Hostname: page-component-745bb68f8f-b95js Total loading time: 0 Render date: 2025-02-10T06:30:11.833Z Has data issue: false hasContentIssue false

Simply laced root systems arising from quantum affine algebras

Published online by Cambridge University Press:  08 February 2022

Masaki Kashiwara
Affiliation:
Kyoto University Institute for Advanced Study, Research Institute for Mathematical Sciences, Kyoto University, Kyoto 606-8502, Japan masaki@kurims.kyoto-u.ac.jp Korea Institute for Advanced Study, Seoul 02455, Korea
Myungho Kim
Affiliation:
Department of Mathematics, Kyung Hee University, Seoul 02447, Korea mkim@khu.ac.kr
Se-jin Oh
Affiliation:
Department of Mathematics, Ewha Womans University, Seoul 03760, Korea sejin092@gmail.com
Euiyong Park
Affiliation:
Department of Mathematics, University of Seoul, Seoul 02504, Korea epark@uos.ac.kr
Rights & Permissions [Opens in a new window]

Abstract

Let $U_q'({\mathfrak {g}})$ be a quantum affine algebra with an indeterminate $q$, and let $\mathscr {C}_{\mathfrak {g}}$ be the category of finite-dimensional integrable $U_q'({\mathfrak {g}})$-modules. We write $\mathscr {C}_{\mathfrak {g}}^0$ for the monoidal subcategory of $\mathscr {C}_{\mathfrak {g}}$ introduced by Hernandez and Leclerc. In this paper, we associate a simply laced finite-type root system to each quantum affine algebra $U_q'({\mathfrak {g}})$ in a natural way and show that the block decompositions of $\mathscr {C}_{\mathfrak {g}}$ and $\mathscr {C}_{\mathfrak {g}}^0$ are parameterized by the lattices associated with the root system. We first define a certain abelian group $\mathcal {W}$ (respectively $\mathcal {W} _0$) arising from simple modules of $\mathscr {C}_{\mathfrak {g}}$ (respectively $\mathscr {C}_{\mathfrak {g}}^0$) by using the invariant $\Lambda ^\infty$ introduced in previous work by the authors. The groups $\mathcal {W}$ and $\mathcal {W} _0$ have subsets $\Delta$ and $\Delta _0$ determined by the fundamental representations in $\mathscr {C}_{\mathfrak {g}}$ and $\mathscr {C}_{\mathfrak {g}}^0$, respectively. We prove that the pair $( \mathbb {R} \otimes _\mathbb {\mspace {1mu}Z\mspace {1mu}} \mathcal {W} _0, \Delta _0)$ is an irreducible simply laced root system of finite type and that the pair $( \mathbb {R} \otimes _\mathbb {\mspace {1mu}Z\mspace {1mu}} \mathcal {W} , \Delta )$ is isomorphic to the direct sum of infinite copies of $( \mathbb {R} \otimes _\mathbb {\mspace {1mu}Z\mspace {1mu}} \mathcal {W} _0, \Delta _0)$ as a root system.

Type
Research Article
Copyright
© 2022 The Author(s). The publishing rights in this article are licensed to Foundation Compositio Mathematica under an exclusive licence

1. Introduction

Let $q$ be an indeterminate and let $U_q'({\mathfrak {g}})$ be a quantum affine algebra. The category $\mathscr {C}_{\mathfrak {g}}$ of finite-dimensional integrable $U_q'({\mathfrak {g}})$-modules has a rich structure. For example, the category $\mathscr {C}_{\mathfrak {g}}$ is not semi-simple and has a rigid monoidal category structure. Because of its rich structure, it has been studied actively in various research areas of mathematics and physics (see [Reference Akasaka and KashiwaraAK97, Reference Chari and PressleyCP94, Reference Frenkel and ReshetikhinFR99, Reference KashiwaraKas02, Reference NakajimaNak01] for examples).

The category $\mathscr {C}_{\mathfrak {g}}$ has been studied from the viewpoint of cluster algebras. Suppose that ${\mathfrak {g}}$ is of simply laced affine ADE type. In [Reference Hernandez and LeclercHL10], Hernandez and Leclerc defined the full subcategory $\mathscr {C}_{\mathfrak {g}}^0$ of $\mathscr {C}_{\mathfrak {g}}$ such that all simple subquotients of its objects are obtained via simple subquotients of tensor products of certain fundamental representations. They then introduced certain monoidal subcategories $\mathscr {C}_\ell$ ($\ell \in \mathbb {\mspace {1mu}Z\mspace {1mu}}_{>0 }$) and studied their Grothendieck rings using cluster algebras. As any simple module in $\mathscr {C}_{\mathfrak {g}}$ can be obtained from a tensor product of suitable parameter shifts of simple modules in $\mathscr {C}_{\mathfrak {g}}^0$, the category $\mathscr {C}_{\mathfrak {g}}^0$ has an essential position in $\mathscr {C}_{\mathfrak {g}}$. Note that an algorithm for computing $q$-characters of Kirillov–Reshetikhin modules for any untwisted quantum affine algebras was described in [Reference Hernandez and LeclercHL16], by studying the cluster algebra structure of the Grothendieck ring of the subcategory $\mathscr {C}_{\mathfrak {g}}^-$ of $\mathscr {C}_{\mathfrak {g}}^0$. On the other hand, Hernandez and Leclerc introduced another abelian monoidal subcategory ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ which categorifies the coordinate ring $\mathbb {C}[N]$ of the unipotent group associated with the finite-dimensional simple Lie algebra ${\mathfrak {g}}_0$ inside ${\mathfrak {g}}$ [Reference Hernandez and LeclercHL15]. For each Dynkin quiver $Q$, they defined an abelian subcategory ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ of $\mathscr {C}_{\mathfrak {g}}^0$ which contains some fundamental representations parameterized by the coordinates of vertices of the Auslander–Reiten quiver of $Q$, and proved that ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ is stable under taking tensor products and that its complexified Grothendieck ring $\mathbb {C} \otimes _\mathbb {\mspace {1mu}Z\mspace {1mu}} K({\mathscr {C}_{\mathfrak {g}}^{Q}})$ is isomorphic to the coordinate ring $\mathbb {C}[N]$. Moreover, under this isomorphism, the set of isomorphism classes of simple modules in ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ corresponds to the upper global basis of $\mathbb {C}[N]$.

The notion of the categories $\mathscr {C}_{\mathfrak {g}}^0$ and ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ has been extended to all untwisted and twisted quantum affine algebras [Reference Kang, Kashiwara, Kim and OhKKKO16, Reference Kashiwara and OhKO19, Reference Oh and ScrimshawOS19a, Reference Oh and SuhOS19b]. Let ${\sigma ({\mathfrak {g}})} \mathbin {:=} I_0 \times \mathbf {k}^\times / \sim,$ where the equivalence relation is given by $(i,x) \sim (j,y)$ if and only if $V(\varpi _i)_x \simeq V(\varpi _j)_y$. The set ${\sigma ({\mathfrak {g}})}$ has a quiver structure determined by the pole of R-matrices between tensor products of fundamental representations $V(\varpi _i)_x$ ($(i,x)\in {\sigma ({\mathfrak {g}})}$). Let ${\sigma _0({\mathfrak {g}})}$ be a connected component of ${\sigma ({\mathfrak {g}})}$. Then the category $\mathscr {C}_{\mathfrak {g}}^0$ is defined to be the smallest full subcategory of $\mathscr {C}_{\mathfrak {g}}$ that has the following properties:

  1. (a) $\mathscr {C}_{\mathfrak {g}}^0$ contains $V(\varpi _i)_x$ for all $(i,x) \in {\sigma _0({\mathfrak {g}})}$;

  2. (b) $\mathscr {C}_{\mathfrak {g}}^0$ is stable by taking subquotients, extensions and tensor products.

The subcategory ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ was introduced in [Reference Kang, Kashiwara, Kim and OhKKKO16] for twisted affine type $A^{(2)}$ and $D^{(2)}$, in [Reference Kashiwara and OhKO19] for untwisted affine types $B^{(1)}$ and $C^{(1)}$, and in [Reference Oh and ScrimshawOS19a, Reference Oh and SuhOS19b] for exceptional affine type. For a Dynkin quiver $Q$ of a certain type with additional data, a finite subset ${\sigma _Q({\mathfrak {g}})}$ of ${\sigma _0({\mathfrak {g}})}$ was determined. Then the category ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ is defined to be the smallest full subcategory of $\mathscr {C}_{\mathfrak {g}}^0$ for which the following hold:

  1. (a) ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ contains $\mathbf {1}$ and $V(\varpi _i)_x$ for all $(i,x) \in {\sigma _Q({\mathfrak {g}})}$;

  2. (b) ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ is stable by taking subquotients, extensions and tensor products

(see §§ 2.3 and 2.4 for more details).

We can summarize the results of this paper as follows:

  1. (i) we associate a simply laced root system to each quantum affine algebra $U_q'({\mathfrak {g}})$ in a natural way;

  2. (ii) we give the block decomposition of $\mathscr {C}_{\mathfrak {g}}$ parameterized by a lattice $\mathcal {W}$ associated with the root system.

Let $U_q'({\mathfrak {g}})$ be a quantum affine algebra of arbitrary type. We first consider certain subgroups $\mathcal {W}$ and $\mathcal {W} _0$ of the abelian group $\operatorname {Hom}({\sigma ({\mathfrak {g}})}, \mathbb {\mspace {1mu}Z\mspace {1mu}})$ arising from simple modules of $\mathscr {C}_{\mathfrak {g}}$ and $\mathscr {C}_{\mathfrak {g}}^0$, respectively (see (4.2)). The subgroups $\mathcal {W}$ and $\mathcal {W} _0$ have subsets $\Delta$ and $\Delta _0$ determined by the fundamental representations in $\mathscr {C}_{\mathfrak {g}}$ and $\mathscr {C}_{\mathfrak {g}}^0$, respectively. Let $\mathcal {E} \mathbin {:=} \mathbb {R} \otimes _\mathbb {\mspace {1mu}Z\mspace {1mu}} \mathcal {W}$ and $\mathcal {E} _0 \mathbin {:=} \mathbb {R} \otimes _\mathbb {\mspace {1mu}Z\mspace {1mu}} \mathcal {W} _0$. Let ${{\mathfrak {g}}_{\mathrm {fin}}}$ be the simply laced finite-type Lie algebra corresponding to the affine type of ${\mathfrak {g}}$ in table (4.5). When ${\mathfrak {g}}$ is of untwisted affine type ADE, ${{\mathfrak {g}}_{\mathrm {fin}}}$ coincides with ${\mathfrak {g}}_0$. We prove that the pair $( \mathcal {E} _0, \Delta _0)$ is the irreducible root system of the Lie algebra ${{\mathfrak {g}}_{\mathrm {fin}}}$ and the pair $( \mathcal {E} , \Delta )$ is isomorphic to the direct sum of infinite copies of $( \mathcal {E} _0, \Delta _0)$ as a root system (see Theorem 4.6 and Corollary 4.7). Interestingly enough, the quantum affine algebra $U_q'({\mathfrak {g}})$ and its Langlands dual $U_q'(^{L}{\mathfrak {g}})$, whose Cartan matrix is the transpose of that of ${\mathfrak {g}}$, yield the same simply laced root system. This coincidence can also be viewed in terms of the mysterious duality between $U_q'({\mathfrak {g}})$ and its Langlands dual $U_q'(^{L}{\mathfrak {g}})$ (see [Reference Frenkel and HernandezFH11a, Reference Frenkel and HernandezFH11b, Reference Frenkel and ReshetikhinFR98]). We conjecture that the categories of representations of two quantum affine algebras are equivalent if and only if their associated root systems are the same. From this viewpoint, the simply laced finite-type root system plays the role of an invariant for the representation categories of quantum affine algebras. For each simply laced finite-type root system, the corresponding untwisted quantum affine algebra, the one of twisted type (if it exists) and its Langland dual have the same categorical structure.

We then show that there exist direct decompositions of $\mathscr {C}_{\mathfrak {g}}$ and $\mathscr {C}_{\mathfrak {g}}^0$ parameterized by elements of $\mathcal {W}$ and $\mathcal {W} _0$, respectively (Theorem 5.10), and we prove that each direct summand of the decompositions is a block (Theorem 5.14). This approach covers all untwisted and twisted quantum affine algebras in a uniform way and provides a transparent explanation of how the blocks of $\mathscr {C}_{\mathfrak {g}}^0$ exist from the perspective of the root system $( \mathcal {E} _0, \Delta _0)$ and the category ${\mathscr {C}_{\mathfrak {g}}^{Q}}$.

When ${\mathfrak {g}}$ is of untwisted type, the block decomposition was studied in [Reference Chari and MouraCM05, Reference Etingof and MouraEM03, Reference Jakelić and MouraJM11]. Etingof and Moura [Reference Etingof and MouraEM03] found the block decomposition of $\mathscr {C}_{\mathfrak {g}}$ whose blocks are parameterized by the elliptic central characters under the condition $|q|<1$. Later, Chari and Moura [Reference Chari and MouraCM05] gave a different description of the block decomposition of $\mathscr {C}_{\mathfrak {g}}$ by using the quotient group $\mathscr {P}_q / \mathscr {Q}_q$ of the $\ell$-weight lattice $\mathscr {P}_q$ by the $\ell$-root lattice $\mathscr {Q}_q$. In the case of the quantum affine algebra $U_\xi ({{\mathfrak {g}}})$ at roots of unity, its block decomposition was studied in [Reference Jakelić and MouraJM11]. For affine Kac–Moody algebras, the block decomposition of the category of finite-dimensional modules was studied in [Reference Chari and MouraCM04, Reference SenesiSen10]. Note that the block decomposition for affine Kac–Moody algebras does not explain blocks for quantum affine algebras $U'_q({\mathfrak {g}})$. We remark that in the untwisted-type case, the quotient group $\mathscr {P}_q / \mathscr {Q}_q$ given in [Reference Chari and MouraCM05] (and also the result of [Reference Etingof and MouraEM03]) provides another group presentation of $\mathcal {W}$ (see Remark 5.16).

The main tools used to prove our results are the new invariants $\Lambda$, $\Lambda ^\infty$ and $\operatorname {\mathfrak {d}}$ for a pair of modules in $\mathscr {C}_{\mathfrak {g}}$ introduced in [Reference Kashiwara, Kim, Oh and ParkKKOP20]. For non-zero modules $M$ and $N$ in $\mathscr {C}_{\mathfrak {g}}$ such that $R^{\mspace {1mu}\mathrm {univ}}_{M,N_z}$ is rationally renormalizable, the integers $\Lambda (M,N)$, $\Lambda ^\infty (M,N)$ and $\operatorname {\mathfrak {d}}(M,N)$ are defined by using the renormalizing coefficient $c_{M,N}(z)$ (see § 3 for details). These invariants are quantum affine algebra analogues of the invariants (with the same notation) for pairs of graded modules over quiver Hecke algebras arising from the grading of R-matrices. The new invariants play similar roles in the representation theory of quantum affine algebras to those for quiver Hecke algebras.

