ be a reduced sequence of $w \in \mathsf {W}$. For $k$ such that $1\leqslant k\leqslant r$, set

\[ M_k^{\boldsymbol{i}}=M(w^{\boldsymbol{i}}_{\leqslant k}\varpi_{i_k},\varpi_{i_k}) \]

(see Proposition 2.2 for the notation).

Proposition 2.17 [Reference Kashiwara, Kim, Oh and ParkKKOP18, Theorem 4.12]

Let ${\boldsymbol {i}}=(i_1,\ldots,i_r)$ be a reduced sequence of $w \in \mathsf {W}$. For $s < t \in \mathsf {K}$, we have

\[ -\Lambda(M_s^{\boldsymbol{i}},M_t^{\boldsymbol{i}}) = (\varpi_{i_s}-w^{\boldsymbol{i}}_{\leqslant s}\varpi_{i_s}, \varpi_{i_t}+w^{\boldsymbol{i}}_{\leqslant t}\varpi_{i_t} ) = l^{\boldsymbol{i}}_{st} = (L^{\boldsymbol{i}})_{st}. \]

We say that a commuting family $\mathcal {M}=\{M_i\}_{i\in \mathsf {K}}$ in $\mathcal {C}$ is a monoidal cluster if there exists a quantum seed $(\{ X_i \}_{i \in \mathsf {K}}, L=(l_{i,j})_{i,j \in \mathsf {K}}, \widetilde {B}=(b_{i,j})_{i \in \mathsf {K},j\in {\mathsf {K}_\mathrm {ex}}} )$ of $\mathscr {A}$ such that

\[ X_i = \Omega( q^{{1/4}({\rm wt}(M_i),{\rm wt}(M_i))}[M_i]) \quad \text{and} \quad l_{i,j} = -\Lambda(M_i,M_j). \]

Note that every monoidal seed is independent since the quantum cluster monomials in a cluster are linearly independent over $\mathbb {A}$ by the definition.

With Proposition 2.2 and (2.2), Proposition 2.17 says that

(2.7)\begin{equation} \mathcal{M}^{\boldsymbol{i}} \mathbin{:=} \{ M^{\boldsymbol{i}}_k \}_{1 \leqslant k \leqslant r} \text{ is a monoidal cluster in $\mathscr{C}_w$,} \end{equation}

for any reduced sequence ${\boldsymbol {i}}=(i_1,\ldots,i_r)$ of $w$. We call $\mathcal {M}^{\boldsymbol {i}}$ the GLS cluster (associated with ${\boldsymbol {i}}$).

3. Quasi-Laurent family and Laurent family

In this section, we introduce the notions of quasi-Laurent families and Laurent families, which allow us to associate two vectors in $\mathbb {Z}\mspace {1mu}^{J}$ with each simple module.

3.1 Definition

Let $J$ be a finite index set. Let $\mathscr {C}$ be a full monoidal subcategory of $R\text {-}\mathrm {gmod}$ stable by taking subquotients, extensions and grading shifts.

Lemma 3.1 Let $\mathcal {M}=\{ M_j \mid j \in J\}$ be an independent commuting family of affreal simple objects in $\mathscr {C}$ and $X$ a simple module in $\mathscr {C}$.

  1. (i) If $\mathbf {a}$, $\mathbf {a}'$, $\mathbf {b}$, $\mathbf {b}'\in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^{J}$ satisfy $X \triangledown \mathcal {M}(\mathbf {a}) \simeq \mathcal {M}(\mathbf {b})$ and $X \triangledown \mathcal {M}(\mathbf {a}') \simeq \mathcal {M}(\mathbf {b}')$ up to grading shifts, then one has $\mathbf {b}-\mathbf {a}=\mathbf {b}'-\mathbf {a}'$.

  2. (ii) If $\mathbf {a}$, $\mathbf {a}'$, $\mathbf {b}$, $\mathbf {b}'\in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^{J}$ satisfy $\mathcal {M}(\mathbf {a}) \triangledown X \simeq \mathcal {M}(\mathbf {b})$ and $\mathcal {M}(\mathbf {a}') \triangledown X \simeq \mathcal {M}(\mathbf {b}')$ up to grading shifts, then one has $\mathbf {b}-\mathbf {a}=\mathbf {b}'-\mathbf {a}'$.

Proof. Since the proof are similar, we prove only part (i). We have

\begin{align*} \mathcal{M}(\mathbf{b}+\mathbf{a}') \simeq \mathcal{M}(\mathbf{b}) \triangledown \mathcal{M}(\mathbf{a}') &\simeq( X \triangledown \mathcal{M}(\mathbf{a}) ) \triangledown \mathcal{M}(\mathbf{a}') \simeq {\operatorname{hd}}(X \circ \mathcal{M}(\mathbf{a}') \circ \mathcal{M}(\mathbf{a}) )\\ &\simeq( X \triangledown \mathcal{M}(\mathbf{a}') ) \triangledown \mathcal{M}(\mathbf{a}) \simeq \mathcal{M}(\mathbf{b}'+\mathbf{a}), \end{align*}

and, hence, we have $\mathbf {a}'+\mathbf {b}=\mathbf {a}+\mathbf {b}'$ since $\mathcal {M}$ is independent.

Definition 3.2 We say that a commuting family $\mathcal {M}= \{ M_j \mid j \in J \}$ of affreal simple objects of $\mathscr {C}$ is a quasi-Laurent family in $\mathscr {C}$ if $\mathcal {M}$ satisfies the following conditions:

  1. (a) $\mathcal {M}$ is independent; and

  2. (b) if a simple module $X$ commutes with all $M_j$ $(j \in J)$, then there exist $\mathbf {a},\mathbf {b} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^{J}$ such that

    \[ X \circ \mathcal{M}(\mathbf{a}) \simeq \mathcal{M}(\mathbf{b})\quad\text{up to a grading shift}. \]

If $\mathcal {M}$ satisfies part (a) and part (c) below, then we say that $\mathcal {M}$ is a Laurent family:

  1. (c) if a simple module $X$ commutes with all $M_j$ $(j \in J)$, then there exists $\mathbf {b}\in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^{J}$ such that $X\simeq \mathcal {M}(\mathbf {b})$.

Note that a Laurent family is a quasi-Laurent family.

Lemma 3.3 Let $\mathcal {M} = \{ M_j \mid j \in J\}$ be a quasi-Laurent family in $\mathscr {C}$. Then we have the following:

  1. (i) for any simple module $X \in \mathscr {C}$, there exist $\mathbf {a},\mathbf {b} \in \mathbb {Z}\mspace {1mu}^J_{\geqslant 0}$ such that $X \triangledown \mathcal {M}(\mathbf {a}) \simeq \mathcal {M}(\mathbf {b})$ up to a grading shift;

  2. (ii) for any simple module $X \in \mathscr {C}$, there exist $\mathbf {a}',\mathbf {b}' \in \mathbb {Z}\mspace {1mu}^J_{\geqslant 0}$ such that $X \triangle \mathcal {M}(\mathbf {a}') \simeq \mathcal {M}(\mathbf {b}')$ up to a grading shift.

Proof. Since the proofs are similar, we shall only prove the first statement. Let us take $\mathbf {a}^{(1)} \in \mathbb {Z}\mspace {1mu}^J_{\geqslant 0}$ such that $\mathbf {a}^{(1)}_j \gg 0$ for all $j \in J$. Then Proposition 2.8 says that the simple module $Y \mathbin {:=} X \triangledown \mathcal {M}(\mathbf {a}^{(1)})$ commutes with all $M_j$. Since $\mathcal {M}$ is a quasi-Laurent family, there exists $\mathbf {a}^{(2)},\mathbf {b} \in \mathbb {Z}\mspace {1mu}_{\ge 0}^J$ such that $Y \circ \mathcal {M}(\mathbf {a}^{(2)}) \simeq \mathcal {M}(\mathbf {b})$. Hence, by taking $\mathbf {a} = \mathbf {a}^{(1)}+\mathbf {a}^{(2)}$, we have

\[ X \triangledown \mathcal{M}(\mathbf{a}) \simeq \mathcal{M}(\mathbf{b}), \]

as desired.

By applying the similar argument as in the lemma above to composition factors of $X \circ \mathcal {M}(\mathbf {a})$, we have the following corollary.

Corollary 3.4 Let $\mathcal {M}$ be a quasi-Laurent family in $\mathscr {C}$. Then, for any $X\in \mathscr {C}$, there exist $\mathbf {a} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}$, a finite index set $\mathsf {S}$, $c(s) \in \mathbb {Z}\mspace {1mu}$ and $\mathbf {b}(s) \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^J$ $(s \in \mathsf {S})$ such that

(3.1)\begin{equation} [X \circ \mathcal{M}(\mathbf{a})] = \sum_{s \in \mathsf{S}} q^{c(s)}[\mathcal{M}(\mathbf{b}(s))]. \end{equation}

Remark 3.5 The above corollary says that for every module $X$ in $\mathscr {C}$ and a quasi-Laurent family $\mathcal {M} = \{ M_j \mid j \in J\}$, the isomorphism class $[X]$ of $X$ in $K(\mathscr {C})$ can be expressed as an element in the Laurent polynomial ring $\mathbb {Z}\mspace {1mu}[q^{\pm 1}][[M_j]^{\pm 1} \mid j \in J]$ with positive coefficients.

Proposition 3.6 Assume that $\mathcal {K}_\mathbb {A}(\mathscr {C})$ has a quantum cluster algebra structure, and let $\mathcal {M}=\{ M_k \mid k\in \mathsf {K}\}$ be a monoidal cluster in $\mathscr {C}$. Then the commuting family $\mathcal {M}$ is a quasi-Laurent family. In particular, the isomorphism class $[X]$ of $X$ in $K(\mathscr {C})$ can be expressed as an element in the Laurent polynomial $\mathbb {Z}\mspace {1mu}[q^{\pm 1}][\;[M_j] \mid j \in J]$ with positive coefficients.

If, moreover, every $[M_k]$ is prime in $K(\mathscr {C})\vert _{q=1}$, then $\mathcal {M}$ is a Laurent family.

Note that if $K(\mathscr {C})\vert _{q=1}$ is factorial, then every $[M_k]$ is prime in $K(\mathscr {C})\vert _{q=1}$ (see [Reference Geiss, Leclerc and SchröerGLS13b]).

Proof. Let $X$ be a simple module in $\mathscr {C}$ commuting with all $M_k$ $(k\in \mathsf {K})$. The quantum Laurent phenomenon states that there exists $\mathbf {a} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^\mathsf {K}$ such that

(3.2)\begin{equation} [X] = \dfrac{\sum_{s \in \mathsf{S}} c(s)[\mathcal{M}(\mathbf{b}^{(s)})]}{[\mathcal{M}(\mathbf{a})]} \iff [X \circ \mathcal{M}(\mathbf{a})] = \sum_{s \in \mathsf{S}} c(s)[\mathcal{M}(\mathbf{b}^{(s)})] \end{equation}

for some $\ell \in \mathbb {Z}\mspace {1mu}_{\geqslant 1}$, $\mathbf {b}^{(s)} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^\mathsf {K}$ and $c(s) \in \mathbb {Z}\mspace {1mu}[q^{\pm 1/2}]$. Since $X \circ \mathcal {M}(\mathbf {a})$ is simple, the right-hand side of (3.2) must coincide with $q^c\mathcal {M}(\mathbf {c})$ for some $\mathbf {c} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^J$ and $c \in \mathbb {Z}\mspace {1mu}$. Hence, $\mathcal{M}$ is a quasi-Laurent family.

Let us show that $\mathcal {M}$ is a Laurent family. If $X\circ \mathcal {M}(\mathbf {a})\simeq \mathcal {M}(\mathbf {b})$, then we have $\mathbf {a}_k\leqslant \mathbf {b}_k$ for all $k$, since each $[M_k]$ is a prime element of $K(\mathscr {C})\vert _{q=1}$. Hence, setting $\mathbf {c}=\mathbf {b}-\mathbf {a}\in \mathbb {Z}\mspace {1mu}_{\ge 0}^\mathsf {K}$, we have $X\circ \mathcal {M}(\mathbf {a})\simeq \mathcal {M}(\mathbf {c})\circ \mathcal {M}(\mathbf {a})$, which implies that $X\simeq \mathcal {M}(\mathbf {c})$. Hence, $\mathcal {M}$ is a Laurent family.

Definition 3.7 For a simple module $X \in \mathscr {C}$ and a quasi-Laurent family $\mathcal {M}=\{ M_j \mid j \in J \}$, take $\mathbf {a},\mathbf {b},\mathbf {a}',\mathbf {b}' \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^J$ such that $X \triangledown \mathcal {M}(\mathbf {a}) \simeq \mathcal {M}(\mathbf {b})$ and $X \triangle \mathcal {M}(\mathbf {a}') \simeq \mathcal {M}(\mathbf {b}')$ up to grading shifts. Then we define

\[ \mathbf{g}_\mathcal{M}^R(X) \mathbin{:=} \mathbf{b} -\mathbf{a} \quad \text{and} \quad \mathbf{g}_\mathcal{M}^L(X) \mathbin{:=} \mathbf{b}' -\mathbf{a}' \in \mathbb{Z}\mspace{1mu}^J. \]

Remark 3.8

  1. (1) For a quasi-Laurent family $\mathcal {M}$ in $\mathscr {C}$, $\mathbf {g}_\mathcal {M}^R$ and $\mathbf {g}_\mathcal {M}^L$ are well-defined by Lemma 3.1.

  2. (2) For a reduced sequence ${\boldsymbol {i}}$ of $w$ and its quasi-Laurent family $\mathcal {M}^{\boldsymbol {i}}$, we write $\mathbf {g}_{\boldsymbol {i}}^R$ and $\mathbf {g}_{\boldsymbol {i}}^L$ instead of $\mathbf {g}_{\mathcal {M}^{\boldsymbol {i}}}^R$ and $\mathbf {g}_{\mathcal {M}^{\boldsymbol {i}}}^L$, respectively.

  3. (3) The map $\mathbf {g}_{\boldsymbol {i}}^R$ and $\mathbf {g}_{\boldsymbol {i}}^L$ for the quasi-Laurent family $\mathcal {M}^{\boldsymbol {i}}$ can be extended to the set ${\rm Irr}(\widetilde {\mathscr {C}}_w)$ of the isomorphism classes of self-dual simples in $\widetilde {\mathscr {C}}_w$.

The following lemma can be proved by the same arguments in [Reference Kashiwara and KimKK19].

Lemma 3.9 Let $\mathcal {M}$ be a quasi-Laurent family in $\mathscr {C}$ and $X$ a simple module in $\mathscr {C}$.

  1. (i) If $\mathbf {a},\mathbf {b} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^{J}$ satisfy $\mathbf {b}-\mathbf {a} = \mathbf {g}_\mathcal {M}^R(X)$ $($respectively, $\mathbf {b}-\mathbf {a} = \mathbf {g}_\mathcal {M}^L(X))$, then we have

    \[ {X \triangledown \mathcal{M}(\mathbf{a}) \simeq \mathcal{M}(\mathbf{b}) ({\textrm respectively}, X \triangle \mathcal{M}(\mathbf{a}) \simeq \mathcal{M}(\mathbf{b})) textrm {up to a grading shift}.} \]
  2. (ii) For any $\mathbf {a} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^{J}$, we have

    \[ \mathbf{g}_\mathcal{M}^R(X \triangledown \mathcal{M}(\mathbf{a})) = \mathbf{g}_\mathcal{M}^R(X)+\mathbf{a} \quad \text{and}\quad \mathbf{g}_\mathcal{M}^L(X \triangle \mathcal{M}(\mathbf{a})) = \mathbf{g}_\mathcal{M}^L(X)+\mathbf{a}. \]
  3. (iii) The maps $\mathbf {g}_\mathcal {M}^R$ and $\mathbf {g}_\mathcal {M}^L$ from ${\rm Irr}(\mathscr {C})$ to $\mathbb {Z}\mspace {1mu}^{J}$ are injective.

4. PBW decomposition vector and GLS seed

In this section, we recall the PBW basis, and we investigate the relationship between the $g$-vectors and the PBW decomposition vectors.

4.1 PBW decomposition vector

Let us take $w \in \mathsf {W}$ and its reduced sequence ${\boldsymbol {i}}=(i_1,\ldots,i_r)$. Recall the following.

  1. (a) We take an index set $\mathsf {K}=[1,r]$ with a decomposition

    \[ {\mathsf{K}_\mathrm{ex}} \sqcup {\mathsf{K}_\mathrm{fr}} \quad \text{where} \ {\mathsf{K}_\mathrm{ex}}=\{k\in\mathsf{K}\mid k_+\leqslant r\}. \]
  2. (b) For each $1 \leqslant k \leqslant r$, we set ${\mspace {1mu}\beta }^{\boldsymbol {i}}_k \in \Delta ^+ \cap w\Delta ^-$ and define simple modules $S^{\boldsymbol {i}}_k = M(w^{\boldsymbol {i}}_{\leqslant k}\varpi _{i_k},w^{\boldsymbol {i}}_{< k}\varpi _{i_k})$ and $M^{\boldsymbol {i}}_k = M(w^{\boldsymbol {i}}_{\leqslant k}\varpi _{i_k},\varpi _{i_k})$. Note that

    (4.1)\begin{equation} M^{\boldsymbol{i}}_k \simeq S^{\boldsymbol{i}}_k \triangledown M^{\boldsymbol{i}}_{k_-} \quad \text{and} \quad \mathcal{M}^{\boldsymbol{i}} \mathbin{:=} \{M_k^{\boldsymbol{i}}\mid k\in \mathsf{K}\} \text{ forms a commuting family.} \end{equation}

For any $\mathbf {a} = (\mathbf {a}_k)_{1 \leqslant k \leqslant r} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^\mathsf {K}$, the convolution product

\[ P_{\boldsymbol{i}}(\mathbf{a}) \mathbin{:=} q^{{1/2} \sum_{k=1}^r \mathbf{a}_k(\mathbf{a}_k-1) \mathsf{d}_{i_k} } {S^{\boldsymbol{i}}_r}^{\;\circ \mathbf{a}_r} \circ \cdots \circ {S^{\boldsymbol{i}}_1}^{\;\circ \mathbf{a}_1} \]

has a self-dual simple head. Conversely, every self-dual simple module in $\mathscr {C}_w$ is isomorphic to ${\operatorname {hd}}(P_{\boldsymbol {i}}(\mathbf {a}))$ for some $\mathbf {a} \in \mathbb {Z}\mspace {1mu}_{\ge 0}^\mathsf {K}$ in a unique way (see [Reference McNamaraMcN15, Theorem 3.1] and [Reference Tingley and WebsterTW16, Theorem 2.19]). We call $\{ P_{\boldsymbol {i}}(\mathbf {a}) \mid \mathbf {a} \in \mathbb {Z}\mspace {1mu}_{\ge 0}^\mathsf {K} \}$ the PBW basis of $\mathscr {C}_w$ associated with ${\boldsymbol {i}}$.

For a simple module $X$ such that $X\simeq {\operatorname {hd}}(P_{\boldsymbol {i}}(\mathbf {a}))$, we set

\[ {\rm PBW}_{\boldsymbol{i}}(X) \mathbin{:=} \mathbf{a} = (\mathbf{a}_1,\ldots,\mathbf{a}_r). \]

The following lemma says that the operation ${\rm PBW}_{\boldsymbol {i}}( - \triangledown \mathcal {M}^{\boldsymbol {i}}(\mathbf {a}))$ on the set of simple modules behaves very nicely, where $\mathcal {M}^{\boldsymbol {i}}(\mathbf {a})$ is defined in (2.5).