Let us explain our results more precisely. Let $U_q'({\mathfrak {g}})$ be a quantum affine algebra of arbitrary type. For $M \in \mathscr {C}_{\mathfrak {g}}$ such that the universal R-matrix $R^{\mspace {1mu}\mathrm {univ}}_{M, V(\varpi _i)_z}$ is rationally renormalizable for any $i\in I_0$, we define $\mathsf {E} (M) \in \operatorname {Hom}({\sigma ({\mathfrak {g}})}, \mathbb {\mspace {1mu}Z\mspace {1mu}})$ by

\[ \mathsf{E} (M) (i,a) := \Lambda^\infty(M, V(\varpi_i)_a) \quad \text{for } (i,a) \in {\sigma({\mathfrak{g}})} \]

and investigate its properties (Lemma 4.1). For $(i,a) \in {\sigma ({\mathfrak {g}})}$, we set

\[ \mathsf{s} _{i,a} \mathbin{:=} \mathsf{E} (V(\varpi_i)_a) \in \operatorname{Hom}({\sigma({\mathfrak{g}})}, \mathbb{\mspace{1mu}Z\mspace{1mu}}) \]

and

\begin{alignat*}{4} \mathcal{W} & \mathbin{:=} \{ \mathsf{E} (M) \mid M \text{ is simple in } \mathscr{C}_{\mathfrak{g}}\}, & \quad \Delta & \mathbin{:=} \{ \mathsf{s} _{i,a} \mid (i,a) \in {\sigma({\mathfrak{g}})} \} \subset \mathcal{W} , \\ \mathcal{W} _0 & \mathbin{:=} \{ \mathsf{E} (M) \mid M \text{ is simple in } \mathscr{C}_{\mathfrak{g}}^0 \}, & \quad \Delta_0 & \mathbin{:=} \{ \mathsf{s} _{i,a} \mid (i,a) \in {\sigma_0({\mathfrak{g}})} \} \subset \mathcal{W} _0. \end{alignat*}

Then $\mathcal {W}$ and $\mathcal {W} _0$ are abelian subgroups of $\operatorname {Hom}({\sigma ({\mathfrak {g}})}, \mathbb {\mspace {1mu}Z\mspace {1mu}})$. Moreover, we see in Lemma 4.2 that there exists a unique symmetric bilinear form $(-,-)$ on $\mathcal {W}$ such that $( \mathsf {E} (M), \mathsf {E} (N) ) = - \Lambda ^\infty (M,N)$ for any simple modules $M,N \in \mathscr {C}_{\mathfrak {g}}$; it induces a symmetric bilinear form on $\mathcal {E}$. Then we prove that the pair $( \mathcal {E} _0, \Delta _0)$ is an irreducible root system of the simply laced finite type given in (4.5) (Theorem 4.6) and that the pair $( \mathcal {E} , \Delta )$ is isomorphic to the direct sum of infinite copies of $( \mathcal {E} _0, \Delta _0)$ as a root system (Corollary 4.7). Furthermore, the bilinear form $(-,-)$ is invariant under the Weyl group action. Theorem 4.6 is proved in § 6 via a case-by-case approach, using the explicit descriptions of ${\sigma _Q({\mathfrak {g}})}$ for ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ given in § 2.4 and the denominator formulas in Appendix A.

We then consider the block decompositions of $\mathscr {C}_{\mathfrak {g}}$ and $\mathscr {C}_g^0$. For $\alpha \in \mathcal {W}$, let $\mathscr {C}_{{\mathfrak {g}}, \alpha }$ be the full subcategory of $\mathscr {C}_{\mathfrak {g}}$ consisting of objects $X$ such that $\mathsf {E} (S)=\alpha$ for any simple subquotient $S$ of $X$. We show that there exist the direct decompositions

\[ \mathscr{C}_{\mathfrak{g}} = \bigoplus_{\alpha \in \mathcal{W} } \mathscr{C}_{{\mathfrak{g}}, \alpha} \quad \text{and} \quad \mathscr{C}_{\mathfrak{g}}^0 = \bigoplus_{\alpha \in \mathcal{W} _0} \mathscr{C}_{{\mathfrak{g}}, \alpha} \]

by proving that $\operatorname {Ext}^1_{U_q'({\mathfrak {g}})}(M,N)=0$ for $M \in \mathscr {C}_{{\mathfrak {g}}, \alpha }$ and $N \in \mathscr {C}_{{\mathfrak {g}}, \beta }$ with $\alpha \ne \beta$ (Theorem 5.10). We set $\mathcal {P} \mathbin {:=} \bigoplus _{(i,a) \in {\sigma ({\mathfrak {g}})}} \mathbb {\mspace {1mu}Z\mspace {1mu}} \mathsf {e}_{(i,a)}$ and $\mathcal {P}_0 \mathbin {:=} \bigoplus _{(i,a) \in {\sigma _0({\mathfrak {g}})}} \mathbb {\mspace {1mu}Z\mspace {1mu}} \mathsf {e}_{(i,a)}$, where $\mathsf {e}_{(i,a)}$ is a symbol. Then we define a group homomorphism $p\colon \mathcal {P} \twoheadrightarrow \mathcal {W}$ by $p( \mathsf {e}_{(i,a)} )= \mathsf {s} _{i,a}$ and set $p_0 \mathbin {:=} p|_{\mathcal {P}_0}\colon \mathcal {P}_0 \twoheadrightarrow \mathcal {W} _0.$ It turns out that the kernel $\ker p_0$ coincides with the subgroup $\mathcal {Q}_0$ of $\mathcal {P}_0$ generated by elements of the form $\sum _{k=1}^m \mathsf {e}_{(i_k,a_k)}$ ($(i_k,a_k) \in {\sigma _0({\mathfrak {g}})}$) such that the trivial module $\mathbf {1}$ appears in $V(i_1)_{a_1} \otimes \cdots \otimes V(i_m)_{a_m}$ as a simple subquotient (Lemma 5.13). We then prove that $\mathscr {C}_{{\mathfrak {g}}, \alpha }$ is a block for any $\alpha \in \mathcal {W}$ (Theorem 5.14), which implies that the above decompositions are block decompositions of $\mathscr {C}_{\mathfrak {g}}$ and $\mathscr {C}_{\mathfrak {g}}^0$.

This paper is organized as follows. In § 2, we give the necessary background on quantum affine algebras, R-matrices, and the categories $\mathscr {C}_{\mathfrak {g}}$ and ${\mathscr {C}_{\mathfrak {g}}^{Q}}$. In § 3, we review the new invariants introduced in [Reference Kashiwara, Kim, Oh and ParkKKOP20]. In § 4, we investigate properties of $\mathcal {W}$, $\Delta$ and $\mathsf {s} _{i,a}$ and state the main theorem for the root systems $( \mathcal {E} _0, \Delta _0)$ and $( \mathcal {E} , \Delta )$. In § 5, we prove the block decompositions of $\mathscr {C}_{\mathfrak {g}}$ and $\mathscr {C}_{\mathfrak {g}}^0$. Section 6 is devoted to a case-by-case proof of Theorem 4.6.

2. Preliminaries

Convention

  1. (i) For a statement $P$, $\delta (P)$ is $1$ or $0$ according to whether $P$ is true or not.

  2. (ii) For an element $a$ in a field $\mathbf {k}$ and $f(z)\in \mathbf {k}(z)$, we denote by $\mspace {1mu}\mathrm {zero}\mspace {1mu}_{z=a}f(z)$ the order of zero of $f(z)$ at $z=a$.

2.1 Quantum affine algebras

The quintuple $( \mathsf {A}, \mathsf {P}, \Pi, \mathsf {P}^\vee, \Pi ^\vee )$ is called an affine Cartan datum if it consists of the following components:

  1. (i) an affine Cartan matrix $\mathsf {A}=(a_{ij})_{i,j\in I}$ with a finite index set $I$;

  2. (ii) a free abelian group $\mathsf {P}$ of rank $|I|+1$, called the weight lattice;

  3. (iii) a set $\Pi =\{\alpha _i \in \mathsf {P} \mid i \in I\}$, whose elements are called simple roots;

  4. (iv) the group $\mathsf {P}^\vee \mathbin {:=} \operatorname {Hom}_\mathbb {\mspace {1mu}Z\mspace {1mu}}(\mathsf {P},\mathbb {\mspace {1mu}Z\mspace {1mu}})$, called the coweight lattice;

  5. (v) a set $\Pi ^\vee =\{h_i \mid i \in I\} \subset \mathsf {P}^\vee$, whose elements are simple coroots;

and if it satisfies the following properties:

  1. (a) $\langle h_i, \alpha _j \rangle = a_{i,j}$ for any $i,j\in I$;

  2. (b) for any $i\in I$, there exists $\Lambda _i\in \mathsf {P}$ such that $\langle {h_j,\Lambda _i}\rangle =\delta (i=j)$ for any $j\in I$;

  3. (c) $\Pi$ is linearly independent.

Let ${\mathfrak {g}}$ be the affine Kac–Moody algebra associated with $(\mathsf {A},\mathsf {P},\Pi,\mathsf {P}^\vee,\Pi ^\vee )$. We set $\mathsf {Q} \mathbin {:=} \bigoplus _{i \in I}\mathbb {\mspace {1mu}Z\mspace {1mu}}\alpha _i \subset \mathsf {P}$, which is called the root lattice, and $\mathsf {Q}^+ \mathbin {:=} \sum _{i \in I}\mathbb {\mspace {1mu}Z\mspace {1mu}}_{\geqslant 0}\alpha _i\subset \mathsf {Q}$. For $\beta = \sum _{i\in I} b_i \alpha _i \in \mathsf {Q}^+$, we write $|\beta | = \sum _{i\in I} b_i$. We denote by $\delta \in \mathsf {Q}$ the imaginary root and by $c \in \mathsf {Q}^\vee$ the central element. Note that the positive imaginary root $\Delta _+^{\rm im}$ is equal to $\mathbb {\mspace {1mu}Z\mspace {1mu}}_{>0}\, \delta$ and the center of ${\mathfrak {g}}$ is generated by $c$. We write $\mathsf {P}_{\mathrm {cl}\mspace {1mu}} \mathbin {:=} \mathsf {P} / (\mathsf {P}\cap \mathbb {Q} \delta )$, which is called the classical weight lattice, and take $\rho \in \mathsf {P}$ (respectively $\rho ^\vee \in \mathsf {P}^\vee$) such that $\langle h_i,\rho \rangle =1$ (respectively $\langle \rho ^\vee,\alpha _i\rangle =1$) for any $i \in I$. There exists a $\mathbb {Q}$-valued non-degenerate symmetric bilinear form $( \ , \ )$ on $\mathsf {P}$ satisfying

\[ \langle h_i,\lambda \rangle= \frac{2(\alpha_i,\lambda)}{(\alpha_i,\alpha_i)} \quad \text{and} \quad \langle c,\lambda \rangle = (\delta,\lambda) \]

for any $i \in I$ and $\lambda \in \mathsf {P}$. We write $\mathsf {W} \mathbin {:=} \langle s_i \mid i \in I \rangle \subset \textrm {Aut}(\mathsf {P})$ for the Weyl group of $\mathsf {A}$, where $s_i (\lambda ) \mathbin {:=} \lambda - \langle h_i,\lambda \rangle \alpha _i$ for $\lambda \in \mathsf {P}$. We will use the standard convention in [Reference KacKac90] of choosing $0\in I$ except for type $A^{(2)}_{2n}$, in which case we take the longest simple root to be $\alpha _0$, and for types $B_2^{(1)}$, $A_3^{(2)}$ and $E_k^{(1)}$ ($k=6,7,8$), in which cases we take the following Dynkin diagrams.

(2.1)

Note that $B^{(1)}_2$ and $A^{(2)}_3$ in (

2.1

) are denoted by $C_2^{(1)}$ and $D_3^{(2)}$, respectively, in [Reference KacKac90].

Let ${\mathfrak {g}}_0$ be the subalgebra of ${\mathfrak {g}}$ generated by the Chevalley generators $e_i$, $f_i$ and $h_i$ for $i \in I_0 \mathbin {:=} I \setminus \{ 0 \}$, and let $\mathsf {W}_0$ be the subgroup of $\mathsf {W}$ generated by $s_i$ for $i \in I_0$. Note that ${\mathfrak {g}}_0$ is a finite-dimensional simple Lie algebra and $\mathsf {W}_0$ contains the longest element $w_0$.

Let $q$ be an indeterminate and ${{\mathbf {k}}}$ the algebraic closure of the subfield $\mathbb {C}(q)$ in the algebraically closed field $\hat {\mathbf {k}}\mathbin {:=}\bigcup _{m >0}\mathbb {C}((q^{1/m}))$. For $m,n \in \mathbb {\mspace {1mu}Z\mspace {1mu}}_{\geqslant 0}$ and $i\in I$, we define $q_i = q^{(\alpha _i,\alpha _i)/2}$ and

\[ [n]_i =\frac{ q^n_{i} - q^{{-}n}_{i} }{ q_{i} - q^{{-}1}_{i} }, \quad [n]_i! = \prod^{n}_{k=1} [k]_i , \quad \Big[\begin{matrix}m \\ n\\ \end{matrix} \Big]_i= \frac{ [m]_i! }{[m-n]_i! [n]_i! }. \]

Let $d$ be the smallest positive integer such that $d ({(\alpha _i, \alpha _i)}/{2})\in \mathbb {\mspace {1mu}Z\mspace {1mu}}$ for all $i\in I$.

Definition 2.1 The quantum affine algebra $U_q({\mathfrak {g}})$ associated with an affine Cartan datum $(\mathsf {A},\mathsf {P},\Pi,\mathsf {P}^\vee,\Pi ^\vee )$ is the associative algebra over ${{\mathbf {k}}}$ with $1$ generated by $e_i,f_i$ $(i \in I)$ and $q^{h}$ $(h \in d^{-1} \mathsf {P}^{\vee })$ which satisfies the following relations:

  1. (i) $q^0=1$ and $q^{h} q^{h'}=q^{h+h'}$ for $h,h' \in d^{-1} \mathsf {P}^{\vee }$;

  2. (ii) $q^{h}e_i q^{-h}= q^{\langle h, \alpha _i \rangle } e_i$ and $q^{h}f_i q^{-h} = q^{-\langle h, \alpha _i \rangle }f_i$ for $h \in d^{-1}\mathsf {P}^{\vee }$ and $i \in I$;

  3. (iii) $e_if_j - f_je_i = \delta _{ij} (({K_i -K^{-1}_i})/({q_i- q^{-1}_i })), \text { where } K_i=q_i^{ h_i}$;

  4. (iv) $\sum ^{1-a_{ij}}_{k=0} (-1)^ke^{(1-a_{ij}-k)}_i e_j e^{(k)}_i = \sum ^{1-a_{ij}}_{k=0} (-1)^k f^{(1-a_{ij}-k)}_i f_jf^{(k)}_i=0 \text { for } i \ne j$;

here $e_i^{(k)}=e_i^k/[k]_i!$ and $f_i^{(k)}=f_i^k/[k]_i!$.