Lemma 4.1 (cf. [Reference Kashiwara and KimKK19, Lemma 3.11 and Proposition 3.14])

For $M=\mathcal {M}^{\boldsymbol {i}}(\mathbf {a})$ with $\mathbf {a} \in \mathbb {Z}\mspace {1mu}_{\ge 0}^{\mathsf {K}}$ and a simple module $X$, we have

\[ {\rm PBW}_{\boldsymbol{i}}(X \triangledown M) ={\rm PBW}_{\boldsymbol{i}}(X)+{\rm PBW}_{\boldsymbol{i}}(M). \]

In particular, $\mathbf {c}\mathbin {:=}{\rm PBW}_{\boldsymbol {i}}\big (\mathcal {M}^{\boldsymbol {i}}(\mathbf {a})\big )$ is given by $\mathbf {c}_k=\sum _{j}\mathbf {a}_j$ where $j$ ranges over $j\in [1,r]$ such that $j\geqslant k$ and ${\boldsymbol {i}}_j={\boldsymbol {i}}_k$.

Proof. It is enough to show it when $M=M^{\boldsymbol {i}}_k$. Note that

\[ X \simeq {\operatorname{hd}}\bigg(\circ_{1 \leqslant k \leqslant r}^{\to} {S^{\boldsymbol{i}}_k}^{\circ \mathbf{n}_k} \bigg) = {S^{\boldsymbol{i}}_{r}}^{\circ \mathbf{n}_r} \triangledown Y, \]

where $\mathbf {n}=(\mathbf {n}_1,\ldots,\mathbf {n}_r) = {\rm PBW}_{\boldsymbol {i}}(X)$, $Y \simeq {\operatorname {hd}}\Big (\displaystyle \circ _{1 \leqslant k \leqslant r-1}^{\to } {S^{\boldsymbol {i}}_k}^{\circ \mathbf {n}_k } \Big )$ and $\displaystyle \circ _{p \leqslant k \leqslant q}^{\to } X_k$ denotes the ordered convolution product $X_q \circ X_{q-1} \circ \cdots X_{p+1} \circ X_p$ for $X_k$ in $R\text {-}\mathrm {gmod}$.

If $r > k$, then $({S^{\boldsymbol {i}}_r}^{\circ \mathbf {n}_r},M^{\boldsymbol {i}}_k)$ is unmixed and, hence, ${S^{\boldsymbol {i}}_r}^{\circ \mathbf {n}_r} \circ Y \circ M^{\boldsymbol {i}}_k$ has a simple head. We have

\begin{align*} X \triangledown M^{\boldsymbol{i}}_k & \simeq {\operatorname{hd}}({S^{\boldsymbol{i}}_r}^{\circ \mathbf{n}_r} \circ Y \circ M^{\boldsymbol{i}}_k) \\ & \simeq {S^{\boldsymbol{i}}_r}^{\circ \mathbf{n}_r} \triangledown (Y \triangledown M^{\boldsymbol{i}}_k) \simeq {S^{\boldsymbol{i}}_r}^{\circ \mathbf{n}_r} \triangledown {\operatorname{hd}}\bigg( \displaystyle\circ_{1 \leqslant k \leqslant r-1}^{\to} {S^{\boldsymbol{i}}_k}^{\circ \mathbf{c}_k} \bigg), \end{align*}

where $\mathbf {c} ={\rm PBW}_{\boldsymbol {i}}(Y)+{\rm PBW}_{\boldsymbol {i}}(M^{\boldsymbol {i}}_k)$ by induction on $r$. Thus, our assertion follows in this case.

If $r = k$, then $M^{\boldsymbol {i}}_r$ commutes with all the objects of $\mathscr {C}_w$ and, hence, we have

\begin{align*} X \triangledown M^{\boldsymbol{i}}_r & \simeq {\operatorname{hd}}({S^{\boldsymbol{i}}_r}^{\circ \mathbf{n}_r} \circ M^{\boldsymbol{i}}_r \circ {S^{\boldsymbol{i}}_{r-1}}^{\circ \mathbf{n}_{r-1}} \circ \cdots {S^{\boldsymbol{i}}_{1}}^{\circ \mathbf{n}_{1}} ) \\ & \simeq {\operatorname{hd}}({S^{\boldsymbol{i}}_r}^{\circ \mathbf{n}_r} \circ M^{\boldsymbol{i}}_r) \triangledown Y \simeq {\operatorname{hd}}({S^{\boldsymbol{i}}_r}^{\circ \mathbf{n}_r} \circ (S^{\boldsymbol{i}}_r \triangledown M^{\boldsymbol{i}}_{r_-})) \triangledown Y \\ & \simeq {\operatorname{hd}}({S^{\boldsymbol{i}}_r}^{\circ \mathbf{n}_r+1} \triangledown M^{\boldsymbol{i}}_{r_-} ) \triangledown Y \simeq {S^{\boldsymbol{i}}_r}^{\circ \mathbf{n}_r+1} \triangledown (M^{\boldsymbol{i}}_{r_-} \circ Y ), \end{align*}

where the last isomorphism follows from the commutativity of $M^{\boldsymbol {i}}_{r_-}$ and $Y$. Then our assertion follows from the induction hypothesis.

The lemma above gives a direct proof of the following corollary although it follows immediately from Proposition 3.6.

Corollary 4.2 For any reduced sequence ${\boldsymbol {i}}$ for $w$, the commuting family $\mathcal {M}^{\boldsymbol {i}}$ is independent.

The following proposition is proved in [Reference Kashiwara and KimKK19, Proposition 2.11] for symmetric quiver Hecke algebra and the same proof also works for the general case.

Proposition 4.3 For any $\mathbf {a} = (\mathbf {a}_1,\ldots,\mathbf {a}_r)$, $\mathbf {b} = (\mathbf {b}_1,\ldots,\mathbf {b}_r) \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^\mathsf {K}$, the ordered sequence

\[ ({S^{\boldsymbol{i}}_r}^{\circ \mathbf{a}_r},(S_{r-1}^{\boldsymbol{i}})^{\circ \mathbf{a}_{r-1}},\ldots, {S^{\boldsymbol{i}}_1}^{\circ \mathbf{a}_1},{M^{\boldsymbol{i}}_1}^{\circ\mathbf{b}_1},{M^{\boldsymbol{i}}_2}^{\circ\mathbf{b}_2},\ldots,{M^{\boldsymbol{i}}_r}^{\circ\mathbf{b}_r}) \]

is a normal sequence.

The statement and proof of following proposition are the same as [Reference Kashiwara and KimKK19, Proposition 3.14] even though [Reference Kashiwara and KimKK19] dealt only with symmetric quiver Hecke algebras. Here we repeat it in order to show relations between explicit $\mathbb {Z}\mspace {1mu}_{\geqslant 0}$-vectors associated with a simple module $X$ in $\mathscr {C}_w$ for the readers’ convenience.

Proposition 4.4 For a simple module $X$ in $\mathscr {C}_w$ or $\widetilde {\mathscr {C}}_w$, there exist $\mathbf {a}, \mathbf {b} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^\mathsf {K}$ such that

\[ X \triangledown \mathcal{M}^{\boldsymbol{i}}(\mathbf{a}) \simeq \mathcal{M}^{\boldsymbol{i}}(\mathbf{b}) \quad \text{up to a grading shift.} \]

Proof. In this proof, we sometimes drop $^{\boldsymbol {i}}$ for notational simplicity. It is enough to consider when $X \in \mathscr {C}_w$ by part (D) in (2.6). Note that there exists a unique $\mathbf {c} = (\mathbf {c}_1,\ldots,\mathbf {c}_r) \in \mathbb {Z}\mspace {1mu}^\mathsf {K}_{\geqslant 0}$ such that

\[ X \simeq {\operatorname{hd}}(P_{\boldsymbol{i}}(\mathbf{c})) \simeq {\operatorname{hd}}({S^{\boldsymbol{i}}_r}^{\circ \mathbf{c}_r} \circ \cdots \circ {S^{\boldsymbol{i}}_1}^{\circ \mathbf{c}_1}) . \]

Set $\mathbf {c}_+ \mathbin {:=} \sum _{k=1}^r \mathbf {c}_{k_+} \mathbf {e}_k= \sum _{j \in \mathsf {K}} \mathbf {c}_k \mathbf {e}_{k_-}$, where $\{ \mathbf {e}_j \mid j \in \mathsf {K}\}$ is the standard basis of $\mathbb {Z}\mspace {1mu}^\mathsf {K}$ such that $\mathbf {c} = \sum _{j \in \mathsf {K}} \mathbf {c}_j \mathbf {e}_j$. Then we have $\mathcal {M}^{\boldsymbol {i}}(\mathbf {c}_+) \simeq M_{1_-}^{\circ \mathbf {c}_1} \circ \cdots \circ M_{r_-}^{\circ \mathbf {c}_r}$. By (4.1) and Proposition 4.3, we have