Let us denote by $U_q'({\mathfrak {g}})$ the $\mathbf {k}$-subalgebra of $U_q({\mathfrak {g}})$ generated by $e_i,f_i$ and $K^{\pm 1}_i$ $(i \in I)$. The coproduct $\Delta$ of $U_q'({\mathfrak {g}})$ is given by

\[ \Delta(q^h)=q^h \mathop\otimes q^h, \quad \Delta(e_i)=e_i \mathop\otimes K_i^{{-}1}+1 \mathop\otimes e_i, \quad \Delta(f_i)=f_i \mathop\otimes 1 +K_i \mathop\otimes f_i, \]

and the bar involution $\bar{}$ of $U_q'({\mathfrak {g}})$ is defined as

\[ q^{1/m} \to q^{{-}1/m}, \quad e_i \mapsto e_i, \quad f_i \mapsto f_i, \quad K_i \mapsto K_i^{{-}1}. \]

Let $\mathscr {C}_{\mathfrak {g}}$ be the category of finite-dimensional integrable $U_q'({\mathfrak {g}})$-modules, i.e. finite-dimensional modules $M$ with a weight decomposition

\[ M=\bigoplus_{\lambda\in\mathsf{P}_{\mathrm{cl}\mspace{1mu}}}M_\lambda \quad \text{where } M_\lambda=\{{u\in M\mid K_iu=q_i^{\langle{h_i,\lambda}\rangle}u }\}. \]

Note that the trivial module $\mathbf {1}$ is contained in $\mathscr {C}_{\mathfrak {g}}$ and the tensor product $\otimes$ gives a monoidal category structure on $\mathscr {C}_{\mathfrak {g}}$. It is known that the Grothendieck ring $K(\mathscr {C}_{\mathfrak {g}})$ is a commutative ring [Reference Frenkel and ReshetikhinFR99]. A simple module $L$ in $\mathscr {C}_{\mathfrak {g}}$ contains a non-zero vector $u \in L$ of weight $\lambda \in \mathsf {P}_{\mathrm {cl}\mspace {1mu}}$ such that (i) $\langle h_i,\lambda \rangle \geqslant 0$ for all $i \in I_0$ and (ii) all the weights of $L$ are contained in $\lambda - \sum _{i \in I_0} \mathbb {\mspace {1mu}Z\mspace {1mu}}_{\geqslant 0}\, {\mathrm {cl}\mspace {1mu}}(\alpha _i)$, where ${\mathrm {cl}\mspace {1mu}}\colon \mathsf {P}\to \mathsf {P}_{\mathrm {cl}\mspace {1mu}}$ is the canonical projection. Such a $\lambda$ is unique, and $u$ is unique up to a constant multiple. We call $\lambda$ the dominant extremal weight of $L$ and $u$ a dominant extremal weight vector of $L$.

Let $\mathsf {P}_{\mathrm {cl}\mspace {1mu}}^0\mathbin {:=}\{\lambda \in \mathsf {P}_{\mathrm {cl}\mspace {1mu}}\mid \langle {c,\lambda }\rangle =0\}$. For each $i \in I_0$ we set

\[ \varpi_i \mathbin{:=} {\rm gcd}(\mathsf{c}_0,\mathsf{c}_i)^{{-}1}{\mathrm{cl}\mspace{1mu}}(\mathsf{c}_0\Lambda_i-\mathsf{c}_i \Lambda_0) \in \mathsf{P}_{\mathrm{cl}\mspace{1mu}}^0, \]

where the central element $c$ is equal to $\sum _{ i\in I} \mathsf {c}_i h_i$. Note that $\mathsf {P}_{\mathrm {cl}\mspace {1mu}}^0= \bigoplus _{i\in I_0}\mathbb {\mspace {1mu}Z\mspace {1mu}}\varpi _i$. For any $i\in I_0$, there exists a unique simple module $V(\varpi _i)$ in $\mathscr {C}_{\mathfrak {g}}$ satisfying certain good conditions (see [Reference KashiwaraKas02, § 5.2]), which is called the $i$th fundamental representation. Note that the dominant extremal weight of $V(\varpi _i)$ is $\varpi _i$.

For simple modules $M$ and $N$ in $\mathscr {C}_{\mathfrak {g}}$, we say that $M$ and $N$ commute or $M$ commutes with $N$ if $M\mathop \otimes N\simeq N\mathop \otimes M$. We say that $M$ and $N$ strongly commute or $M$ strongly commutes with $N$ if $M\mathop \otimes N$ is simple. Note that $M\mathop \otimes N$ is simple if and only if $N\mathop \otimes M$ is simple, since $K(\mathscr {C}_{\mathfrak {g}})$ is a commutative ring. It is clear that if simple modules $M$ and $N$ strongly commute, then they commute. We say that a simple module $M$ is real if $M$ strongly commutes with itself.

For an integrable $U_q'({\mathfrak {g}})$-module $M$, we denote by $M_z$ the affinization of $M$ and by $z_M \colon M_z \to M_z$ the $U_q'({\mathfrak {g}})$-module automorphism of weight $\delta$. Note that $M_z \simeq \mathbf {k}[z^{\pm 1}]\otimes _{\mathbf {k}} M$ for an indeterminate $z$ as a $\mathbf {k}$-vector space. For $x \in {{\mathbf {k}}}^\times$, we define

\[ M_x \mathbin{:=} M_z / (z_M -x)M_z. \]

We call $x$ a spectral parameter. The functor $T_x$ defined by $T_x(M)=M_x$ is an endofunctor of $\mathscr {C}_{\mathfrak {g}}$ that commutes with tensor products (see [Reference KashiwaraKas02, § 4.2] for details).

It is known that a fundamental representation is a good module, which is a simple $U_q'({\mathfrak {g}})$-module with good properties including a bar involution, a crystal basis with simple crystal graph, and a global basis (see [Reference KashiwaraKas02] for the precise definition). We say that a $U_q'({\mathfrak {g}})$-module $M$ is quasi-good if

\[ M \simeq V_c \]

for some good module $V$ and $c \in \mathbf {k}^\times$. Note that any quasi-good module is a simple $U_q'({\mathfrak {g}})$-module. Moreover the tensor product $M^{\otimes k} \mathbin {:=} \underbrace {M \mathop \otimes \cdots \mathop \otimes M}_{k\text { times}}$ for a quasi-good module $M$ and $k \in \mathbb {\mspace {1mu}Z\mspace {1mu}}_{\geqslant 1}$ is again quasi-good.

For a $U_q'({\mathfrak {g}})$-module $M$, we denote by $\bar {M}=\{ \bar {u} \mid u \in M \}$ the $U_q'({\mathfrak {g}})$-module defined by $x \bar {u} \mathbin {:=} \overline { \bar {x} u }$ for $x \in U_q'({\mathfrak {g}})$. Then we have

\[ \overline{M_a} \simeq (\bar{M})_{\,\bar{a}} \quad \text{and} \quad \overline{M \otimes N} \simeq \bar{N} \otimes \bar{M}\quad\text{for any $M,N\in\mathscr{C}_{\mathfrak{g}}$ and $a\in\mathbf{k}^\times$.}\quad \]

Note that $V(\varpi _i)$ is bar-invariant, i.e. $\overline {V(\varpi _i)}\simeq V(\varpi _i)$ (see [Reference Akasaka and KashiwaraAK97, Appendix A]).

Let $m_i$ be a positive integer such that

\[ \mathsf{W}\pi_i\cap\bigl(\pi_i+\mathbb{\mspace{1mu}Z\mspace{1mu}}\delta\bigr)=\pi_i+\mathbb{\mspace{1mu}Z\mspace{1mu}} m_i\delta, \]

where $\pi _i$ is an element of $\mathsf {P}$ such that ${\mathrm {cl}\mspace {1mu}}(\pi _i)=\varpi _i$. Note that $m_i=(\alpha _i,\alpha _i)/2$ in the case where ${\mathfrak {g}}$ is the dual of an untwisted affine algebra, and $m_i=1$ otherwise. Then for $x,y\in {{\mathbf {k}}}^\times$ we have (see [Reference Akasaka and KashiwaraAK97, § 1.3])

(2.2)\begin{equation} V(\varpi_i)_x \simeq V(\varpi_i)_y \quad \text{if and only if} \quad \text{$x^{m_i}=y^{m_i}$}. \end{equation}

We set

(2.3)\begin{equation} {\sigma({\mathfrak{g}})} \mathbin{:=} I_0 \times \mathbf{k}^\times{/} \sim,\end{equation}

where the equivalence relation is given by

(2.4)\begin{equation} (i,x) \sim (j,y) \iff V(\varpi_i)_x \simeq V(\varpi_j)_y \iff \text{$i=j$ and $x^{m_i}=y^{m_j}$.} \end{equation}

We denote by $[(i,a)]$ the equivalence class of $(i,a)$ in ${\sigma ({\mathfrak {g}})}$. When confusion is unlikely to arise, we simply write $(i,a)$ for the equivalence class $[(i,a)]$.

The monoidal category $\mathscr {C}_{\mathfrak {g}}$ is rigid. For $M\in \mathscr {C}_{\mathfrak {g}}$, we denote by ${}^*\mspace {-3mu} M$ and $M^*$ the right and left duals of $M$, respectively. We set

(2.5)\begin{equation} p^* \mathbin{:=} ({-}1)^{ \langle \rho^\vee, \delta \rangle} q^{\langle c, \rho \rangle} \quad \text{and}\quad \tilde{p} \mathbin{:=} (p^*)^2 = q^{2 \langle c, \rho \rangle}. \end{equation}

The integer $\langle \rho ^\vee, \delta \rangle$ is called the Coxeter number, and $\langle c, \rho \rangle$ is called the dual Coxeter number (see [Reference KacKac90, Ch. 6]). For the reader's convenience we list $p^*$ for all types in the following table.

(2.6) \begin{align} \small \begin{array}{|c||c|c|c|c|c|c|c|} \hline \text{Type of ${\mathfrak{g}}$} & A_n^{(1)^{\vphantom{1}}} & B_n^{(1)} & C_n^{(1)} & D_n^{(1)} & A_{2n}^{(2)} & A_{2n-1}^{(2)} & D_{n+1}^{(2)} \\ & (n\geqslant1) & (n\geqslant2) & (n\geqslant3) & (n\geqslant4) & (n\geqslant1) & (n\geqslant2) & (n\geqslant3)\\ \hline p^* & ({-}q)^{n+1^{\vphantom{1}}} & q^{2n-1} & q^{n+1} & q^{2n-2} & -q^{2n+1} & -q^{2n} & ({-}1)^{n+1} q^{2n} \\ \hline \hline \text{Type of ${\mathfrak{g}}$} & E_6^{(1)} & E_7^{(1)} & E_8^{(1)^{\vphantom{1}}} & F_4^{(1)} & G_{2}^{(1)} & E_{6}^{(2)} & D_{4}^{(3)} \\ \hline p^* & q^{12} & q^{18^{\vphantom{1}}} & q^{30} & q^9 & q^4 & -q^{12} & q^6 \\ \hline\end{array} \end{align}

Then for any $M \in \mathscr {C}_{\mathfrak {g}}$ we have

\[ M^{**} \simeq M_{ (\tilde{p})^{{-}1} }\quad \text{and}\quad {}^{**} M \simeq M_{\tilde{p} }, \]

and for $i\in I_0$ and $x\in \mathbf {k}^\times$ we have

(2.7)\begin{equation} \bigl( V(\varpi_i)_x\bigr)^* \simeq V(\varpi_{i^*})_{(p^*)^{{-}1}x} \quad \text{and}\quad {}^*\bigl(V(\varpi_i)_x \bigr) \simeq V(\varpi_{i^*})_{p^*x}, \end{equation}

where $i^*\in I_0$ is defined by $\alpha _{i^*}=-w_0\,\alpha _i$ (see [Reference Akasaka and KashiwaraAK97, Appendix A]). Note that the involution $i \mapsto i^*$ is the identity for all types except $A_n, D_n$ and $E_6$, which are given as follows:

  1. (a) (type $A_n$) $i^* = n+1-i$;

  2. (b) (type $D_n$) $i^* = \begin {cases} n-(1-\epsilon ) & \text {if } n \hbox { is odd and } i=n-\epsilon\ (\epsilon =0,1), \\ i & \text {otherwise;} \end {cases}$

  3. (c) (type $E_6$) the map $i \mapsto i^*$ is determined by

    \[ i^* = \begin{cases} 6 & \text{if } i=1, \\ i & \text{if } i=2,4, \\ 5 & \text{if } i=3, \end{cases} \]
    where the Dynkin diagram of type $E_6$ is given in (A.3) in Appendix A.

2.2 R-matrices

We recall the notion of R-matrices on $U_q'({\mathfrak {g}})$-modules and their coefficients (see [Reference Drinfel'dDri86], as well as [Reference Akasaka and KashiwaraAK97, Appendices A and B] and [Reference KashiwaraKas02, § 8], for details). Choose a basis $\{P_\nu \}_\nu$ of $U_q^+({\mathfrak {g}})$ and a basis $\{Q_\nu \}_\nu$ of $U_q^-({\mathfrak {g}})$ that are dual to each other with respect to a suitable coupling between $U_q^+({\mathfrak {g}})$ and $U_q^-({\mathfrak {g}})$. For $U_q'({\mathfrak {g}})$-modules $M$ and $N$, we define

\[ R^{\mspace{1mu}\mathrm{univ}}_{M,N}(u\otimes v) \mathbin{:=} q^{(\operatorname{wt}(u),\operatorname{wt}(v))} \sum_\nu P_\nu v\otimes Q_\nu u \quad\text{for $u\in M$ and $v\in N$,}\quad \]

so that $R^{\mspace {1mu}\mathrm {univ}}_{M,N}$ gives a $U_q'({\mathfrak {g}})$-linear homomorphism $M\otimes N \rightarrow N\otimes M$, called the universal R-matrix, provided that the infinite sum has a meaning. As $R^{\mspace {1mu}\mathrm {univ}}_{M,N_z}$ converges in the $z$-adic topology for $M,N\in \mathscr {C}_{\mathfrak {g}}$, we have a morphism of ${{\mathbf {k}}}((z))\mathop \otimes U_q'({\mathfrak {g}})$-modules

\[ R^{\mspace{1mu}\mathrm{univ}}_{M,N_z} \colon {{\mathbf{k}}}((z))\mathop\otimes_{{{\mathbf{k}}}[z^{\pm1}]} (M \mathop\otimes N_z) \xrightarrow{\,{\hspace{2ex}}\,} {{\mathbf{k}}}((z))\mathop\otimes_{{{\mathbf{k}}}[z^{\pm1}]} (N_z\mathop\otimes M). \]

Note that $R^{\mspace {1mu}\mathrm {univ}}_{M,N_z}$ is an isomorphism.

Let $M$ and $N$ be non-zero modules in $\mathscr {C}_{\mathfrak {g}}$. The universal R-matrix $R^{\mspace {1mu}\mathrm {univ}}_{M,N_z}$ is rationally renormalizable if there exists $f(z) \in {{\mathbf {k}}}((z))^\times$ such that

\[ f(z) R^{\mspace{1mu}\mathrm{univ}}_{M,N_z}\bigl( M\mathop\otimes N_z\bigr)\subset N_z\mathop\otimes M. \]

In this case, we can choose $c_{M,N}(z) \in {{\mathbf {k}}}((z))^\times$ such that for any $x \in {{\mathbf {k}}}^\times$, the specialization of $R^{\mspace {1mu}\mathrm {ren}}_{M,N_z} \mathbin {:=} c_{M,N}(z)R^{\mspace {1mu}\mathrm {univ}}_{M,N_z} \colon M \otimes N_z \to N_z \otimes M$ at $z=x$,

\[ R^{\mspace{1mu}\mathrm{ren}}_{M,N_z}\big\vert_{z=x} \colon M \otimes N_x \to N_x \otimes M, \]

does not vanish. Note that $R^{\mspace {1mu}\mathrm {ren}}_{M,N_z}$ and $c_{M,N}(z)$ are unique up to a multiple of ${\mathbf {k}[z^{\pm 1}]}^\times = \bigsqcup _{n \in \mathbb {\mspace {1mu}Z\mspace {1mu}}}{{\mathbf {k}}}^\times z^n$. We call $c_{M,N}(z)$ the renormalizing coefficient. We denote by ${{\mathbf {r}}_{{\scriptstyle {M,N}}}}$ the specialization at $z=1$,

(2.8)\begin{equation} {{\mathbf{r}}_{\mspace{-2mu}{{\scriptstyle{M,N}}}}} \mathbin{:=} R^{\mspace{1mu}\mathrm{ren}}_{M,N_z}\vert_{z=1} \colon M \otimes N \to N \otimes M, \end{equation}

and call it the R-matrix. The R-matrix ${{\mathbf {r}}_ {{\scriptstyle {M,N}}}}$ is well-defined up to a constant multiple whenever $R^{\mspace {1mu}\mathrm {univ}}_{M,N_z}$ is rationally renormalizable. By the definition, ${{\mathbf {r}}_{{\scriptstyle {M,N}}}}$ never vanishes.

Suppose that $M$ and $N$ are simple $U_q'({\mathfrak {g}})$-modules in $\mathscr {C}_{\mathfrak {g}}$. Let $u$ and $v$ be dominant extremal weight vectors of $M$ and $N$, respectively. Then there exists $a_{M,N}(z)\in {{\mathbf {k}}}[[z]]^\times$ such that

\[ R^{\mspace{1mu}\mathrm{univ}}_{M,N_z}( u \mathop\otimes v_z)= a_{M,N}(z)( v_z\mathop\otimes u ). \]

Thus we have a unique ${{\mathbf {k}}}(z)\mathop \otimes U_q'({\mathfrak {g}})$-module isomorphism

\[ R^{\mathrm{norm}}_{M,N_z}\mathbin{:=} a_{M,N}(z)^{{-}1} R^{\mspace{1mu}\mathrm{univ}}_{M,N_z}\big \vert_{{{\mathbf{k}}}(z)\otimes_{{{\mathbf{k}}}[z^{\pm1}]} ( M \otimes N_z)} \]

from ${{\mathbf {k}}}(z)\otimes _{{{\mathbf {k}}}[z^{\pm 1}]} ( M \otimes N_z)$ to ${{\mathbf {k}}}(z)\otimes _{{{\mathbf {k}}}[z^{\pm 1}]} ( N_z \otimes M )$, which satisfies

\[ R^{\mathrm{norm}}_{M, N_z}( u \otimes v_z) = v_z\otimes u . \]

We call $a_{M,N}(z)$ the universal coefficient of $M$ and $N$, and call $R^{\mathrm {norm}}_{M,N_z}$ the normalized $R$-matrix.

Let $d_{M,N}(z) \in {{\mathbf {k}}}[z]$ be a monic polynomial of the smallest degree such that the image of $d_{M,N}(z) R^{\mathrm {norm}}_{M,N_z}(M\mathop \otimes N_z)$ is contained in $N_z \otimes M$; we call it the denominator of $R^{\mathrm {norm}}_{M,N_z}$. Then we have

\[ R^{\mspace{1mu}\mathrm{ren}}_{M,N_z} = d_{M,N}(z)R^{\mathrm{norm}}_{M,N_z} \colon M \otimes N_z \xrightarrow{\,{\hspace{2ex}}\,} N_z \otimes M \quad\text{up to a multiple of ${\mathbf{k}[z^{\pm1}]}^\times$.}\quad \]

Thus

\[ R^{\mspace{1mu}\mathrm{ren}}_{M,N_z} =a_{M,N}(z)^{{-}1}d_{M,N}(z)R^{\mspace{1mu}\mathrm{univ}}_{M,N_z} \quad \text{and} \quad c_{M,N}(z)= \frac{d_{M,N}(z)}{a_{M,N}(z)} \]

up to a multiple of ${{\mathbf {k}}}[z^{\pm 1}]^\times$. In particular, $R^{\mspace {1mu}\mathrm {univ}}_{M,N_z}$ is rationally renormalizable whenever $M$ and $N$ are simple.

The denominator formulas between fundamental representations are summarized for all types in Appendix A.

The next theorem follows from the results of [Reference Akasaka and KashiwaraAK97, Reference ChariCha10, Reference KashiwaraKas02, Reference Kang, Kashiwara, Kim and OhKKKO15]. In the theorem, (ii) follows essentially from [Reference Kang, Kashiwara, Kim and OhKKKO15, Corollary 3.16] together with properties of R-matrices (see also [Reference Kashiwara, Kim, Oh and ParkKKOP20, Proposition 3.16 and Corollary 3.17]), and (i), (iii) and (iv) were conjectured in [Reference Akasaka and KashiwaraAK97, § 2] and proved in [Reference Akasaka and KashiwaraAK97, § 4] for affine types $A$ and $C$, in [Reference KashiwaraKas02, § 9] for general cases in terms of good modules, and in [Reference ChariCha10, §§ 4 and 6] using the braid group actions.

Theorem 2.2 [Reference Akasaka and KashiwaraAK97, Reference ChariCha10, Reference KashiwaraKas02, Reference Kang, Kashiwara, Kim and OhKKKO15]

  1. (i) For good modules $M$ and $N$, the zeros of $d_{M,N}(z)$ belong to $\mathbb {C}[[q^{1/m}]]q^{1/m}$ for some $m\in \mathbb {\mspace {1mu}Z\mspace {1mu}}_{>0}$.

  2. (ii) For simple modules $M$ and $N$ such that one of them is real, $M_x$ and $N_y$ strongly commute with each other if and only if $d_{M,N}(z)d_{N,M}(1/z)$ does not vanish at $z=y/x$.

  3. (iii) Let $M_k$ be a good module with a dominant extremal vector $u_k$ of weight $\lambda _k$, and let $a_k\in {{\mathbf {k}}}^\times$ for $k=1,\ldots, t$. Assume that $a_j/a_i$ is not a zero of $d_{M_i, M_j}(z)$ for any $1\leqslant i< j\leqslant t$. Then the following statements hold.

    1. (a) $(M_1)_{a_1}\otimes \cdots \otimes (M_t)_{a_t}$ is generated by $u_1\otimes \cdots \otimes u_t$.

    2. (b) The head of $(M_1)_{a_1}\otimes \cdots \otimes (M_t)_{a_t}$ is simple.

    3. (c) Any non-zero submodule of $(M_t)_{a_t}\otimes \cdots \otimes (M_1)_{a_1}$ contains the vector $u_t\otimes \cdots \otimes u_1$.

    4. (d) The socle of $(M_t)_{a_t}\otimes \cdots \otimes (M_1)_{a_1}$ is simple.

    5. (e) Let ${{\mathbf {r}}_{\mspace {-2mu}{{\scriptstyle {}}}}}\colon (M_1)_{a_1}\!\otimes\! \cdots \!\otimes\! (M_t)_{a_t} \!\to\! (M_t)_{a_t}\otimes \cdots \otimes (M_1)_{a_1}$ be the specialization of ${{\mathbf {r}}_{\mspace {-2mu}{{\scriptstyle {M_1,\ldots, M_t}}}}}\mathbin {:=}\prod _{1\leqslant j< k\leqslant t}{{\mathbf {r}}_{\mspace {-2mu}{{\scriptstyle {M_j,M_k}}}}}$ at $z_k=a_k$; see (2.8). Then the image of ${\mathbf {r}}$ is simple and coincides with the head of $(M_1)_{a_1}\otimes \cdots \otimes (M_t)_{a_t}$ and also with the socle of $(M_t)_{a_t}\otimes \cdots \otimes (M_1)_{a_1}$.

  4. (iv) For any simple integrable $U_q'({\mathfrak {g}})$-module $M$, there exists a finite sequence in ${\sigma ({\mathfrak {g}})}$ (see (2.3)) such that $M$ has $\sum _{k=1}^t \varpi _{i_k}$ as a dominant extremal weight and is isomorphic to a simple subquotient of $V(\varpi _{i_1})_{a_1}\mathop \otimes \cdots V(\varpi _{i_t})_{a_t}$. Moreover, such a sequence $\bigl ((i_1,a_1),\ldots, (i_t,a_t)\bigr )$ is unique up to a permutation.

    We call $\sum _{k=1}^t(i_k,a_k)\in \mathbb {\mspace {1mu}Z\mspace {1mu}}^{\oplus {\sigma ({\mathfrak {g}})}}$ the affine highest weight of $M$.

2.3 Hernandez–Leclerc categories

Recall ${\sigma ({\mathfrak {g}})}$ in (2.3). For $(i,x)$ and $(j,y) \in {\sigma ({\mathfrak {g}})}$, we put $d$ arrows from $(i,x)$ to $(j,y)$, where $d$ is the order of the zeros of $d_{ V(\varpi _i), V(\varpi _j) } ( z_{V(\varpi _j)} / z_{V(\varpi _i)} )$ at $z_{V(\varpi _j)} / z_{V(\varpi _i)} = y/x$. Then ${\sigma ({\mathfrak {g}})}$ has a quiver structure. Note that $(i, x)$ and $(j,y)$ are linked in ${\sigma ({\mathfrak {g}})}$ if and only if the tensor product $V(\varpi _i)_x \otimes V(\varpi _j)_y$ is reducible [Reference Akasaka and KashiwaraAK97, Corollary 2.4]. The denominator formulas are explicitly given in Appendix A.

We choose a connected component ${\sigma _0({\mathfrak {g}})}$ of ${\sigma ({\mathfrak {g}})}$. Since a connected component of ${\sigma ({\mathfrak {g}})}$ is unique up to a spectral parameter shift, ${\sigma _0({\mathfrak {g}})}$ is uniquely determined up to a quiver isomorphism. We set

(2.9)\begin{equation} q_s=q^{1/2}\quad\text{{and}}\quad q_t=q^{1/3}. \end{equation}

The distance $\mathrm {d}(u,v)$ between two vertices $u$ and $v$ in a finite Dynkin diagram is the length of the path connecting them. For example, $\mathrm {d}(1,4)=2$ in a Dynkin diagram of type $D_4$, and $\mathrm {d}(1,3)=2$ in a Dynkin diagram of type $F_4$. We denote by $\mathrm {d}_\circ (i,j)$ the distance between $i$ and $j$ in the Dynkin diagram of ${\mathfrak {g}}_0$. For the rest of this paper, we make the following choices of $\sigma _0({\mathfrak {g}})$ (see table (2.6) for the range of $n$):

\begin{align*} &\sigma_0(X) \mathbin{:=} \{(i,({-}q)^{p}) \in I_0 \times {{\mathbf{k}}}^{{\times}} \mid p \equiv_2 \mathrm{d}_\circ(1,i) \} \quad ( X=A^{(1)}_{n}, D^{(1)}_{n}, E^{(1)}_{k} (k=6,7,8) ), \\ & \sigma_0({B_{n}^{(1)}}) \mathbin{:=} \{{ (i,({-}1)^{ n+i} q_s q^m), (n,q^m)}\mid{1\leqslant i \leqslant n-1,\: m \in \mathbb{\mspace{1mu}Z\mspace{1mu}}}\}, \\ &\sigma_0(C^{(1)}_{n}) \mathbin{:=} \{(i,({-}q_s)^{p}) \in I_0 \times {{\mathbf{k}}}^{{\times}} \mid p \equiv_2 \mathrm{d}_\circ(1,i) \}, \\ &\sigma_0(F^{(1)}_{4}) \mathbin{:=} \{(i,({-}1)^iq_s^{2p-\delta_{i,3}}) \in I_0 \times {{\mathbf{k}}}^{{\times}} \mid p \in \mathbb{\mspace{1mu}Z\mspace{1mu}} \}, \\ &\sigma_0(G^{(1)}_{2}) \mathbin{:=} \{(i,({-}q_t)^{p}) \in I_0 \times {{\mathbf{k}}}^{{\times}} \mid p \equiv_2 \mathrm{d}_\circ(2, i ) \}, \\ &\sigma_0(A^{(2)}_{2n}) \mathbin{:=} \{(i,({-}q)^{p}) \in I_0 \times {{\mathbf{k}}}^{{\times}} \mid p \in \mathbb{\mspace{1mu}Z\mspace{1mu}} \}, \\ &\sigma_0(A^{(2)}_{2n-1}) \mathbin{:=} \{( i,\pm({-}q)^{p}), ( n, ({-}q)^{r}) \mid 1 \leqslant i < n,\: p \equiv_2 i+1, \: r \equiv_2 n+1 \}, \\ & \sigma_0(D^{(2)}_{n+1}) \mathbin{:=} \{ (i, (\sqrt{-1}^{n+1-i}) ({-}q)^p), (n, \pm ({-}q)^{r}) \mid 1\leqslant i < n,\: p \equiv_2 i+1 ,\: r \equiv_2 n+1 \}, \\ & \sigma_0(E^{(2)}_{6}) \mathbin{:=} \{ (i,\pm q^r), (i', \sqrt{-1} ({-}q )^{r'}) \mid i \in \{ 1,2 \}, \: i' \in \{3,4\}, \: r \equiv_2 i+1, \: r'\equiv_2 i'+1 \}, \\ & \sigma_0(D^{(3)}_{4}) \mathbin{:=} \{ (1, q^r),(1, \omega q^r), (1, \omega^2 q^r), (2,-q^{r+1}) \mid r\equiv_2 0 \} \quad (\omega^2+\omega+1=0), \end{align*}

where $a \equiv _2 b$ means that $a\equiv b\mspace {3mu}\mathbin {\mathrm {mod}}\mspace {1mu} 2$ (see [Reference Hernandez and LeclercHL10, § 3.7], [Reference Kang, Kashiwara, Kim and OhKKKO16, § 4.1], [Reference Kashiwara and OhKO19, § 6] and [Reference Oh and ScrimshawOS19a, § 6]). Note that in [Reference Oh and ScrimshawOS19a, § 6] the category ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ and ${\sigma _Q({\mathfrak {g}})}$ were dealt with only in exceptional cases, but it is easy to obtain ${\sigma _0({\mathfrak {g}})}$ using ${\sigma _Q({\mathfrak {g}})}$. We use the notation $B_2^{(1)}$ and $A^{(2)}_{3}$ instead of $C_2^{(1)}$ and $D_{3}^{(2)}$, respectively. Here we use the standard convention for Dynkin diagrams in [Reference KacKac90, Ch. 4] except for $A^{(2)}_{2n}$, $A_3^{(2)}$, $B_2^{(1)}$ and $E^{(1)}_{k}$ ($k=6,7,8$), which are given in (2.1).

We define $\mathscr {C}_{\mathfrak {g}}^0$ to be the smallest full subcategory of $\mathscr {C}_{\mathfrak {g}}$ for which the following hold:

  1. (a) $\mathscr {C}_{\mathfrak {g}}^0$ contains $V(\varpi _i)_x$ for all $(i,x) \in {\sigma _0({\mathfrak {g}})}$;

  2. (b) $\mathscr {C}_{\mathfrak {g}}^0$ is stable by taking subquotients, extensions and tensor products.

For symmetric affine types, this category was introduced in [Reference Hernandez and LeclercHL10]. Note that every simple module in $\mathscr {C}_{\mathfrak {g}}$ is isomorphic to a tensor product of certain spectral parameter shifts of some simple modules in $\mathscr {C}_{\mathfrak {g}}^0$ (see [Reference Hernandez and LeclercHL10, § 3.7]).

2.4 The categories ${\mathscr {C}_{\mathfrak {g}}^{Q}}$

In this subsection, we recall very briefly a certain subcategory ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ of $\mathscr {C}_{\mathfrak {g}}$ categorifying the coordinate ring $\mathbb {C}[N]$ of the maximal unipotent group $N$ associated with a certain simple Lie algebra.