\begin{align*} X \triangledown \mathcal{M}^{\boldsymbol{i}}(\mathbf{c}_+) &\simeq {\operatorname{hd}}(S_r^{\circ \mathbf{c}_r} \circ S_{r-1}^{\circ \mathbf{c}_{r-1}} \circ \cdots \circ S_1^{\circ \mathbf{c}_{1}} \circ \mathcal{M}^{\boldsymbol{i}}(\mathbf{c}_+) ) \\ &\simeq {\operatorname{hd}}(S_r^{\circ \mathbf{c}_r} \circ S_{r-1}^{\circ \mathbf{c}_{r-1}} \circ \cdots \circ S_2^{\circ \mathbf{c}_{2}} \circ S_1^{\circ \mathbf{c}_{1}} \circ M_{1_-}^{\circ \mathbf{c}_1} \circ M_{2_-}^{\circ \mathbf{c}_2} \circ \cdots \circ M_{r_-}^{\circ \mathbf{c}_r} ) \\ &\simeq {\operatorname{hd}}((S_r^{\circ \mathbf{c}_r} \circ S_{r-1}^{\circ \mathbf{c}_{r-1}} \circ \cdots \circ S_2^{\circ \mathbf{c}_{2}}) \circ (S_1^{\circ \mathbf{c}_{1}} \triangledown M_{1_-}^{\circ \mathbf{c}_1}) \circ (M_{2_-}^{\circ \mathbf{c}_2} \circ \cdots \circ M_{r_-}^{\circ \mathbf{c}_r} )) \\ &\simeq {\operatorname{hd}}((S_r^{\circ \mathbf{c}_r} \circ S_{r-1}^{\circ \mathbf{c}_{r-1}} \circ \cdots \circ S_1^{\circ \mathbf{c}_{2}}) \circ M_{1}^{\circ \mathbf{c}_1} \circ (M_{2_-}^{\circ \mathbf{c}_2} \circ \cdots \circ M_{r_-}^{\circ \mathbf{c}_r} )) \\ &\simeq {\operatorname{hd}}((S_r^{\circ \mathbf{c}_r} \circ S_{r-1}^{\circ \mathbf{c}_{r-1}} \circ \cdots \circ S_2^{\circ \mathbf{c}_{2}}) \circ (M_{2_-}^{\circ \mathbf{c}_2} \circ \cdots \circ M_{r_-}^{\circ \mathbf{c}_r} )\circ M_{1}^{\circ \mathbf{c}_1} ) \\ &\simeq \cdots \simeq {\operatorname{hd}}(M_r^{\circ \mathbf{c}_r} \circ \cdots \circ M_1^{\circ \mathbf{c}_1}) \simeq \mathcal{M}^{\boldsymbol{i}}(\mathbf{c}), \end{align*}

which implies our assertion.

As seen by the proof of the above proposition and Proposition 3.6, we have the following.

Proposition 4.5 The commuting family $\mathcal {M}^{\boldsymbol {i}}$ is a Laurent family. Moreover, for a simple module $M$, two vectors $\mathbf {a}={\rm PBW}_{\boldsymbol {i}}(M)$ and $\mathbf {g}\mathbin {:=}\mathbf {g}_{\boldsymbol {i}}^R(M)$ are related by

\[ \mathbf{g}_k=\mathbf{a}_k-\mathbf{a}_{k_+},\quad \mathbf{a}_k=\sum_{j; j\geqslant k,\ {\boldsymbol{i}}_j={\boldsymbol{i}}_k}\mathbf{g}_j, \]

where $\mathbf {a}_{r+1}=0$.

The following corollary can be proved by the same arguments in [Reference Kashiwara and KimKK19].

Corollary 4.6 Let ${\boldsymbol {i}}$ be a reduced sequence of $w$.

  1. (i) For a dual pair of simples $(L,R)$ in $\widetilde {\mathscr {C}}_w$, we have

    \[ \mathbf{g}_{\boldsymbol{i}}^R(L)+ \mathbf{g}_{\boldsymbol{i}}^L (R)=0. \]
  2. (ii) The maps $\mathbf {g}_{\boldsymbol {i}}^R, \mathbf {g}_{\boldsymbol {i}}^L \colon {\rm Irr}(\widetilde {\mathscr {C}}_w) \to \mathbb {Z}\mspace {1mu}^\mathsf {K}$ are bijective.

5. Skew-symmetric pairings

In this section, we study skew-symmetric pairings induced by the $\mathbb {Z}\mspace {1mu}$-vectors associated with simple modules.

5.1 Skew-symmetric pairing associated with a quasi-Laurent family

Let $\mathscr {C}$ be a full monoidal subcategory of $R\text {-}\mathrm {gmod}$ stable by taking subquotients, extensions and grading shifts, and let $\mathcal {M}=\{ M_j \mid j \in J\}$ be a quasi-Laurent family in $\mathscr {C}$ labeled by a finite index set $J$.

For $X,Y \in {\rm Irr}(\mathscr {C})$, let us define

(5.1) \begin{align} {\mathrm G}^R_\mathcal{M}(X,Y) &\mathbin{:=} \sum\limits_{ a,b \in J} (\mathbf{g}^R_\mathcal{M}(X))_a (\mathbf{g}^R_\mathcal{M}(Y))_b \, \Lambda(M_a,M_b) \quad \text{and}\nonumber\\ {\mathrm G}^L_\mathcal{M}(X,Y) &\mathbin{:=} \sum\limits_{ a,b \in J} (\mathbf{g}^L_\mathcal{M}(X))_a (\mathbf{g}^L_\mathcal{M}(Y))_b \, \Lambda(M_a,M_b). \end{align}

The following lemma immediately follows from Lemma 3.9.

Lemma 5.1 For $M=\mathcal {M}(\mathbf {a})$ with $\mathbf {a} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^{J}$ and $X,Y \in {\rm Irr}(\mathscr {C})$, we have

\begin{align*} &{\mathrm G}^R_\mathcal{M}(X \triangledown M,Y) = {\mathrm G}^R_\mathcal{M}(X,Y)+ {\mathrm G}^R_\mathcal{M}(M,Y), \quad {\mathrm G}^R_\mathcal{M}(X,Y) = -{\mathrm G}^R_\mathcal{M}(Y,X), \\ &{\mathrm G}^L_\mathcal{M}(X \triangle M,Y) = {\mathrm G}^L_\mathcal{M}(X,Y)+ {\mathrm G}^L_\mathcal{M}(M,Y), \quad {\mathrm G}^L_\mathcal{M}(X,Y) = -{\mathrm G}^L_\mathcal{M}(Y,X). \end{align*}

Proposition 5.2 Let $X$ be a simple module in $\mathscr {C}$. Then for any $\mathbf {c}\in \mathbb {Z}\mspace {1mu}_{\ge 0}^J$, we have

\[ {\rm{(i)}} \ \Lambda(X,\mathcal{M}(\mathbf{c})) ={\mathrm G}^R_\mathcal{M}(X,\mathcal{M}(\mathbf{c})) \quad \text{and} \quad {\rm{(ii)}}\ \Lambda(\mathcal{M}(\mathbf{c}),X) ={\mathrm G}^L_\mathcal{M}(\mathcal{M}(\mathbf{c}),X) . \]

Proof. If $X$ is also of the form $\mathcal {M}(\mathbf {d})$, it is obvious. Set $Y=\mathcal {M}(\mathbf {c})$.

  1. (i) Note that there exist $\mathbf {a},\mathbf {b} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^{J}$ such that $X \triangledown \mathcal {M}(\mathbf {a}) \simeq \mathcal {M}(\mathbf {b})$. Then we have

    \begin{align*} \Lambda(X,Y)+\Lambda(\mathcal{M}(\mathbf{a}),Y) = \Lambda(X\triangledown \mathcal{M}(\mathbf{a}),Y) & = {\mathrm G}^R_\mathcal{M}(X\triangledown \mathcal{M}(\mathbf{a}),Y) \\ & = {\mathrm G}^R_\mathcal{M}(X,Y) + {\mathrm G}^R_\mathcal{M}(\mathcal{M}(\mathbf{a}),Y). \end{align*}
    Since $\Lambda (\mathcal {M}(\mathbf {a}),Y) = {\mathrm G}^R_\mathcal {M}(\mathcal {M}(\mathbf {a}),Y)$, our assertion follows.
  2. (ii) Similarly, there exist $\mathbf {a},\mathbf {b} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^{J}$ such that $\mathcal {M}(\mathbf {a}) \triangledown X \simeq \mathcal {M}(\mathbf {b})$. Then we have

    \begin{align*} \Lambda(Y,\mathcal{M}(\mathbf{a})) + \Lambda(Y,X) = \Lambda(Y,\mathcal{M}(\mathbf{a})\triangledown X) & = {\mathrm G}^L_\mathcal{M}(Y,\mathcal{M}(\mathbf{a})\triangledown X) \\ & = {\mathrm G}^L_\mathcal{M}(Y,\mathcal{M}(\mathbf{a})) + {\mathrm G}^L_\mathcal{M}(Y,X). \end{align*}
    Then our assertion follows from the fact that $\Lambda (Y,\mathcal {M}(\mathbf {a})) = {\mathrm G}^L_\mathcal {M}(Y,\mathcal {M}(\mathbf {a}))$.

Proposition 5.3 For any simple modules $X,Y$ in $\mathscr {C}$ such that one of them is affreal, we have

\[ {\mathrm G}^R_\mathcal{M}(X,Y), {\mathrm G}^L_\mathcal{M}(X,Y) \leqslant \Lambda(X,Y). \]

Proof. Since the proofs are similar, we will consider the case of ${\mathrm G}^R_\mathcal {M}$. Take $\mathbf {a},\mathbf {b}\in \mathbb {Z}\mspace {1mu}_{\ge 0}^J$ such that $Y \triangledown \mathcal {M}(\mathbf {a}) \simeq \mathcal {M}(\mathbf {b})$. Then we have

\begin{align*} {\mathrm G}^R_\mathcal{M}(X,Y) + {\mathrm G}^R_\mathcal{M}(X,\mathcal{M}(\mathbf{a}))& = {\mathrm G}^R_\mathcal{M}(X,Y\triangledown\mathcal{M}(\mathbf{a})) \\ & = \Lambda(X,Y \triangledown \mathcal{M}(\mathbf{a})) \leqslant \Lambda(X,Y)+ \Lambda(X,\mathcal{M}(\mathbf{a})) \\ & = \Lambda(X,Y)+ {\mathrm G}^R_\mathcal{M}(X,\mathcal{M}(\mathbf{a})) , \end{align*}

which yields our assertion. Here, the inequality follows from [Reference Kang, Kashiwara, Kim and OhKKKO18, Proposition 3.2.10].