This subcategory ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ was introduced in [Reference Hernandez and LeclercHL15] for simply laced affine type ADE, in [Reference Kang, Kashiwara, Kim and OhKKKO16] for twisted affine types $A^{(2)}$ and $D^{(2)}$, in [Reference Kashiwara and OhKO19, Reference Oh and SuhOS19b] for untwisted affine types $B^{(1)}$ and $C^{(1)}$, and in [Reference Oh and ScrimshawOS19a] for exceptional affine type. The quantum affine Schur–Weyl duality functor between the finite-dimensional module category of a quiver Hecke algebra and ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ was also constructed in [Reference Kang, Kashiwara and KimKKK15] for untwisted affine types $A^{(1)}$ and $D^{(1)}$, in [Reference Kang, Kashiwara, Kim and OhKKKO16] for twisted affine types $A^{(2)}$ and $D^{(2)}$, in [Reference Kashiwara and OhKO19] for untwisted affine types $B^{(1)}$ and $C^{(1)}$, in [Reference Oh and ScrimshawOS19a] for exceptional affine type, and in [Reference FujitaFuj20] for simply laced affine type ADE in a geometric manner.

We shall describe ${\sigma _Q({\mathfrak {g}})}$ and ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ by using $Q$-data [Reference Fujita and OhFO21]. A $Q$-datum generalizes a Dynkin quiver with a height function, which provides a uniform way of describing the Hernandez–Leclerc category ${\mathscr {C}_{\mathfrak {g}}^{Q}}$. Our brief explanation follows [Reference Fujita and OhFO21, § 3] (see also [Reference Fujita, Hernandez, Oh and OyaFHOO21, § 4] and [Reference Kashiwara, Kim, Oh and ParkKKOP21, § 6]). Let ${\mathfrak {g}}$ be an affine Kac–Moody algebra and let ${{\mathfrak {g}}_{\mathrm {fin}}}$ be the simply laced finite-type Lie algebra corresponding to the affine type of ${\mathfrak {g}}$ in table (4.5). Let ${\mathrm {I_{fin}}}$ be the index set of ${{\mathfrak {g}}_{\mathrm {fin}}}$ and let $\mathsf {D}_{\rm fin}$ be the Dynkin diagram for ${{\mathfrak {g}}_{\mathrm {fin}}}$.

We first assume that ${\mathfrak {g}}$ is of untwisted type. We define an Dynkin diagram automorphism $\varrho$ of $\mathsf {D}_{\rm fin}$ as follows. For ${\mathfrak {g}}=A_n^{(1)}, D_n^{(1)}$ or $E_k^{(1)}$ type ($k=6,7,8$) we set $\varrho :={\rm id}$, and for the remaining types $\varrho$ is defined as follows (see [Reference Fujita and OhFO21, § 3.1]).

Let $I_0 =\{ 1,2,\ldots, n \}$ be the index set of ${\mathfrak {g}}_0$. Note that ${\mathrm {I_{fin}}} = I_0$ when ${\mathfrak {g}}=A_n^{(1)}, D_n^{(1)}, E_k^{(1)}$ ($k=6,7,8$). Let $\mathrm {ord}(\varrho )$ be the order of $\varrho$. For $i\in {\mathrm {I_{fin}}}$, we denote by $\mathrm {orb}(i)$ the orbit of $i$ under the action $\varrho$ and set $\mathsf {d}_i := | \mathrm {orb}(i) |$. We identify the set of orbits of ${\mathrm {I_{fin}}}$ with $I_0$ by mapping $\mathrm {orb}(i) \mapsto \min \{ \mathrm {orb}(i) \}$ for ${\mathfrak {g}} \ne F_4^{(1)}$ and mapping $\mathrm {orb}(1) \mapsto 1$, $\mathrm {orb}(3) \mapsto 2$, $\mathrm {orb}(4) \mapsto 3$ and $\mathrm {orb}(2) \mapsto 4$ for ${\mathfrak {g}} = F_4^{(1)}$. We write $\pi \colon {\mathrm {I_{fin}}}\to I_0$ for the projection via this identification.

Definition 2.3 [Reference Fujita and OhFO21, Definition 3.5]

A function $\xi \colon {\mathrm {I_{fin}}} \to \mathbb {\mspace {1mu}Z\mspace {1mu}}$ is called a height function on $(\mathsf {D}_{\rm fin}, \varrho )$ if the following two conditions are satisfied.

  1. (i) For any $i, j \in {\mathrm {I_{fin}}}$ such that $\mathrm {d}(i,j)=1$ and $\mathsf {d}_i= \mathsf {d}_j$, we have $|\xi _{i} - \xi _{j}| =\mathsf {d}_i$.

  2. (ii) For any $i,j \in {\mathrm {I_{fin}}}$ such that $\mathrm {d}(i,j)=1$ and $1=\mathsf {d}_i < \mathsf {d}_j=\mathrm {ord}(\varrho )$, there exists a unique element $j^{\circ } \in \mathrm {orb}(j)$ such that $|\xi _{i} - \xi _{j^{\circ }}| = 1$ and $\xi _{\varrho ^{k}(j^{\circ })} = \xi _{j^{\circ }} - 2k$ for any $0 \leqslant k < \mathrm {ord}(\varrho )$.

Here $\mathrm {d}(i,j)$ denotes the distance between $i$ and $j$ in the Dynkin diagram $\mathsf {D}_{\rm fin}$. We call the triple $Q = (\mathsf {D}_{\rm fin}, \varrho, \xi )$ a $Q$-datum for ${\mathfrak {g}}$.

For a $Q$-datum $Q = (\mathsf {D}_{\rm fin}, \varrho, \xi )$ associated to ${\mathfrak {g}}$, let

\[ \hat{I}_Q \mathbin{:=} \{ ( i ,p) \in {\mathrm{I_{fin}}} \times \mathbb{\mspace{1mu}Z\mspace{1mu}} \mid p -\xi_i \in 2\mathsf{d}_i\mathbb{\mspace{1mu}Z\mspace{1mu}}\}. \]

The generalized $\varrho$-Coxeter element $\tau _Q \in \mathsf {W}_{\rm fin} \rtimes \textrm {Aut} (\mathsf {D}_{\rm fin})$ associated with $Q$ is defined in [Reference Fujita and OhFO21, Definition 3.33] and can be understood as a generalization of a Coxeter element. Here $\mathsf {W}_{\rm fin}$ is the Weyl group of ${{\mathfrak {g}}_{\mathrm {fin}}}$.

For $i\in I_0$, we denote by $o(i)$ the corresponding orbit of ${\mathrm {I_{fin}}}$. For each $i\in I_0$, we denote by $i^\circ$ the unique vertex in the orbit $o( i)$ satisfying $\xi _{i^\circ } = \max \{ \xi _j \mid j\in o(i) \}$. In this paper, we assume further that the height function $\xi$ satisfies

(2.10)\begin{equation} \xi_{\varrho^k(i^\circ)} = \xi_{i^\circ} - 2k \quad \text{for each $i\in I_0$ and $0 \leqslant k < \mathsf{d}_i $.} \end{equation}

Let $\{ i_1, i_2, \ldots, i_n\}$ be a total order of $I_0$ satisfying $\xi _{i_1^\circ } \geqslant \xi _{i_2^\circ } \geqslant \cdots \geqslant \xi _{i_n^\circ }$. Then we have

\[ \tau_Q = s_{i_1^\circ} s_{i_2^\circ} \cdots s_{i_n^\circ} \varrho \in \mathsf{W}_{\rm fin} \rtimes \textrm{Aut} (\mathsf{D}_{\rm fin}) \]

(see [Reference Fujita and OhFO21, § 3.6] and also [Reference Fujita, Hernandez, Oh and OyaFHOO21, Proposition 4.4] for more details).

Let ${\Delta ^+_{Q}}$ be the set of positive roots of ${{\mathfrak {g}}_{\mathrm {fin}}}$, and let $\hat {\Phi } \mathbin {:=} {\Delta ^+_{Q}} \times \mathbb {\mspace {1mu}Z\mspace {1mu}}$. For each $i \in {\mathrm {I_{fin}}}$ we define

\[ \gamma^Q_i \mathbin{:=} (1-\tau_Q^{\mathsf{d}_i})\Lambda_i \in {\Delta^+_{Q}}, \]

where $\Lambda _i$ is the $i$th fundamental weight of ${{\mathfrak {g}}_{\mathrm {fin}}}$. It is shown in [Reference Hernandez and LeclercHL15, § 2.2] and [Reference Fujita and OhFO21, Theorem 3.35] that there exists a unique bijection $\psi _Q \colon \hat {I}_Q \to \hat {\Phi }$ defined inductively as follows:

  1. (i) $\psi _Q(i,\xi _i)=(\gamma ^Q_i,0)$;

  2. (ii) if $\psi _Q(i,p)=(\beta,m)$, then define:

    1. (a) $\psi _Q(i,p\pm 2\mathsf {d}_i) = (\tau _Q^{\mp \mathsf {d}_i} (\beta ),m)$ if $\tau _Q^{ \mp \mathsf {d}_i}(\beta ) \in {\Delta ^+_{Q}}$;

    2. (b) $\psi _Q(i,p\pm 2\mathsf {d}_i) = (-\tau _Q^{\mp \mathsf {d}_i}(\beta ),m\pm 1)$ if $\tau _Q^{\mp \mathsf {d}_i}(\beta ) \in - {\Delta ^+_{Q}}$.

Let $I_Q\mathbin {:=} \psi _Q^{-1}({\Delta ^+_{Q}}\times \{0\})\subset {\mathrm {I_{fin}}}\times \mathbb {\mspace {1mu}Z\mspace {1mu}}.$ Then one can describe

\[ I_Q=\{{(i,p)\in\hat{I}_Q\mid \xi_{ i^*} - \mathrm{ord}(\varrho)\mathsf{h}^\vee{<} p \leqslant \xi_{i}}\}, \]

where $\mathsf {h}^\vee$ is the dual Coxeter number of ${\mathfrak {g}}_0$ (see [Reference Fujita and OhFO21, Theorem 3.35] and also [Reference Fujita, Hernandez, Oh and OyaFHOO21, Proposition 4.15]). We define

\[ {\sigma_Q({\mathfrak{g}})} := \{ \zeta(i,p ) \mid (i,p) \in I_Q \}, \]

where we set ${{q}_{\mspace {1mu}{\scriptstyle {\mathrm {sh}}}}} \mathbin {:=} q^{1/\mathrm {ord}(\varrho )}$ and

\begin{align*} \zeta(i,p) := \begin{cases} (\pi(i), (-{{q}_{\mspace{1mu}{\scriptstyle{\mathrm{sh}}}}})^p) & \text{if } {\mathfrak{g}} = A^{(1)}_n, \; C_n^{(1)}, \; D^{(1)}_n, \; E_{6,7,8}^{(1)}, \; G_2^{(1)}, \\ (\pi(i), ({-}1)^{i+n}({{q}_{\mspace{1mu}{\scriptstyle{\mathrm{sh}}}}})^p) & \text{if } {\mathfrak{g}} = B_n^{(1)}, \\ (\pi(i), ({-}1)^{ \pi(i) }({{q}_{\mspace{1mu}{\scriptstyle{\mathrm{sh}}}}})^p) & \text{if } {\mathfrak{g}} = F_4^{(1)} \end{cases} \end{align*}

(see [Reference Fujita and OhFO21, § 5.4]). We define

(2.11)\begin{equation} \phi_Q\colon {\Delta^+_{Q}} \mathop{\xrightarrow{\sim}}{\sigma_Q({\mathfrak{g}})} \end{equation}

by $\phi _Q( \beta ) := \zeta \circ \psi _Q^{-1} (\beta, 0)$ for $\beta \in {\Delta ^+_{Q}}$. The map $\phi _Q$ is bijective.

For the rest of this paper, we make the following choices of $Q$-data:

  • for simply laced ADE type, $\mathrm {ord}(\varrho )=1$ and the height function $\xi$ is defined in Appendix A.1;

  • for ${\mathfrak {g}}=B_n^{(1)}$, $\mathrm {ord}(\varrho )=2$ and ;

  • for ${\mathfrak {g}}=C_n^{(1)}$, $\mathrm {ord}(\varrho )=2$ and ;

  • for ${\mathfrak {g}}=F_4^{(1)}$, $\mathrm {ord}(\varrho )=2$ and ;

  • for ${\mathfrak {g}}=G_2^{(1)}$, $\mathrm {ord}(\varrho )=3$ and .

Here an underlined integer stands for the value of $\xi _i$ at each vertex $i \in \mathsf {D}_{\rm fin}$ and an arrow $i \to j$ means that $\xi _i>\xi _j$ and $\mathrm {d}(i,j)=1$ in the Dynkin diagram $\mathsf {D}_{\rm fin}$. Note that our choice of $Q$ satisfies (2.10). Then $\tau _Q$ is given as follows:

  • for simply laced ADE type, $\tau _Q$ is the same as $\tau$ in Appendix A.1;

  • for ${\mathfrak {g}}=B_n^{(1)}, C_n^{(1)}$, $\tau _Q = s_1 s_2 \cdots s_n \varrho$;

  • for ${\mathfrak {g}}=F_4^{(1)}$, $\tau _Q = s_1 s_2 s_3 s_4 \varrho$;

  • for ${\mathfrak {g}}=G_2^{(1)}$, $\tau _Q = s_2 s_1 \varrho$.

In this case the set ${\sigma _Q({\mathfrak {g}})}$ is contained in ${\sigma _0({\mathfrak {g}})}$ in § 2.3 and can be written explicitly as follows (where $a\leqslant _2 b$ means that $a\leqslant b$ and $a\equiv b\mspace {3mu}\mathbin {\mathrm {mod}}\mspace {1mu} 2$):

\begin{align*} &\sigma_Q(A^{(1)}_{n }) \mathbin{:=} \{(i,({-}q)^{k}) \in \sigma_0(A^{(1)}_{n }) \mid i-2n +1 \leqslant_2 k \leqslant_2 -i +1 \}, \\ &\sigma_Q(B^{(1)}_{n}) \mathbin{:=} \{ (i,({-}1)^{n+i}q_s^{k}), (n,q^{k'}) \in \sigma_0(B^{(1)}_{n}) \mid i < n, -2n-2i+3 \leqslant_2 k \leqslant_2 2n-2i-1, \\ &\hspace{48ex} -\!2n+2 \leqslant k' \leqslant 0 \}, \\ & \sigma_Q({C_{n}^{(1)}}) \mathbin{:=} \{ (i, ({-}q_s)^{k}) \in \sigma_0(C^{(1)}_{n}) \mid{-}\mathrm{d}_\circ(1,i)-2n \leqslant_2 k \leqslant_2 -\mathrm{d}_\circ(1,i) \} , \\ & \sigma_Q({D_{n}^{(1)}}) \mathbin{:=} \{ (i, ({-}q)^{k}) \in \sigma_0(D^{(1)}_{n}) \mid{-}\mathrm{d}_\circ(1,i)-2n+4 \leqslant_2 k \leqslant_2 -\mathrm{d}_\circ(1,i) \} , \\ & \sigma_Q({E_{6}^{(1)}}) \mathbin{:=} \{ (i,({-}q)^k) \in \sigma_0(E^{(1)}_{6}) \mid \mathrm{d}_\circ(1,i)-14 \leqslant_2 k \leqslant_2 -\mathrm{d}_\circ(1,i) + 2\delta_{i,2} \}, \\ & \sigma_Q({E_{7}^{(1)}}) \mathbin{:=} \{ (i,({-}q)^k) \in \sigma_0(E^{(1)}_{7}) \mid{-}\mathrm{d}_\circ(1,i)-16 + 2\delta_{i,2} \leqslant_2 k \leqslant_2-\mathrm{d}_\circ(1,i) + 2\delta_{i,2} \}, \\ & \sigma_Q({E_{8}^{(1)}}) \mathbin{:=} \{ (i,({-}q)^k) \in \sigma_0(E^{(1)}_{8}) \mid{-}\mathrm{d}_\circ(1,i)-28 + 2\delta_{i,2} \leqslant_2 k \leqslant_2 -\mathrm{d}_\circ(1,i) + 2\delta_{i,2} \}, \\ & \sigma_Q({F_{4}^{(1)}}) \mathbin{:=} \biggl\{ (i, ({-}1)^iq^{k}) \in \sigma_0(F^{(1)}_{4}) \biggm| \mathrm{d}_\circ(i,3)-10 + \frac{\delta_{i,3}}{2} \leqslant k \leqslant \mathrm{d}_\circ(i,3)-2 + \frac{\delta_{i,3}}{2} \biggr\}, \\ & \sigma_Q({G_{2}^{(1)}}) \mathbin{:=} \{ (i, ({-}q_t)^{k}) \in \sigma_0(G^{(1)}_{2}) \mid{-}\mathrm{d}_\circ(2,i)-10 \leqslant_2 k \leqslant_2 -\mathrm{d}_\circ(2,i) \} , \end{align*}

where $\mathrm {d}_\circ (i,j)$ denotes the distance between $i$ and $j$ in the Dynkin diagram of ${\mathfrak {g}}_0$.