Proposition 5.4 If simple modules $X$ and $Y$ in $\mathscr {C}$ commute and one of them is affreal, then we have

\[ \Lambda(X,Y) = {\mathrm G}^R_\mathcal{M}(X,Y) = {\mathrm G}^L_\mathcal{M}(X,Y). \]

Proof. Since the proofs are similar, we will only give the proof for ${\mathrm G}^R_\mathcal {M}$. By the preceding proposition, we have

\begin{align*} 0&=\big(\Lambda(X,Y)+\Lambda(Y,X)\big)-\big({\mathrm G}^R_\mathcal{M}(X,Y)+{\mathrm G}^R_\mathcal{M}(Y,X)\big)\\ &=\big(\Lambda(X,Y)-{\mathrm G}^R_\mathcal{M}(X,Y)\big)+\big( \Lambda(Y,X)-{\mathrm G}^R_\mathcal{M}(Y,X)\big) \ge0, \end{align*}

which implies $\Lambda (X,Y)-{\mathrm G}^R_\mathcal {M}(X,Y)=0$.

Remark 5.5 The two invariants ${{\mathrm G}}^R_{{\mathcal {M}}}(X,Y)$ and ${{\mathrm G}}^L_{{\mathcal {M}}}(X,Y)$ are different in general and depend on the choice of $\mathcal {M}$.

Let $w_0$ be the longest element of finite type $A_2$. For a reduced sequence ${\boldsymbol {i}}=(1,2,1)$ of $w_0$, we have

(5.2)\begin{equation} \{\!S_1^{{\boldsymbol{i}}}=\langle 1\rangle , S_2^{{\boldsymbol{i}}}=\langle 12\rangle , S_3^{{\boldsymbol{i}}}=\langle 2\rangle \}\quad \text{and}\quad \mathcal{M}^{\boldsymbol{i}} =\{ M_1^{{\boldsymbol{i}}}=\langle 1\rangle , M_2^{{\boldsymbol{i}}}=\langle 12\rangle , M_3^{{\boldsymbol{i}}}=\langle 21\rangle\!\}, \end{equation}

while

(5.3)\begin{equation} \{\!S_1^{\boldsymbol{j}}=\langle 2\rangle , S_2^{\boldsymbol{j}}=\langle 21\rangle , S_3^{\boldsymbol{j}}=\langle 1\rangle \}\quad \text{and}\quad \mathcal{M}^{\boldsymbol{j}} =\{ M_1^{\boldsymbol{j}}=\langle 2\rangle , M_2^{\boldsymbol{j}}=\langle 21\rangle , M_3^{\boldsymbol{j}}=\langle 12\rangle\!\} \end{equation}

for the other reduced sequence ${\boldsymbol {j}}=(2,1,2)$ of $w_0$. Here $\langle k\rangle$ $(k=1,2)$ is a one-dimensional $R({\mspace {1mu}\alpha }_k)$-module, and $\langle 12 \rangle$ and $\langle 21 \rangle$ are one-dimensional $R({\mspace {1mu}\alpha }_1+{\mspace {1mu}\alpha }_2)$-modules (see [Reference Kang, Kashiwara and KimKKK18] for more details on these modules).

Since $A_2$ is symmetric, $\mathcal {M}' \mathbin {:=} \mu _1(\mathcal {M}^{\boldsymbol {i}})$ is also a Laurent family given as follows:

\[ \mathcal{M}' = \{ M'_1 =\mu_1(M^{\boldsymbol{i}}_1) \simeq \langle 2\rangle, M'_2 = M^{\boldsymbol{i}}_2 \simeq \langle 12 \rangle, M'_3 = M^{\boldsymbol{i}}_3 \simeq \langle 21 \rangle\}. \]

Note that $\mathcal {M}'=\mathcal {M}^{\boldsymbol {j}}$ (up to an index permutation). We have

\begin{align*} & \mathbf{g}^R_{\mathcal{M}^{\boldsymbol{i}}}(\langle 1 \rangle) = \mathbf{g}^L_{\mathcal{M}^{\boldsymbol{i}}}(\langle 1 \rangle) = (1,0,0), \quad \mathbf{g}^R_{\mathcal{M}^{\boldsymbol{i}}}(\langle 2 \rangle ) = (-1,0,1), \quad \mathbf{g}^L_{\mathcal{M}^{\boldsymbol{i}}}(\langle 2 \rangle ) = (-1,1,0), \\ & \mathbf{g}^R_{\mathcal{M}'}(\langle 1 \rangle) = (-1,1,0), \quad \mathbf{g}^L_{\mathcal{M}'}(\langle 1 \rangle) = (-1,0,1), \quad \mathbf{g}^R_{\mathcal{M}'}(\langle 2 \rangle) = \mathbf{g}^L_{\mathcal{M}'}(\langle 2 \rangle) = (1,0,0), \end{align*}

and

\begin{align*} & \Lambda(\langle 1 \rangle,\langle 2\rangle) = \Lambda(\langle 2\rangle,\langle 1\rangle)=1, \quad \Lambda(\langle 1\rangle,\langle 12\rangle)=1, \\ & \Lambda(\langle 1\rangle,\langle 21\rangle) =-1, \quad \Lambda(\langle 12\rangle,\langle 2\rangle) = 1, \quad \Lambda(\langle 21\rangle,\langle 2\rangle) = -1. \end{align*}

Note that $\Lambda (X,X)=0$ for an affreal simple module $X$. Thus, we have

\begin{align*} \begin{array}{ll} {\mathrm G}^R_{\mathcal{M}^{\boldsymbol{i}}}(\langle 1\rangle,\langle 2\rangle) = 1 \times \Lambda(\langle 1\rangle,\langle 21\rangle) =-1, \quad & {\mathrm G}^R_{\mathcal{M}'}(\langle 1\rangle,\langle 2\rangle) =1 \times \Lambda(\langle 12 \rangle,\langle 2 \rangle) =1, \\ {\mathrm G}^L_{\mathcal{M}^{\boldsymbol{i}}}(\langle 1\rangle,\langle 2\rangle) = 1 \times \Lambda(\langle 1\rangle,\langle 12\rangle) =1, \quad & {\mathrm G}^L_{\mathcal{M}'}(\langle 1\rangle,\langle 2\rangle) =1 \times \Lambda(\langle 21 \rangle,\langle 2 \rangle) =-1. \end{array} \end{align*}

Thus, for a non-commuting pair of simple modules $(X,Y)$ in $\mathscr {C}$, the $\mathbb {Z}\mspace {1mu}$-values ${\mathrm G}^R_\mathcal {M}(X,Y)$ and ${\mathrm G}^L_\mathcal {M}(X,Y)$ do depend on the choice of a quasi-Laurent commuting family $\mathcal {M}$.

5.2 Skew-symmetric pairing associated with the GLS cluster

Let $w \in \mathsf {W}$ and ${\boldsymbol {i}}=(i_1,\ldots,i_r)$ a reduced sequence of $w$. Let $\mathcal {M}^{\boldsymbol {i}}$ be the associated GLS cluster. For such a Laurent family, we can define ${\mathrm G}_{\mathcal {M}^{\boldsymbol {i}}}^R$ in terms of PBW decompositions.

We define a skew-symmetric $\mathbb {Z}\mspace {1mu}$-valued map $\lambda ^{\boldsymbol {i}}\colon [1,r]\times [1,r]\to \mathbb {Z}\mspace {1mu}$ by

(5.4)\begin{equation} \lambda^{\boldsymbol{i}} _{a,b} \mathbin{:=} (-1)^{\delta(a>b)} \delta(a \ne b) ({\mspace{1mu}\beta}^{\boldsymbol{i}}_a,{\mspace{1mu}\beta}^{\boldsymbol{i}}_b) \end{equation}

for $1 \leqslant a,b\leqslant r$.

Remark 5.6 The skew-symmetric map $\lambda ^{\boldsymbol {i}}$ in (5.4) is known when $\mathfrak {g}$ is of finite type and ${\boldsymbol {i}}$ is adapted to a $Q$-datum (see [Reference Hernandez and LeclercHL15, Proposition 3.2], [Reference Fujita and OhFO21, Proposition 5.21] and [Reference Kashiwara and OhKaOh23, Theorem 5.4]).

Let us recall the notion of ${\boldsymbol {i}}$-box and an affreal simple module $M^{\boldsymbol {i}}[a,b]$ in $\mathscr {C}_w$ for an ${\boldsymbol {i}}$-box $[a,b]$, which are introduced in [Reference Kashiwara, Kim, Oh and ParkKKOP24b].

  1. (a) For $1 \leqslant a \leqslant b \leqslant r$ such that $i_a=i_b$, we call an interval $[a,b]$ an ${\boldsymbol {i}}$-box.

  2. (b) For an ${\boldsymbol {i}}$-box $[a,b]$, we set $[a,b]_{\boldsymbol {i}} \mathbin {:=} \{ u \mid a \leqslant u \leqslant b, \; i_a=i_u\}$.

  3. (c) For an ${\boldsymbol {i}}$-box $[a,b]$, we set

    \begin{align*} M^{\boldsymbol{i}}[a,b] \mathbin{:=} M(w^{\boldsymbol{i}}_{\leqslant b}\varpi_{i_a},w^{\boldsymbol{i}}_{< a}\varpi_{i_a})& \simeq {\operatorname{hd}}\Big( \circ_{u \in [a,b]_{\boldsymbol{i}}}^{\to} S^{\boldsymbol{i}}_u \Big) \\ &\simeq S^{\boldsymbol{i}}_b \triangledown M^{\boldsymbol{i}}[a,b_-] \simeq M^{\boldsymbol{i}}[a_+,b] \triangledown S^{\boldsymbol{i}}_a, \end{align*}
    up to grading shifts. In particular, $M^{\boldsymbol {i}}_k = M^{\boldsymbol {i}}[k_{\min },k]$ and $S^{\boldsymbol {i}}_k = M^{\boldsymbol {i}}[k,k]$.

Note that $M^{\boldsymbol {i}}[a,b]$ is an affreal simple module in $\mathscr {C}_w$.

Proposition 5.7 For ${\boldsymbol {i}}$-boxes $[x,y]$ and $[x',y']$ in an interval $[1,r]$, assume that

(5.5)\begin{equation} {\rm{(a)}} \ \ x > x'_- \quad \text{or} \quad {\rm{(b)}} \ \ y_+ > y'. \end{equation}

Then we have

(5.6)\begin{equation} \Lambda( M^{\boldsymbol{i}}[x,y], M^{\boldsymbol{i}}[x',y']) = \sum_{ u \in [x,y]_{\boldsymbol{i}} , \ v \in [x',y']_{\boldsymbol{i}} } \lambda_{u,v}. \end{equation}

Proof. Since the proof is similar, we shall give only the proof of case (a). Let us divide into sub-cases as below.

(i) [$x=y > x'_-$]  If $x>x'$, we have

\begin{align*} \Lambda(S^{\boldsymbol{i}}_x, M^{\boldsymbol{i}}[x',y']) & = \Lambda(S^{\boldsymbol{i}}_x,M^{\boldsymbol{i}}[x'_+,y'] \triangledown S^{\boldsymbol{i}}_{x'}) \\ & \underset{(1)}= \Lambda(S^{\boldsymbol{i}}_x,M^{\boldsymbol{i}}[ x'_+,y'])+ \Lambda(S^{\boldsymbol{i}}_x, S^{\boldsymbol{i}}_{x'})= \Lambda(S^{\boldsymbol{i}}_x,M^{\boldsymbol{i}}[x'_+,y'])+ \lambda^{\boldsymbol{i}}_{x,x'}. \end{align*}

Here $\underset {(1)}=$ holds by [Reference Kashiwara, Kim, Oh and ParkKKOP23, Proposition 2.12] and the fact that $(S^{\boldsymbol {i}}_x,S^{\boldsymbol {i}}_{x'})$ is an unmixed pair. Then by the induction hypothesis on $|[x',y']_{\boldsymbol {i}}|$, we have

\[ \Lambda(S^{\boldsymbol{i}}_x, M^{\boldsymbol{i}}[x',y']) = \lambda^{\boldsymbol{i}}_{x,x'} + \sum_{v \in [x'_+,y']_{\boldsymbol{i}}} \lambda^{\boldsymbol{i}}_{x,v} = \sum_{v \in [x', y' ]_{\boldsymbol{i}}} \lambda^{\boldsymbol{i}}_{x,v}, \]

as we desired.

Now, the remainder of case (i) can be described as follows:

\[ x'_-< x =y \leqslant x' \leqslant y'. \]

Since $S^{\boldsymbol {i}}_x$ commutes with $M^{\boldsymbol {i}}[x',y']$ and $M^{\boldsymbol {i}}[x',y'_-]$ by [Reference Kashiwara, Kim, Oh and ParkKKOP21, Proposition 3.27],

\begin{align*} \Lambda(S^{\boldsymbol{i}}_x,M^{\boldsymbol{i}}[x',y']) & = - \Lambda(M^{\boldsymbol{i}}[x',y'],S^{\boldsymbol{i}}_x) = -\Lambda( S^{\boldsymbol{i}}_{y'} \triangledown M^{\boldsymbol{i}}[x',y'_-],S^{\boldsymbol{i}}_x) \\ & = -\Lambda( S^{\boldsymbol{i}}_{y'} ,S^{\boldsymbol{i}}_x) -\Lambda( M^{\boldsymbol{i}}[x',y'_-],S^{\boldsymbol{i}}_x) \\ & = ({\mspace{1mu}\beta}_{y'},{\mspace{1mu}\beta}_x) + \Lambda( S^{\boldsymbol{i}}_x, M^{\boldsymbol{i}}[x',y'_-]) = \lambda^{\boldsymbol{i}}_{x,y'} + \Lambda( S^{\boldsymbol{i}}_x, M^{\boldsymbol{i}}[x',y'_-]), \end{align*}

then our assertion follows from the induction hypothesis on $|\,[x',y']_{\boldsymbol {i}}|$.

(ii) [$x< y$] Assume first that $y>y'$. Then we have

\begin{align*} \Lambda(M^{\boldsymbol{i}}[x,y],M^{\boldsymbol{i}}[x',y']) & = \Lambda(S^{\boldsymbol{i}}_y \triangledown M^{\boldsymbol{i}}[x,y_-],M^{\boldsymbol{i}}[x',y']) \\ & = \Lambda(S^{\boldsymbol{i}}_y,M^{\boldsymbol{i}}[x',y']) + \Lambda(M^{\boldsymbol{i}}[x,y_-],M^{\boldsymbol{i}}[x',y']). \end{align*}

Note that $(S^{\boldsymbol {i}}_y,M^{\boldsymbol {i}}[x',y'])$ is an unmixed pair. Then, by induction on $|[x,y]_{\boldsymbol {i}}|$, we have

\begin{align*} \Lambda(M^{\boldsymbol{i}}[x,y],M^{\boldsymbol{i}}[x',y'])& = \Lambda(S^{\boldsymbol{i}}_y \triangledown M^{\boldsymbol{i}}[x,y_-],M^{\boldsymbol{i}}[x',y'])\\ & = \sum_{v \in [x',y']_{\boldsymbol{i}}} \lambda^{\boldsymbol{i}}_{y,v} + \sum_{ u \in [x,y_-]_{\boldsymbol{i}}; \ v \in [x',y']_{\boldsymbol{i}} } \lambda^{\boldsymbol{i}}_{u,v}, \end{align*}

which yields our assertion for this case.

Now let us assume that $y \leqslant y'$. Then we have

\[ x'_- < x < y \leqslant y'. \]

Then for any $u \in [x,y]_{\boldsymbol {i}}$, $S^{\boldsymbol {i}}_u$ commutes with $M^{\boldsymbol {i}}[x',y']$ by [Reference Kashiwara, Kim, Oh and ParkKKOP21, Proposition 3.27]. By [Reference Kang, Kashiwara, Kim and OhKKKO18, Proposition 3.2.13], we have

\[ \Lambda(M^{\boldsymbol{i}}[x,y],M^{\boldsymbol{i}}[x',y']) = \sum_{u \in [x,y]_{\boldsymbol{i}}} \Lambda(S^{\boldsymbol{i}}_u,M^{\boldsymbol{i}}[x',y']). \]

Then our assertion follows from case (i).

We say that ${\boldsymbol {i}}$-boxes $[a_1,b_1]$ and $[a_2,b_2]$ commute if we have either

\[ (a_1)_- < a_2 \leqslant b_2 < (b_1)_+ \quad \text{or} \quad (a_2)_- < a_1 \leqslant b_1 < (b_2)_+. \]

The following corollary is proved in [Reference Kashiwara, Kim, Oh and ParkKKOP24b, Theorem 4.21] in the quantum affine case.

Corollary 5.8 For commuting ${\boldsymbol {i}}$-boxes $[a_1,b_1]$ and $[a_2,b_2]$, the modules $M^{\boldsymbol {i}}[a_1,b_1]$ and $M^{\boldsymbol {i}}[a_2,b_2]$ commute.

Proof. By Proposition 5.7, we have

\[ \Lambda(M^{\boldsymbol{i}}[a_1,b_1],M^{\boldsymbol{i}}[a_2,b_2]) = \sum_{\substack{u \in [a_1,b_1]_{\boldsymbol{i}} \\ v \in [a_2,b_2]_{\boldsymbol{i}} }} \lambda^{\boldsymbol{i}}_{u,v} = - \Lambda(M^{\boldsymbol{i}}[a_2,b_2],M^{\boldsymbol{i}}[a_1,b_1]), \]

which implies $\mathfrak {d}(M^{\boldsymbol {i}}[a_1,b_1],M^{\boldsymbol {i}}[a_2,b_2])=0$. Thus, our assertion follows from Proposition 2.7(v).

Proposition 5.9 For a commuting pair $(M^{\boldsymbol {i}}[x,y],M^{\boldsymbol {i}}[x',y'])$(5.6) holds.

Proof. If the ${\boldsymbol {i}}$-boxes $[x,y]$ and $[x',y']$ satisfy (5.5), our assertion holds. Thus, it is enough to consider when $x \leqslant x'_-$ and $y_+ \leqslant y'$. Since they commute,

\[ \Lambda(M^{\boldsymbol{i}}[x,y],M^{\boldsymbol{i}}[x',y']) =- \Lambda(M^{\boldsymbol{i}}[x',y'],M^{\boldsymbol{i}}[x,y]). \]

If $x'> x_-$ or $y'_+> y$, Proposition 5.7 says that

\[ \Lambda(M^{\boldsymbol{i}}[x,y],M^{\boldsymbol{i}}[x',y']) = - \sum_{ u \in [x,y]_{\boldsymbol{i}} ; \ v \in [x',y']_{\boldsymbol{i}} } \lambda^{\boldsymbol{i}}_{v,u} = \sum_{ u \in [x,y]_{\boldsymbol{i}} ; \ v \in [x',y']_{\boldsymbol{i}} } \lambda^{\boldsymbol{i}}_{u,v}, \]

which implies the assertion. Thus, we may assume that $x' \leqslant x_-$. However, in this case, we have

\[ x' \leqslant x_- \leqslant x \leqslant x'_-, \]

which yields a contradiction.

Let us define the skew-symmetric pairing ${\mathrm L}_{\boldsymbol {i}}$ on ${\rm Irr}(\mathscr {C}_w)$ as follows:

(5.7)\begin{equation} {\mathrm L}_{\boldsymbol{i}}(X,Y) \mathbin{:=} \sum_{1 \leqslant a,b \leqslant r} ({\rm PBW}_{\boldsymbol{i}}(X))_a ({\rm PBW}_{\boldsymbol{i}}(Y))_b \; \lambda^{\boldsymbol{i}}_{a,b}. \end{equation}

The following lemma follows from Lemma 4.1 and (5.7).