We now assume that ${\mathfrak {g}}$ is of twisted type. Then one can define

\begin{align*} &\sigma_Q(A^{(2)}_{N}) \mathbin{:=} \{(i, ({-}q)^{k} )^\star \mid (i,({-}q)^{k}) \in \sigma_Q(A^{(1)}_{N}) \} \quad (N=2n-1 \text{ or } 2n) , \\ &\sigma_Q(D^{(2)}_{n+1}) \mathbin{:=} \{(i, ({-}q)^{k})^\star\mid (i,({-}q)^{k}) \in \sigma_Q(D^{(1)}_{n+1}) \}, \\ &\sigma_Q(E^{(2)}_{6}) \mathbin{:=} \{(i,({-}q)^{k} )^\star\mid (i,({-}q)^{k}) \in \sigma_Q(E^{(1)}_{6}) \}, \\ &\sigma_Q(D^{(3)}_{4}) \mathbin{:=} \{(i,({-}q)^{k})^{{\dagger}}\mid (i,({-}q)^{k}) \in \sigma_Q(D^{(1)}_{4}) \}, \end{align*}

where for $(i,a) \in \sigma _0({\mathfrak {g}}^{(1)}_{N})$ we set

\[ \begin{aligned} (i,a)^\star{=} \begin{cases} (i,a) & \text{if } \mathfrak{g}=A^{(1)}_{N}, \, i \leqslant \lfloor (N+1)/2 \rfloor \:\text{ or }\: \mathfrak{g}=E^{(1)}_{6}, \, i =1, \\ (N+1-i,({-}1)^Na) & \text{if } \mathfrak{g}=A^{(1)}_{N}, \: i > \lfloor (N+1)/2 \rfloor, \\ (i,\sqrt{-1}^{n+1-i}a) & \text{if } \mathfrak{g}=D^{(1)}_{n+1}, \: i \leqslant n-1, \\ (n,({-}1)^ia) & \text{if } \mathfrak{g}=D^{(1)}_{n+1}, \: i\in \{n,n+1\}, \\ (2,a) & \text{if } \mathfrak{g}=E^{(1)}_{6}, \: i=3, \\ (2,-a) & \text{if } \mathfrak{g}=E^{(1)}_{6}, \: i=5, \\ (1,-a) & \text{if } \mathfrak{g}=E^{(1)}_{6}, \: i=6 , \\ (3,\sqrt{-1}a) & \text{if } \mathfrak{g}=E^{(1)}_{6}, \: i=4 , \\ (4,\sqrt{-1}a) & \text{if } \mathfrak{g}=E^{(1)}_{6}, \: i=2 \end{cases} \end{aligned} \]

and

\begin{align*} (i,a)^{{\dagger}}= \begin{cases} (2,a) & \text{if } i=2, \\ (1,(\delta_{i,1}+\delta_{i,3} \omega+\delta_{i,4} \omega^2)a) & \text{if } i \ne 2 \end{cases} \end{align*}

(see [Reference Kang, Kashiwara, Kim and OhKKKO16, Proposition 4.3] and [Reference Oh and ScrimshawOS19a, Proposition 6.5] for details of $\star$ and ${{\dagger}}$). The bijection $\phi _Q\colon {\Delta ^+_{Q}} \mathop {\xrightarrow {\sim}}{\sigma _Q({\mathfrak {g}})}$ is defined by composing the bijection for untwisted type with the maps $\star$ and ${{\dagger}}$.

Comparing the above descriptions of ${\sigma _Q({\mathfrak {g}})}$ with the descriptions of ${\sigma _0({\mathfrak {g}})}$ given in § 2.3, one can easily show that

(2.12)\begin{equation} \begin{aligned} & {\sigma_0({\mathfrak{g}})} = \bigsqcup_{k \in \mathbb{\mspace{1mu}Z\mspace{1mu}} } {\sigma_Q({\mathfrak{g}})} ^{*k}, \\ & {\sigma_Q({\mathfrak{g}})} ^{*k} \cap {\sigma_Q({\mathfrak{g}})} ^{*k'} = \emptyset \quad \text{for } k,k' \in \mathbb{\mspace{1mu}Z\mspace{1mu}} \text{ with } k \ne k', \end{aligned} \end{equation}

where ${\sigma _Q({\mathfrak {g}})} ^{*k} \mathbin {:=} \{ (i^{*k},(p^*)^ka) \mid (i,a) \in {\sigma _Q({\mathfrak {g}})} \}$ with $i^{*k} =i$ if $k$ is even and $i^{*k} =i^*$ if $k$ is odd (see [Reference Fujita and OhFO21, Proposition 5.9]). Note that $p^*$ is given in (2.5).

Let ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ be the smallest full subcategory of $\mathscr {C}_{\mathfrak {g}}^0$ with the following properties:

  1. (a) ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ contains $\mathbf {1}$ and $V(\varpi _i)_x$ for all $(i,x) \in {\sigma _Q({\mathfrak {g}})}$;

  2. (b) ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ is stable by taking subquotients, extensions and tensor products.

It was shown in [Reference Hernandez and LeclercHL15, Theorem 6.1], [Reference Kang, Kashiwara, Kim and OhKKKO16, Corollary 5.6], [Reference Kashiwara and OhKO19, Corollary 6.6] and [Reference Oh and ScrimshawOS19a, § 6] that the Grothendieck ring $K({\mathscr {C}_{\mathfrak {g}}^{Q}})$ of the monoidal category ${\mathscr {C}_{\mathfrak {g}}^{Q}}$ is isomorphic to the coordinate ring $\mathbb {C}[N]$ of the maximal unipotent group $N$ associated with ${{\mathfrak {g}}_{\mathrm {fin}}}$. The set ${\Delta ^+_{Q}}$ has a convex order $\prec _Q$ arising from $Q$.

Let $\beta \in {\Delta ^+_{Q}}$ and write $(i,a) = \phi _Q(\beta )$. Then set

\[ V_Q(\beta) \mathbin{:=} V(\varpi_i)_a \in {\mathscr{C}_{\mathfrak{g}}^{Q}}. \]

Under the categorification, the modules $V_Q(\beta )$ correspond to the dual PBW vectors of $\mathbb {C}[N]$ with respect to the convex order $\prec _Q$ on ${\Delta ^+_{Q}}$.

The proposition below follows from [Reference Kang, Kashiwara and KimKKK15, § 4.3], [Reference Kang, Kashiwara, Kim and OhKKKO16, Proposition 4.9 and Theorem 5.1], [Reference Kashiwara and OhKO19, § 4.3] and [Reference Oh and ScrimshawOS19a, § 6].

Proposition 2.4 [Reference Kang, Kashiwara and KimKKK15, Reference Kang, Kashiwara, Kim and OhKKKO16, Reference Kashiwara and OhKO19, Reference Oh and ScrimshawOS19a]

For a minimal pair $(\alpha, \beta )$ of a positive root $\gamma \in {\Delta ^+_{Q}}$, $V_Q(\gamma )$ is isomorphic to the head of $V_Q(\alpha ) \otimes V_Q(\beta )$. Here, $(\alpha, \beta )$ is called a minimal pair of $\gamma$ if $\alpha \prec _Q \beta$, $\gamma = \alpha +\beta$ and there exists no pair $(\alpha ', \beta ')$ such that $\gamma = \alpha '+\beta '$ and $\alpha \prec _Q \alpha ' \prec _Q \beta ' \prec _Q \beta$.

3. New invariants for pairs of modules

In this section, we recall several properties of the new invariants arising from R-matrices introduced in [Reference Kashiwara, Kim, Oh and ParkKKOP20].

We set

\[ \varphi(z) \mathbin{:=} \prod_{s=0}^\infty (1-{\tilde{p}}^{s}z) =\sum_{n=0}^\infty\frac{({-}1)^n{\tilde{p}}^{n(n-1)/2}}{\prod_{k=1}^n(1-{\tilde{p}}^k)} z^n \in{{\mathbf{k}}}[[z]]\subset\hat{\mathbf{k}}[[z]], \]

where $\tilde {p}$ is given in (2.5). We consider the subgroup $\mathcal {G}$ of $\mathbf {k}((z))^\times$ given by

\[ \mathcal{G} \mathbin{:=} \biggl\{ cz^m \prod_{a \in \mathbf{k}^\times} \varphi(az)^{\eta_a} \biggm| \begin{matrix} c \in \mathbf{k}^\times, \ m \in \mathbb{\mspace{1mu}Z\mspace{1mu}} , \\ \eta_a \in \mathbb{\mspace{1mu}Z\mspace{1mu}} \text{ vanishes except for finitely many $a$} \end{matrix} \biggr\}. \]

Note that if $R^{\mspace {1mu}\mathrm {univ}}_{M,N_z}$ is rationally renormalizable for $M, N \in \mathscr {C}_{\mathfrak {g}}$, then the renormalizing coefficient $c_{M,N}(z)$ belongs to $\mathcal {G}$ (see [Reference Kashiwara, Kim, Oh and ParkKKOP20, Proposition 3.2]). In particular, for simple modules $M$ and $N$ in $\mathscr {C}_{\mathfrak {g}}$, the universal coefficient $a_{M,N}(z)$ belongs to $\mathcal {G}$.

For a subset $S$ of $\mathbb {\mspace {1mu}Z\mspace {1mu}}$, let ${\tilde {p}}^{S} \mathbin {:=} \{ {\tilde {p}}^k \mid k \in S\}$. We define the group homomorphisms

\[ \mathrm{Deg} \colon \mathcal{G} \to \mathbb{\mspace{1mu}Z\mspace{1mu}} \quad \text{and} \quad \mathrm{Deg}^\infty \colon \mathcal{G} \to \mathbb{\mspace{1mu}Z\mspace{1mu}} \]

by

\[ \mathrm{Deg}(f(z)) = \sum_{a \in {\tilde{p}}^{\,\mathbb{\mspace{1mu}Z\mspace{1mu}}_{\leqslant 0}} }\eta_a - \sum_{a \in {\tilde{p}}^{\,\mathbb{\mspace{1mu}Z\mspace{1mu}}_{> 0}} } \eta_a \quad \text{and} \quad \mathrm{Deg}^\infty(f(z)) = \sum_{a \in {\tilde{p}}^{\,\mathbb{\mspace{1mu}Z\mspace{1mu}}}} \eta_a \]

for $f(z)=cz^m \prod _{ a\in \mathbf {k}^\times } \varphi (az)^{\eta _a} \in \mathcal {G}$.

Lemma 3.1 [Reference Kashiwara, Kim, Oh and ParkKKOP20, Lemma 3.4]

Let $f(z)\in \mathcal {G}$.

  1. (i) If $f(z)\in \mathbf {k}(z)^\times$, then we have $f(z)\in \mathcal {G}$,

    \[ \mathrm{Deg}^\infty(f(z))=0\quad \text{and}\quad \mathrm{Deg}(f(z))=2\,\mspace{1mu}\mathrm{zero}\mspace{1mu}_{z=1}f(z). \]
  2. (ii) If $g(z),h(z)\in \mathcal {G}$ satisfy $g(z)/h(z)\in {\mathbf {k}[z^{\pm 1}]}$, then $\mathrm {Deg}(h(z))\leqslant \mathrm {Deg}(g(z))$.

  3. (iii) We have that $\mathrm {Deg}^\infty f(z)= -\mathrm {Deg}\bigl ( f({\tilde {p}}^{n}z)\bigr )=\mathrm {Deg}\bigl ( f({\tilde {p}}^{-n}z)\bigr )$ for $n \gg 0$.

  4. (iv) If $\mathrm {Deg}^\infty \bigl ( f(cz)\bigr )=0$ for any $c\in \mathbf {k}^\times$, then $f(z)\in \mathbf {k}(z)^\times$.

The following invariants for a pair of modules $M$ and $N$ in $\mathscr {C}_{\mathfrak {g}}$ such that $R^{\mspace {1mu}\mathrm {univ}}_{M,N_z}$ is rationally renormalizable were introduced in [Reference Kashiwara, Kim, Oh and ParkKKOP20] by using the homomorphisms $\mathrm {Deg}$ and $\mathrm {Deg}^\infty$.

Definition 3.2 For non-zero modules $M$ and $N$ in $\mathscr {C}_{\mathfrak {g}}$ such that $R^{\mspace {1mu}\mathrm {univ}}_{M,N_z}$ is rationally renormalizable, we define the integers $\Lambda (M,N)$ and $\Lambda ^\infty (M,N)$ by

\begin{align*} \Lambda(M,N) & \mathbin{:=}\mathrm{Deg}(c_{M,N}(z)),\\ \Lambda^\infty(M,N) & \mathbin{:=}\mathrm{Deg}^\infty(c_{M,N}(z)). \end{align*}

We have $\Lambda (M,N)\equiv \Lambda ^\infty (M,N)\mspace {3mu}\mathbin {\mathrm {mod}}\mspace {1mu} 2$.

Proposition 3.3 [Reference Kashiwara, Kim, Oh and ParkKKOP20, Lemma 3.7]

For any simple modules $M,N\in \mathscr {C}_{\mathfrak {g}}$ and $x \in \mathbf {k}^\times$, we have

\[ \Lambda(M,N)=\Lambda(M_x,N_x) \quad\text{{and}}\quad \Lambda^\infty(M,N)=\Lambda^\infty(M_x,N_x). \]

Proposition 3.4 [Reference Kashiwara, Kim, Oh and ParkKKOP20, Lemmas 3.7 and 3.8 and Corollary 3.23]

Let $M$ and $N$ be simple modules in $\mathscr {C}_{\mathfrak {g}}$. Then the following hold:

  1. (i) $\Lambda ^\infty (M,N) =-\mathrm {Deg}^\infty (a_{M,N}(z))$;

  2. (ii) $\Lambda ^\infty (M,N) = \Lambda ^\infty (N,M)$;

  3. (iii) $\Lambda ^\infty (M,N) = - \Lambda ^\infty (M^*, N) = - \Lambda ^\infty (M,{}^*\mspace {-3mu} N)$;

  4. (iv) in particular, $\Lambda ^\infty (M,N) = \Lambda ^\infty (M^*, N^*) = \Lambda ^\infty ( {}^*\mspace {-3mu} M,{}^*\mspace {-3mu} N)$.

Proposition 3.5 [Reference Kashiwara, Kim, Oh and ParkKKOP20, Lemma 3.7 and Proposition 3.18]

Let $M$ and $N$ be simple modules in $\mathscr {C}_{\mathfrak {g}}$. Then the following hold:

  1. (i) $\Lambda (M,N) = \Lambda (N^*, M) = \Lambda (N,{}^*\mspace {-3mu} M)$;

  2. (ii) in particular,

    \[ \Lambda(M,N) = \Lambda(M^*, N^*) = \Lambda( {}^*\mspace{-3mu} M,{}^*\mspace{-3mu} N). \]

Proposition 3.6 [Reference Kashiwara, Kim, Oh and ParkKKOP20, Proposition 3.9]

Let $M$ and $N$ be modules in $\mathscr {C}_{\mathfrak {g}}$, and let $M'$ and $N'$ be non-zero subquotients of $M$ and $N$, respectively. Assume that $R^{\mspace {1mu}\mathrm {univ}}_{M,N_z}$ is rationally renormalizable. Then $R^{\mspace {1mu}\mathrm {univ}}_{M',N'_z}$ is rationally renormalizable, and we have

\[ \Lambda(M',N')\leqslant \Lambda(M,N)\quad\text{{and}}\quad\Lambda^\infty(M',N')=\Lambda^\infty(M,N). \]

Proposition 3.7 [Reference Kashiwara, Kim, Oh and ParkKKOP20, Proposition 3.11]

Let $M$, $N$ and $L$ be non-zero modules in $\mathscr {C}_{\mathfrak {g}}$, and let $S$ be a non-zero subquotient of $M\mathop \otimes N$.

  1. (i) Assume that $R^{\mspace {1mu}\mathrm {univ}}_{M,L_z}$ and $R^{\mspace {1mu}\mathrm {univ}}_{N,L_z}$ are rationally renormalizable. Then $R^{\mspace {1mu}\mathrm {univ}}_{S,L_z}$ is rationally renormalizable, and we have

    \[ \Lambda(S,L)\leqslant \Lambda(M,L) + \Lambda(N,L) \quad\text{{and}}\quad \Lambda^\infty(S,L)=\Lambda^\infty(M,L) + \Lambda^\infty(N,L). \]
  2. (ii) Assume that $R^{\mspace {1mu}\mathrm {univ}}_{L,M_z}$ and $R^{\mspace {1mu}\mathrm {univ}}_{L,N_z}$ are rationally renormalizable. Then $R^{\mspace {1mu}\mathrm {univ}}_{L,S_z}$ is rationally renormalizable, and we have

    \[ \Lambda(L,S)\leqslant\Lambda(L,M) + \Lambda(L,N) \quad\text{{and}}\quad\Lambda^\infty(L,S)=\Lambda^\infty(L,M) + \Lambda^\infty(L,N). \]

Corollary 3.8 [Reference Kashiwara, Kim, Oh and ParkKKOP20, Corollary 3.12]

Let $M$ and $N$ be simple modules in $\mathscr {C}_{\mathfrak {g}}$. Suppose that $M$ (respectively $N$) is isomorphic to a subquotient of $V(\varpi _{i_1 })_{a_1} \otimes V(\varpi _{i_2 })_{a_2}\otimes \cdots \otimes V(\varpi _{i_k})_{a_k}$ (respectively $V(\varpi _{j_1 })_{b_1} \otimes V(\varpi _{j_2 })_{b_2}\otimes \cdots \otimes V(\varpi _{j_l})_{b_l}$). Then we have

\[ \Lambda^{\infty}(M,N) = \sum_{ 1 \leqslant \nu \leqslant k, 1 \leqslant \mu \leqslant l} \Lambda^{\infty}(V(\varpi_{i_\nu})_{a_{\nu}},V(\varpi_{j_\mu})_{b_{\mu}}). \]

For simple modules $M$ and $N$ in $\mathscr {C}_{\mathfrak {g}}$, we define $\operatorname {\mathfrak {d}}(M,N)$ by

\[ \operatorname{\mathfrak{d}}(M,N) \mathbin{:=} \tfrac{1}{2}\bigl(\Lambda(M,N) + \Lambda(M^*,N)\bigr). \]

Proposition 3.9 [Reference Kashiwara, Kim, Oh and ParkKKOP20, Proposition 3.16 and Corollary 3.19]

Let $M$ and $N$ be simple modules in $\mathscr {C}_{\mathfrak {g}}$. Then the following hold:

  1. (i) $\operatorname {\mathfrak {d}}(M,N)=\mspace {1mu}\mathrm {zero}\mspace {1mu}_{z=1}\bigl ( d_{M,N}(z)d_{N,M}(z^{-1})\bigr )$;

  2. (ii) $\operatorname {\mathfrak {d}}(M,N)=\frac {1}{2}\bigl (\Lambda (M,N)+\Lambda (N,M)\bigr )$;

  3. (iii) in particular, $\operatorname {\mathfrak {d}}(M,N) = \operatorname {\mathfrak {d}}(N,M)$.

Corollary 3.10 [Reference Kashiwara, Kim, Oh and ParkKKOP20, Corollaries 3.17 and 3.20]

Let $M$ and $N$ be simple modules in $\mathscr {C}_{\mathfrak {g}}$.

  1. (i) Suppose that one of $M$ and $N$ is real. Then $M$ and $N$ strongly commute if and only if $\operatorname {\mathfrak {d}}(M,N)=0$.

  2. (ii) In particular, if $M$ is real, then $\Lambda (M,M)=0.$

Proposition 3.11 [Reference Kashiwara, Kim, Oh and ParkKKOP20, Proposition 3.22]

For simple modules $M$ and $N$ in $\mathscr {C}_{\mathfrak {g}}$, we have

\begin{align*} \Lambda(M,N) & = \sum_{k \in \mathbb{\mspace{1mu}Z\mspace{1mu}}} ({-}1)^{k+\delta(k<0)} \operatorname{\mathfrak{d}}(M,\mathscr{D}^{k}N) , \\ \Lambda^\infty(M,N) &= \sum_{k \in \mathbb{\mspace{1mu}Z\mspace{1mu}}} ({-}1)^{k} \operatorname{\mathfrak{d}}(M,\mathscr{D}^{k}N), \end{align*}

where $\mathscr {D}^kN$ is defined as

\[ \mathscr{D}^kN \mathbin{:=} \begin{cases} (\cdots(N^* \underbrace{ )^* \cdots )^* }_{\text{$({-}k)$ times}} & \text{if } k <0, \\ \underbrace{{}^*\mspace{-3mu} ( \cdots ( }_{\text{$k$ times}} {}^*\mspace{-3mu} N )\cdots) & \text{if } k \geqslant 0. \end{cases} \]

4. Root systems associated with $\mathscr {C}_{\mathfrak {g}}$

Let $\operatorname {Hom}({\sigma ({\mathfrak {g}})}, \mathbb {\mspace {1mu}Z\mspace {1mu}})$ be the set of $\mathbb {\mspace {1mu}Z\mspace {1mu}}$-valued functions on ${\sigma ({\mathfrak {g}})}$. It is obvious that $\operatorname {Hom}({\sigma ({\mathfrak {g}})}, \mathbb {\mspace {1mu}Z\mspace {1mu}})$ forms a torsion-free abelian group under addition. Let $M \in \mathscr {C}_{\mathfrak {g}}$ be a module such that $R^{\mspace {1mu}\mathrm {univ}}_{M, V(\varpi _i)_z}$ is rationally renormalizable for any $i\in I_0$. Then we define $\mathsf {E} (M) \in \operatorname {Hom}({\sigma ({\mathfrak {g}})}, \mathbb {\mspace {1mu}Z\mspace {1mu}})$ by

(4.1)\begin{equation} \mathsf{E} (M)(i,a)\mathbin{:=} \Lambda^\infty(M, V(\varpi_i)_a) \quad \text{for } (i,a) \in {\sigma({\mathfrak{g}})}, \end{equation}

which is well-defined by (2.4).

Lemma 4.1 Let $M$ and $N$ be simple modules in $\mathscr {C}_{\mathfrak {g}}$.

  1. (i) We have $\mathsf {E} (M) = - \mathsf {E} (M^*) = - \mathsf {E} ({}^*\mspace {-3mu} M )$.

  2. (ii) Let $\{{M_k}\}_{1\leqslant k\leqslant r}$ be a sequence of simple modules. Then for any non-zero subquotient $S$ of $M_1\mathop \otimes \cdots \mathop \otimes M_r$, we have

    \[ \mathsf{E} (S) =\sum_{k=1}^r \mathsf{E} (M_k). \]
  3. (iii) $\mathsf {E} (M) = \mathsf {E} (N)$ if and only if ${a_{M,V(\varpi _i)}(z) }/{a_{N,V(\varpi _i)}(z) } \in \mathbf {k}(z)^\times$ for any $i\in I_0$.

Proof. Assertions (i) and (ii) easily follow from Propositions 3.4 and 3.7.

Let us show (iii). For $(i,a) \in {\sigma ({\mathfrak {g}})}$, the condition $\Lambda ^\infty (M, V(\varpi _i)_a) = \Lambda ^\infty (N, V(\varpi _i)_a)$ is equivalent to

\[ \mathrm{Deg}^\infty(a_{M, V(\varpi_i)} (az) ) = \mathrm{Deg}^\infty(a_{N, V(\varpi_i)}(az) ). \]

Since $\mathrm {Deg}^\infty \colon \mathcal {G} \rightarrow \mathbb {\mspace {1mu}Z\mspace {1mu}}$ is a group homomorphism, it is equivalent to

\[ \mathrm{Deg}^\infty \bigg( \frac{a_{M, V(\varpi_i)}(az)}{a_{N, V(\varpi_i)}(az)} \bigg)=0 \quad \text{for any }a\in \mathbf{k}^\times. \]

Then (iii) follows from Lemma 3.1(iv).

For $(i,a) \in {\sigma ({\mathfrak {g}})}$, we set

\[ \mathsf{s} _{i,a} \mathbin{:=} \mathsf{E} (V(\varpi_i)_a) \in \operatorname{Hom}({\sigma({\mathfrak{g}})}, \mathbb{\mspace{1mu}Z\mspace{1mu}}) \]

and

(4.2)\begin{alignat}{4} \mathcal{W} & \mathbin{:=} \{ \mathsf{E} (M) \mid M \text{ is simple in } \mathscr{C}_{\mathfrak{g}}\}, & \quad \Delta & \mathbin{:=} \{ \mathsf{s} _{i,a} \mid (i,a) \in {\sigma({\mathfrak{g}})} \} \subset \mathcal{W} , \nonumber \\[-6pt] \mathcal{W} _0 & \mathbin{:=} \{ \mathsf{E} (M) \mid M \text{ is simple in } \mathscr{C}_{\mathfrak{g}}^0 \}, & \quad \Delta_0 & \mathbin{:=} \{ \mathsf{s} _{i,a} \mid (i,a) \in {\sigma_0({\mathfrak{g}})} \} \subset \mathcal{W} _0. \end{alignat}

It is obvious that $\mathcal {W} _0 \subset \mathcal {W}$ and $\Delta _0 \subset \Delta$.

Lemma 4.2

  1. (i) We have $\mathcal {W} = \sum _{(i,a) \in {\sigma ({\mathfrak {g}})}} \mathbb {\mspace {1mu}Z\mspace {1mu}} \mathsf {s} _{i,a}$ and $\mathcal {W} _0 = \sum _{(i,a) \in {\sigma _0({\mathfrak {g}})}} \mathbb {\mspace {1mu}Z\mspace {1mu}} \mathsf {s} _{i,a}=\sum _{(i,a) \in {\sigma _Q({\mathfrak {g}})}} \mathbb {\mspace {1mu}Z\mspace {1mu}} \mathsf {s} _{i,a}$. In particular, $\mathcal {W} _0$ is a finitely generated free $\mathbb {\mspace {1mu}Z\mspace {1mu}}$-module.

  2. (ii) There exists a unique symmetric bilinear form $(-,-)$ on $\mathcal {W}$ such that

    \[ ( \mathsf{E} (M), \mathsf{E} (N) ) ={-} \Lambda^\infty(M,N) \]
    for any simple modules $M,N \in \mathscr {C}_{\mathfrak {g}}$.

Proof. Assertion (i) follows from Theorem 2.2(iv), Lemma 4.1 and (2.12).

Let us show (ii). By Corollary 3.8, it reduces to the existence of a bilinear form $(-,-)$ on $\mathcal {W}$ such that

\[ \bigl( \mathsf{s} _{i,a}, \mathsf{s} _{j,b}\bigr)={-}\Lambda^\infty\bigl( V(\varpi_i)_a,V(\varpi_j)_b\bigr). \]

Therefore it is enough to show that for a sequence $\{{(i_k,a_k)}\}_{k=1,\ldots,r}$ in ${\sigma ({\mathfrak {g}})}$ such that ${\sum _{k=1}^r \mathsf {s} _{i_k,a_k}=0}$, we have $\sum _{k=1}^r\Lambda ^\infty \bigl ( V(\varpi _{i_k})_{a_k},V(\varpi _j)_b\bigr )=0$ for any $(j,b)\in {\sigma ({\mathfrak {g}})}$. Let us take a simple subquotient $M$ of $V(\varpi _{i_1})_{a_1}\mathop \otimes \cdots \mathop \otimes V(\varpi _{i_r})_{a_r}$. Then we have $\mathsf {E} (M)= \sum _{k=1}^r \mathsf {s} _{i_k,a_k}=0$. Hence we obtain

\[ \sum _{k=1}^r\Lambda^\infty\bigl( V(\varpi_{i_k})_{a_k},V(\varpi_j)_b\bigr)= \Lambda^\infty(M,V(\varpi_j)_b)={-} \mathsf{E} (M)(j,b)=0. \]

Lemma 4.3 For $i\in I_0$ and $a\in \mathbf {k}^\times$, we have

(4.3)\begin{equation} \operatorname{\mathfrak{d}}( V(\varpi_i), \mathscr{D}^k V(\varpi_{i}))= \delta(k=\pm1)\quad\text{for $k\in\mathbb{\mspace{1mu}Z\mspace{1mu}}$.}\quad\end{equation}

In particular,

\[ ( \mathsf{s} _{i,a}, \mathsf{s} _{i,a})={-}\Lambda^\infty( V(\varpi_i), V(\varpi_{i}) ) = 2. \]

Proof. The statement $\Lambda ^\infty ( V(\varpi _i), V(\varpi _{i}) ) = -2$ follows from (4.3) and Proposition 3.11.