Lemma 5.10 For $M=\mathcal {M}^{\boldsymbol {i}}(\mathbf {a})$ with $\mathbf {a} \in \mathbb {Z}\mspace {1mu}_{\ge 0}^{\mathsf {K}}$, we have

\[ {\mathrm L}_{\boldsymbol{i}}(X \triangledown M,Y) = {\mathrm L}_{\boldsymbol{i}}(X,Y)+ {\mathrm L}_{\boldsymbol{i}}(M,Y) \quad \text{and} \quad {\mathrm L}_{\boldsymbol{i}}(X,Y) = -{\mathrm L}_{\boldsymbol{i}}(Y,X). \]

Proposition 5.11 For any simple $X,Y$ in $\mathscr {C}_w$, we have

\[ {\mathrm L}_{\boldsymbol{i}}(X,Y)={\mathrm G}^R_{\mathcal{M}^{\boldsymbol{i}}}(X,Y). \]

Proof. Let $\mathcal {S}$ be the set of simple modules $Y$ in $\mathscr {C}_w$ such that ${\mathrm L}_{\boldsymbol {i}}(X,Y)={\mathrm G}^R_{\mathcal {M}^{\boldsymbol {i}}}(X,Y)$ for any simple $X\in \mathscr {C}_w$, and let $\mathcal {S}'$ be the set of simple modules $Y$ in $\mathscr {C}_w$ such that ${\mathrm L}_{\boldsymbol {i}}(\mathcal {M}^{\boldsymbol {i}}(\mathbf {a}),Y)={\mathrm G}^R_{\mathcal {M}^{\boldsymbol {i}}}(\mathcal {M}^{\boldsymbol {i}}(\mathbf {a}),Y)$ for any $\mathbf {a}\in \mathbb {Z}\mspace {1mu}_{\ge 0}^\mathsf {K}$. By Proposition 5.9, we have

\[ {\mathrm L}_{\boldsymbol{i}}(M^{\boldsymbol{i}}_s,M^{\boldsymbol{i}}_t)=\Lambda(M^{\boldsymbol{i}}_s,M^{\boldsymbol{i}}_t) \quad \text{for any $s,t \in \mathsf{K}$.} \]

Thus, we have $\mathcal {M}^{\boldsymbol {i}}(\mathbf {a}) \in \mathcal {S}'$ by Lemma 5.10. Now, let us show $\mathcal {S}'\subset \mathcal {S}$. Let $Y\in \mathcal {S}'$. For any simple $X$, there exist $\mathbf {a},\mathbf {b}\in \mathbb {Z}\mspace {1mu}_{\ge 0}^\mathsf {K}$ such that $X\triangledown \mathcal {M}^{\boldsymbol {i}}(\mathbf {a})\simeq \mathcal {M}^{\boldsymbol {i}}(\mathbf {b})$. Hence, we have

\begin{align*} {\mathrm L}_{\boldsymbol{i}}(X,Y)+{\mathrm L}_{\boldsymbol{i}}(\mathcal{M}^{\boldsymbol{i}}(\mathbf{a}),Y)&\underset{(1)}={\mathrm L}_{\boldsymbol{i}}(\mathcal{M}^{\boldsymbol{i}}(\mathbf{b}),Y)= {\mathrm G}^R_{\mathcal{M}^{\boldsymbol{i}}}(\mathcal{M}^{\boldsymbol{i}}(\mathbf{b}),Y)\\ &\underset{(2)}={\mathrm G}^R_{\mathcal{M}^{\boldsymbol{i}}}(X,Y)+{\mathrm G}^R_{\mathcal{M}^{\boldsymbol{i}}}(\mathcal{M}^{\boldsymbol{i}}(\mathbf{a}),Y)= {\mathrm G}^R_{\mathcal{M}^{\boldsymbol{i}}}(X,Y)+{\mathrm L}_{\boldsymbol{i}}(\mathcal{M}^{\boldsymbol{i}}(\mathbf{a}),Y). \end{align*}

Here $\underset {(1)}=$ follows from Lemma 5.10 and $\underset {(2)}=$ follows from Lemma 5.1. Hence, we have ${\mathrm L}_{\boldsymbol {i}}(X,Y)={\mathrm G}^R_{\mathcal {M}^{\boldsymbol {i}}}(X,Y)$. Thus, we have proved $\mathcal {S}'\subset \mathcal {S}$.

Since $\mathcal {M}^{\boldsymbol {i}}(\mathbf {a})\in \mathcal {S}'$, we have $\mathcal {M}^{\boldsymbol {i}}(\mathbf {a})\in \mathcal {S}$, which implies that

\[ {\mathrm L}_{\boldsymbol{i}}(Y,\mathcal{M}^{\boldsymbol{i}}(\mathbf{a}))={\mathrm G}^R_{\mathcal{M}^{\boldsymbol{i}}}(Y,\mathcal{M}^{\boldsymbol{i}}(\mathbf{a})) \]

for any simple $Y$. Hence any simple $Y$ belongs to $\mathcal {S}'$ and hence to $\mathcal {S}$.

5.3 Degree and codegree

In this subsection, we see the relationship between $\mathbf {g}^R_\mathcal {M}(X)$ (respectively, $\mathbf {g}^L_\mathcal {M}(X)$) and the degree (respectively, codegree) in the (quantum) cluster algebra theory. For a commuting family $\mathcal {M}=\{ M_j \mid j \in J \}$ labeled by a finite index set $J$, let us recall the preorder $\preccurlyeq _\mathcal {M}$ on $\mathbb {Z}\mspace {1mu}_{\geqslant 0}^{J}$ given in [Reference Kashiwara and KimKK19, § 3.3] (see also [Reference QinQin17, Definition 3.1.1]):

\begin{align*} \mathbf{b}' \preccurlyeq_\mathcal{M} \mathbf{b} \quad \text{if and only if} \quad & {\rm (1)} \ {\rm wt}(\mathcal{M}(\mathbf{b})) ={\rm wt}(\mathcal{M}(\mathbf{b}')), \\ & {\rm (2)} \ \Lambda(\mathcal{M}(\mathbf{b}'),M_j) \leqslant \Lambda(\mathcal{M}(\mathbf{b}),M_j) \text{ for all } j \in J. \end{align*}

The preorder $\preccurlyeq _\mathcal {M}$ can be extended to the one on $\mathbb {Z}\mspace {1mu}^{J}$ as follows: for $\mathbf {b},\mathbf {b}' \in \mathbb {Z}\mspace {1mu}^{J}$,

\[ \mathbf{b}' \preccurlyeq_\mathcal{M} \mathbf{b} \text{ if } \mathbf{b}'+\mathbf{a} \preccurlyeq_\mathcal{M} \mathbf{b}+\mathbf{a} \text{ for some } \mathbf{a} \in \mathbb{Z}\mspace{1mu}_{\geqslant 0}^{J} \text{ such that } \mathbf{b}+\mathbf{a},\mathbf{b}'+\mathbf{a} \in \mathbb{Z}\mspace{1mu}_{\geqslant 0}^{J}. \]

We write $\mathbf {b}'\prec _\mathcal {M}\mathbf {b}$, if $\mathbf {b}'\preccurlyeq _\mathcal {M}\mathbf {b}$ holds but $\mathbf {b}\preccurlyeq _\mathcal {M}\mathbf {b}'$ does not hold. Hence, $\mathbf {b}'\prec _\mathcal {M}\mathbf {b}$ if and only if $\mathbf {b}'\preccurlyeq _\mathcal {M}\mathbf {b}$ and there exists $j\in J$ such that $\Lambda (\mathcal {M}(\mathbf {b}),M_j)<\Lambda (\mathcal {M}(\mathbf {b}'),M_j)$.

Lemma 5.12 (cf. [Reference Kashiwara and KimKK19, Lemma 3.6])

Let $X$ be a simple module and $\mathcal {M}=\{ M_j \mid j \in J\}$ be a quasi-Laurent family in $\mathscr {C}$. Let $\mathbf {a} \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}$, $\mathsf {S}$, $c(s) \in \mathbb {Z}\mspace {1mu}$ and $\mathbf {b}(s) \in \mathbb {Z}\mspace {1mu}_{\geqslant 0}^J$ $(s \in \mathsf {S})$ be as in (3.1). Then we have the following.

  1. (i) There exists a unique $s_0 \in \mathsf {S}$ such that $X \triangledown \mathcal {M}(\mathbf {a}) \simeq q^{\,c(s_0)} \mathcal {M}(\mathbf {b}(s_0))$. Moreover, we have $\mathbf {b}(s) \prec _\mathcal {M} \mathbf {b}(s_0)$ for any $s \in \mathsf {S} \setminus \{ s_0 \}$.

  2. (ii) There exists a unique $s_1 \in \mathsf {S}$ such that $X \triangle \mathcal {M}(\mathbf {a}) \simeq q^{\,c(s_1)} \mathcal {M}(\mathbf {b}(s_1))$. Moreover, we have $\mathbf {b}(s_1) \prec _\mathcal {M} \mathbf {b}(s)$ for any $s \in \mathsf {S} \setminus \{ s_1 \}$.

  3. (iii) If $s_0 = s_1$, then $\mathsf {S} = \{ s_0\}$ and $X \circ \mathcal {M}(\mathbf {a}) \simeq \mathcal {M}(\mathbf {b}(s_0))$.

  4. (iv) If $s_0 \ne s_1$ and there exists no $\mathbf {c} \in \mathbb {Z}\mspace {1mu}^\mathsf {K}$ such that $\mathbf {b}(s_1) - \mathbf {a} \prec _\mathcal {M} \mathbf {c} \prec _\mathcal {M} \mathbf {b}(s_0) - \mathbf {a}$, then

    \[ [X \circ \mathcal{M}(\mathbf{a})] = [X \triangledown \mathcal{M}(\mathbf{a})] + [X \triangle \mathcal{M}(\mathbf{a})] \quad \text{in $K(\mathscr{C}_w)$.} \]

Proof. It follows from Proposition 2.7 and (3.1).

The following proposition is proved for symmetric quiver Hecke algebras and can be extended to general quiver Hecke algebras using almost the same argument:

Proposition 5.13 [Reference Kashiwara and KimKK19, Proposition 3.3]

For a monoidal cluster $\mathcal {M}=\{ M_k \mid k \in \mathsf {K}\}$ associated with a quantum seed $(\{ X_k\}_{k\in \mathsf {K}}, L,\widetilde {B})$,

\[ \mathbf{b}' \preccurlyeq_\mathcal{M} \mathbf{b} \quad \text{if and only if} \quad \mathbf{b}-\mathbf{b}' = \widetilde{B} {\underline{v}} \text{ for some } {\underline{v}} \in \mathbb{Z}\mspace{1mu}_{\ge0}^{\mathsf{K}_\mathrm{ex}}. \]

In particular, the relation $\preccurlyeq _\mathcal {M}$ is an order on $\mathbb {Z}\mspace {1mu}^\mathsf {K}$.

Corollary 5.14 Let $\mathcal {M}=\{ M_k \mid k \in \mathsf {K}\}$ be a monoidal cluster associated with a quantum seed $\mathcal {S}=( \{X_k\}_{k \in \mathsf {K}}, L, B)$. Then $[X]$ is $\mathcal {T}(L)$-pointed and $\mathcal {T}(L)$-copointed for any simple module $X \in \mathscr {C}_w$.

Remark 5.15 For a monoidal cluster $\mathcal {M}$ associated with a quantum seed $\mathcal {S}$ and a simple module $M \in \mathscr {C}$, the above corollary says that $\mathbf {g}^R_\mathcal {M}(M)$ and $\mathbf {g}^L_\mathcal {M}(M)$ coincide with the degree and codegree of $[M] \in \mathcal {K}_\mathbb {A}(\mathscr {C}) \simeq \mathscr {A}_\mathbb {A}(\mathcal {S})$, respectively.

Acknowledgements

The second, third and fourth authors gratefully acknowledge the hospitality of RIMS (Kyoto University) during their visit in 2023. The authors would like to thank the referee for reading our manuscript carefully and making constructive comments.

Conflicts of interest

None.

Financial support

The research of M. Kashiwara was supported by Grant-in-Aid for Scientific Research (B) 23K20206, Japan Society for the Promotion of Science. The research of M. Kim was supported by the National Research Foundation of Korea (NRF) Grant funded by the Korea government (MSIT) (NRF-2022R1F1A1076214 and NRF-2020R1A5A1016126). The research of S.-j. Oh was supported by the National Research Foundation of Korea (NRF) Grant funded by the Korea government (MSIT) (NRF-2022R1A2C1004045). The research of E. Park was supported by the National Research Foundation of Korea (NRF) Grant funded by the Korea Government (MSIT) (RS-2023-00273425 and NRF-2020R1A5A1016126).

Journal information

Compositio Mathematica is owned by the Foundation Compositio Mathematica and published by the London Mathematical Society in partnership with Cambridge University Press. All surplus income from the publication of Compositio Mathematica is returned to mathematics and higher education through the charitable activities of the Foundation, the London Mathematical Society and Cambridge University Press.

References

Berenstein, A. and Zelevinsky, A., Quantum cluster algebras, Adv. Math. 195 (2005), 405455; MR 2146350.CrossRefGoogle Scholar
Davison, B., Positivity for quantum cluster algebras, Ann. of Math. (2) 187 (2018), 157219.CrossRefGoogle Scholar
Derksen, H., Weyman, J. and Zelevinsky, A., Quivers with potentials and their representations II: applications to cluster algebras, J. Amer. Math. Soc. 23 (2010), 749790; MR 2629987.CrossRefGoogle Scholar
Fomin, S. and Zelevinsky, A., Cluster algebras. I. Foundations, J. Amer. Math. Soc. 15 (2002), 497529; MR 1887642.CrossRefGoogle Scholar
Fomin, S. and Zelevinsky, A., Cluster algebras. IV. Coefficients, Compos. Math. 143 (2007), 112164; MR 2295199.CrossRefGoogle Scholar
Fujita, R. and Oh, S.-j., Q-data and representation theory of untwisted quantum affine algebras, Comm. Math. Phys. 384 (2021), 13511407; MR 4259388.CrossRefGoogle Scholar
Geiss, C., Leclerc, B. and Schröer, J., Cluster structures on quantum coordinate rings, Selecta Math. (N.S.) 19 (2013), 337397; MR 3090232.CrossRefGoogle Scholar
Geiss, C., Leclerc, B. and Schröer, J., Factorial cluster algebras, Doc. Math. 18 (2013), 249274.CrossRefGoogle Scholar
Goodearl, K. R. and Yakimov, M. T., Quantum cluster algebra structures on quantum nilpotent algebras, Mem. Amer. Math. Soc. 247 (2017); MR 3633289.Google Scholar
Gross, M., Hacking, P., Keel, S. and Kontsevich, M., Canonical bases for cluster algebras, J. Amer. Math. Soc. 31 (2018), 497608; MR 3758151.CrossRefGoogle Scholar
Hernandez, D. and Leclerc, B., Cluster algebras and quantum affine algebras, Duke Math. J. 154 (2010), 265341; MR 2682185.CrossRefGoogle Scholar
Hernandez, D. and Leclerc, B., Quantum Grothendieck rings and derived Hall algebras, J. Reine Angew. Math. 701 (2015), 77126; MR 3331727.CrossRefGoogle Scholar
Kang, S.-J., Kashiwara, M. and Kim, M., Symmetric quiver Hecke algebras and $R$-matrices of quantum affine algebras, Invent. Math. 211 (2018), 591685.CrossRefGoogle Scholar
Kang, S.-J., Kashiwara, M., Kim, M. and Oh, S.-j., Simplicity of heads and socles of tensor products, Compos. Math. 151 (2015), 377396.CrossRefGoogle Scholar
Kang, S.-J., Kashiwara, M., Kim, M. and Oh, S.-j., Monoidal categorification of cluster algebras, J. Amer. Math. Soc. 31 (2018), 349426; MR 3758148.CrossRefGoogle Scholar
Kashiwara, M. and Kim, M., Laurent phenomenon and simple modules of quiver Hecke algebras, Compos. Math. 155 (2019), 22632295; MR 4016058.CrossRefGoogle Scholar
Kashiwara, M., Kim, M., Oh, S.-j. and Park, E., Monoidal categories associated with strata of flag manifolds, Adv. Math. 328 (2018), 9591009.CrossRefGoogle Scholar
Kashiwara, M., Kim, M., Oh, S.-j. and Park, E., Localizations for quiver Hecke algebras, Pure Appl. Math. Q. 17 (2021), 14651548.CrossRefGoogle Scholar
Kashiwara, M., Kim, M., Oh, S.-j. and Park, E., Localizations for quiver hecke algebras II, Proc. Lond. Math. Soc. 127 (2023), 11341184.CrossRefGoogle Scholar
Kashiwara, M., Kim, M., Oh, S.-j. and Park, E., Affinizations, $R$-matrices and reflection functors, Adv. Math. 443 (2024), 109598.CrossRefGoogle Scholar
Kashiwara, M., Kim, M., Oh, S.-j. and Park, E., Monoidal categorification and quantum affine algebras II, Invent. Math. 236 (2024), 837924.CrossRefGoogle Scholar
Kashiwara, M. and Oh, S.-j., The $(q, t)$-Cartan matrix specialized at $q= 1$ and its applications, Math. Z. 303 (2023), article 42.CrossRefGoogle Scholar
Kashiwara, M. and Park, E., Affinizations and $R$-matrices for quiver Hecke algebras, J. Eur. Math. Soc. (JEMS) 20 (2018), 11611193.CrossRefGoogle Scholar
Khovanov, M. and Lauda, A. D., A diagrammatic approach to categorification of quantum groups $I$, Represent. Theory 13 (2009), 309347.CrossRefGoogle Scholar
Khovanov, M. and Lauda, A. D., A diagrammatic approach to categorification of quantum groups $II$, Trans. Amer. Math. Soc. 363 (2011), 26852700.CrossRefGoogle Scholar
Kimura, Y., Quantum unipotent subgroup and dual canonical basis, Kyoto J. Math. 52 (2012), 277331; MR 2914878.CrossRefGoogle Scholar
Kimura, Y. and Oya, H., Twist automorphisms on quantum unipotent cells and dual canonical bases, Int. Math. Res. Not. IMRN 2021 (2021), 67726847.CrossRefGoogle Scholar
Lee, K. and Schiffler, R., Positivity for cluster algebras, Ann. of Math. (2) 182 (2015), 73125.CrossRefGoogle Scholar
McNamara, P. J., Finite dimensional representations of Khovanov–Lauda–Rouquier algebras $I$: finite type, J. Reine Angew. Math. 707 (2015), 103124.CrossRefGoogle Scholar
Qin, F., Dual canonical bases and quantum cluster algebras, Preprint (2020), arXiv:2003.13674.Google Scholar
Qin, F., Triangular bases in quantum cluster algebras and monoidal categorification conjectures, Duke Math. J. 166 (2017), 23372442; MR 3694569.CrossRefGoogle Scholar
Rouquier, R., 2-Kac-Moody algebras, Preprint (2008), arXiv:0812.5023.Google Scholar
Tingley, P. and Webster, B., Mirković–Vilonen polytopes and Khovanov–Lauda–Rouquier algebras, Compos. Math. 152 (2016), 16481696.CrossRefGoogle Scholar
Tran, T., $F$-polynomials in quantum cluster algebras, Algebr. Represent. Theory 14 (2011), 10251061; MR 2844755.CrossRefGoogle Scholar