Let us show (4.3). Let $h^\vee$ be the dual Coxeter number of ${\mathfrak {g}}$, and write

\[ d_{i,j}(z) \mathbin{:=} d_{V(\varpi_i), V(\varpi_j)} (z) \quad \text{for } i,j \in I. \]

The denominator formula for $d_{i,j}(z)$ is given in Appendix A. Using this formula, one can easily check that if $\epsilon q^t$ ($|\epsilon |=1$) is a zero of $d_{i,i}(z)$, then $t$ should be between $1$ and $h^\vee$. Combining this with Proposition 3.9, we obtain

\[ \operatorname{\mathfrak{d}}( V(\varpi_i), V(\varpi_{j})_{ (p^*)^{k}} ) = \mspace{1mu}\mathrm{zero}\mspace{1mu}_{z=1}\bigl( d_{i,j }( (p^*)^k z) d_{j,i}( (p^*)^{{-}k} z^{{-}1})\bigr) = 0 \quad \text{unless } k={\pm} 1. \]

Now we shall show that $\operatorname {\mathfrak {d}}( V(\varpi _i), \mathscr {D}^{\pm 1} V(\varpi _{ i }) )=1$.

Case of simply laced affine ADE type. In this case, the dual Coxeter number is equal to the Coxeter number. Then from the denominator formula in Appendix A it follows that

\[ \operatorname{\mathfrak{d}}( V(\varpi_{i}),\mathscr{D}^{\pm1} V(\varpi_{i})) = \tilde{c}_{i,i^*}(h^\vee{-}1). \]

Since $\tilde {c}_{i,j}(k) = \tilde {c}_{j,i^*}(h^\vee - k)$ for $1 \leqslant k \leqslant h^\vee -1$ (see [Reference FujitaFuj22, Lemma 3.7]) and $\tilde {c}_{i ,i}( 1)=1$ by Proposition A.1, we have

\[ \operatorname{\mathfrak{d}}( V(\varpi_{i}),\mathscr{D}^{\pm1} V(\varpi_{i}))= \tilde{c}_{i,i}( 1)= 1. \]

Other case. In this case, we know that $i^* = i$ for any $i\in I_0$. Thus we have

\[ \operatorname{\mathfrak{d}}( V(\varpi_i), \mathscr{D}^{\pm1} V(\varpi_{i})) = \operatorname{\mathfrak{d}}( V(\varpi_i), V(\varpi_{i })_{ p^* } ). \]

Using (2.6) and the denominator formula in Appendix A, one can compute directly that

\[ \operatorname{\mathfrak{d}}( V(\varpi_i), V(\varpi_{i })_{ p^* } ) = 1. \]

For $t \in \mathbf {k}^\times$, $(i,a) \in {\sigma ({\mathfrak {g}})}$ and $f \in \operatorname {Hom}({\sigma ({\mathfrak {g}})}, \mathbb {\mspace {1mu}Z\mspace {1mu}})$, we define

(4.4)\begin{equation} \tau_t( i,a ) \mathbin{:=} ( i, ta ) \quad \text{and} \quad (\tau_t f)( i, a) \mathbin{:=} f(i,t^{{-}1}a). \end{equation}

Lemma 4.4

  1. (i) For $(i,a) \in {\sigma ({\mathfrak {g}})}$, we have $\mathsf {s} _{i,a} = - \mathsf {s} _{i^*,a p^* } = - \mathsf {s} _{i^*,a (p^*)^{-1} }$.

  2. (ii) For $t\in \mathbf {k}^\times$ and $(i,a) \in {\sigma ({\mathfrak {g}})}$, we have $\tau _t ( \mathsf {s} _{i,a}) = \mathsf {s} _{i,ta}$.

Proof. Assertion (i) follows from (2.7) and Lemma 4.1.

(ii) For $(j,b)\in {\sigma ({\mathfrak {g}})}$, we have

\begin{align*} (\tau_t ( \mathsf{s} _{i,a})) (j,b) &= ( \mathsf{s} _{i,a}) (j,t^{{-}1}b) = \Lambda^\infty( V(\varpi_i)_a, V(\varpi_j)_{t^{{-}1}b} ) = \Lambda^\infty( V(\varpi_i)_{ta}, V(\varpi_j)_{b} ) \\ & = ( \mathsf{s} _{i,ta}) (j, b), \end{align*}

where the third equality follows from Proposition 3.3. Thus, we have the desired assertion.

For $t\in \mathbf {k}^\times$, $A \subset {\sigma ({\mathfrak {g}})}$ and $F \subset \operatorname {Hom}({\sigma ({\mathfrak {g}})}, \mathbb {\mspace {1mu}Z\mspace {1mu}})$, we set

\[ A _t \mathbin{:=} \{ \tau_t(a) \mid a \in A \} \quad \text{and } \quad F_t \mathbin{:=} \{ \tau_t(f) \mid f\in F \}. \]

We write ${\mathbf {k}_0}$ for the stabilizer subgroup of ${\sigma _0({\mathfrak {g}})}$ with respect to the action of $\mathbf {k}^\times$ on ${\sigma ({\mathfrak {g}})}$ through $\tau _t$, i.e.

\[ {\mathbf{k}_0} \mathbin{:=} \{ t \in \mathbf{k}^\times{\mid} ( {\sigma_0({\mathfrak{g}})} )_t = {\sigma_0({\mathfrak{g}})} \}. \]

Proposition 4.5 The following hold:

  1. (i) ${\sigma ({\mathfrak {g}})} = \bigsqcup _{a\in \mathbf {k}^\times / {\mathbf {k}_0} } \bigl ({\sigma _0({\mathfrak {g}})}\bigr )_a$;

  2. (ii) $\Delta = \bigsqcup _{a\in \mathbf {k}^\times / {\mathbf {k}_0} } (\Delta _0)_a$;

  3. (iii) for $k,k'\in \mathbf {k}^\times$ such that $k/k'\not \in {\mathbf {k}_0}$, we have $\bigl (( \mathcal {W} _0)_k,( \mathcal {W} _0)_{k'}\bigr )=0$.

Proof. Assertion (i) follows from the fact that any connected component of ${\sigma ({\mathfrak {g}})}$ is a translation of ${\sigma _0({\mathfrak {g}})}$.

(iii) It is enough to show that for $(i,a)\in ({\sigma _0({\mathfrak {g}})})_k$ and $(j,b)\in ({\sigma _0({\mathfrak {g}})})_{k'}$ we have $( \mathsf {s} _{i,a}, \mathsf {s} _{ j,b })=0$. By the definition of ${\sigma _0({\mathfrak {g}})}$, $V(\varpi _i)_a$ and $\mathscr {D}^mV(\varpi _j)_b$ strongly commute for any $m$, which tells us that

\[ \Lambda^\infty(V(\varpi_i)_a, V(\varpi_j)_b)=0 \]

by Corollary 3.10 and Proposition 3.11.

(ii) It is enough to show that

\[ \Delta_0\cap (\Delta_0)_k=\emptyset \quad\text{for $k\in \mathbf{k}^\times{/} {\mathbf{k}_0} $.}\quad \]

For $(i,a) \in {\sigma _0({\mathfrak {g}})}$ and $(j,b) \in {\sigma _0({\mathfrak {g}})}_k$, we have $( \mathsf {s} _{i,a}, \mathsf {s} _{i,a})= 2$ by Lemma 4.3 and $( \mathsf {s} _{i,a}, \mathsf {s} _{j,b}) = 0$ by (iii). Thus we conclude that $\mathsf {s} _{i,a} \ne \mathsf {s} _{j,b}$.

We set

\[ \mathcal{E} \mathbin{:=} \mathbb{R} \otimes_\mathbb{\mspace{1mu}Z\mspace{1mu}} \mathcal{W} \quad \text{and}\quad \mathcal{E} _0 \mathbin{:=} \mathbb{R} \otimes_\mathbb{\mspace{1mu}Z\mspace{1mu}} \mathcal{W} _0. \]

Then the pairing $(-,-)$ gives a symmetric bilinear form on $\mathcal {E}$. Theorem 4.6 below is the main theorem of this section; its proof is postponed until § 6.

Theorem 4.6

  1. (i) The pair $( \mathcal {E} _0, \Delta _0)$ is an irreducible simply laced root system of type shown in the following table.

    (4.5) \begin{align} \small \begin{array}{|c||c|c|c|c|c|c|c|} \hline \text{Type of ${\mathfrak{g}}$} & A_n^{(1)} & B_n^{(1)} & C_n^{(1)^{\vphantom{1}}} & D_n^{(1)} & A_{2n}^{(2)} & A_{2n-1}^{(2)} & D_{n+1}^{(2)} \\ & (n\geqslant1) & (n\geqslant2) & (n\geqslant3) & (n\geqslant4) & (n\geqslant1) & (n\geqslant2) & (n\geqslant3)\\ \hline \text{Type of $( \mathcal{E} _0, \Delta_0)$} & A_n & A_{2n-1} & D_{n+1} & D_n & A_{2n} & A_{2n-1} & D_{n+1} \\ \hline \hline \text{Type of ${\mathfrak{g}}$} & E_6^{(1)^{\vphantom{1}}} & E_7^{(1)} & E_8^{(1)} & F_4^{(1)} & G_{2}^{(1)} & E_{6}^{(2)} & D_{4}^{(3)} \\ \hline \text{Type of $( \mathcal{E} _0, \Delta_0)$} & E_6 & E_{7} & E_{8} & E_6 & D_{4} & E_{6} & D_{4} \\ \hline \end{array} \end{align}
  2. (ii) The bilinear form $(-,-)\vert _{ \mathcal {W} _0}$ is positive definite. Moreover, it is Weyl group invariant, i.e. $s_\alpha (\Delta _0)\subset \Delta _0$ for any $\alpha \in \Delta _0$. Here $s_\alpha \in \operatorname {End}( \mathcal {E} _0)$ is the reflection defined by $s_\alpha (\lambda )= \lambda -(\alpha,\lambda )\alpha$.

The next corollary follows from Proposition 4.5 and Theorem 4.6.

Corollary 4.7

  1. (i) We have $\mathcal {W} = \bigoplus _{k \in \mathbf {k}^\times / {\mathbf {k}_0} } ( \mathcal {W} _0)_k$.

  2. (ii) As a root system, $( ( \mathcal {E} _0)_k, (\Delta _0)_k)$ is isomorphic to $( \mathcal {E} _0, \Delta _0)$ for $k \in \mathbf {k}^\times / {\mathbf {k}_0}$, and

    \[ ( \mathcal{E} , \Delta) = \bigsqcup_{k \in \mathbf{k}^\times{/} {\mathbf{k}_0} } ( ( \mathcal{E} _0)_k, (\Delta_0)_k). \]

Proof. We know already that $\mathcal {W} = \sum _{k \in \mathbf {k}^\times / {\mathbf {k}_0} } ( \mathcal {W} _0)_k$. Since $( \mathcal {W} _0)_k$ and $( \mathcal {W} _0)_{k'}$ are orthogonal if $k/k'\not \in {\mathbf {k}_0}$, the non-degeneracy of $(-,-)\vert _{ \mathcal {E} }$ implies that $\mathcal {W} = \bigoplus _{k \in \mathbf {k}^\times / {\mathbf {k}_0} } ( \mathcal {W} _0)_k$.

Assertion (ii) easily follows from (i) and Theorem 4.6.

The following corollary is an immediate consequence of Theorem 4.6.

Corollary 4.8 The following hold:

  1. (i) $(\lambda,\lambda )\in 2\mathbb {\mspace {1mu}Z\mspace {1mu}}_{>0}$ for any $\lambda \in \mathcal {W} _0\setminus \{{0}\}$;

  2. (ii) $\Delta _0=\{{\lambda \in \mathcal {W} _0\mid (\lambda,\lambda )=2}\}$.

Hence the root system $( \mathcal {E} _0, \Delta _0)$ is completely determined by the pair $( \mathcal {W} _0,\,(-,-)\vert _{ \mathcal {W} _0})$.

5. Block decomposition of $\mathscr {C}_{\mathfrak {g}}$

In this section, we give a block decomposition of $\mathscr {C}_{\mathfrak {g}}$ parameterized by $\mathcal {W}$.

5.1 Blocks

We recall the notion of blocks. Let $\mathcal {C}$ be an abelian category such that any object of $\mathcal {C}$ has finite length.

Definition 5.1 A block $\mathcal {B}$ of $\mathcal {C}$ is a full abelian subcategory with the following properties:

  1. (i) there is a decomposition $\mathcal {C} = \mathcal {B} \oplus \mathcal {C}'$ for some full abelian subcategory $\mathcal {C}'$,

  2. (ii) there is no non-trivial decomposition $\mathcal {B} = \mathcal {B}' \oplus \mathcal {B}''$ with full abelian subcategories $\mathcal {B}'$ and $\mathcal {B}''$.

The following lemma is obvious.

Lemma 5.2 Let $\mathcal {B}$ be a full subcategory of $\mathcal {C}$ satisfying condition (i) in Definition 5.1. Then $\mathcal {B}$ has the following properties:

  1. (i) $\mathcal {B}$ is stable by taking subquotients and extensions;

  2. (ii) for simple objects $S,S' \in \mathcal {C}$ such that $\operatorname {Ext}^1_\mathcal {C}(S,S') \not \simeq 0$, if one of them belongs to $\mathcal {B}$ then so does the other.

Lemma 5.3 Let $X, X' \in \mathcal {C}$. Suppose that $\operatorname {Ext}^{1}_\mathcal {C}( S,S' )=0$ for any simple subquotients $S$ and $S'$ of $X$ and $X'$, respectively. Then we have $\operatorname {Ext}^{1}_\mathcal {C}( X,X' )=0$.

Proof. Let $\ell$ and $\ell '$ be the lengths of $X$ and $X'$, respectively. We use induction on $\ell +\ell '$. If $X$ and $X'$ are simple, then the claimed result is clear by the assumption.

Suppose that $X'$ is not simple. Then there exists an exact sequence $0 \rightarrow M \rightarrow X' \rightarrow N \rightarrow 0$ with a simple $M$. It in turn gives the exact sequence

\[ \operatorname{Ext}^1_\mathcal{C}(X, M) \rightarrow \operatorname{Ext}^1_\mathcal{C}(X, X') \rightarrow \operatorname{Ext}^1_\mathcal{C}(X, N). \]

By the induction hypothesis we have $\operatorname {Ext}^1_\mathcal {C}(X, M) = \operatorname {Ext}^1_\mathcal {C}(X, N) = 0$, which tells us that $\operatorname {Ext}^{1}_\mathcal {C}( X,X' )=0$.

The case where $X$ is not simple can be proved in the same manner.

Lemma 5.4 Let $\mathfrak {c}$ be the set of isomorphism classes of simple objects of $\mathcal {C}$, and let $\mathfrak {c}=\bigsqcup _{a\in A}\mathfrak {c}_a$ be a partition of $\mathfrak {c}$. We assume that

For $a\in A$, let $\mathcal {C}_a$ be the full subcategory of $\mathcal {C}$ consisting of objects $X$ such that any simple subquotient of $X$ belongs to $\mathfrak {c}_a$. Then $\mathcal {C}=\bigoplus _{a\in A}\mathcal {C}_a$.

Proof. It is enough to show that any object $X$ of $\mathcal {C}$ has a decomposition $X\simeq \bigoplus _{a\in A}X_a$ with $X_a\in \mathcal {C}_a$. In order to prove this, we shall argue by induction on the length of $X$. We may assume that X is non-zero. Let us take a subobject $Y$ of $X$ such that $X/Y$ is simple. Then the induction hypothesis implies that $Y = \bigoplus _{a\in A}Y_a$ with $Y_a\in \mathcal {C}_a$.

Take $a_0\in A$ such that $X/Y$ belongs to $\mathfrak {c}_{a_0}$. Then define $Z\in \mathcal {C}$ by the exact sequence

(5.1)\begin{equation} 0 \rightarrow \bigoplus_{a\not=a_0}Y_a \rightarrow X \rightarrow Z \rightarrow 0. \end{equation}

Since we have an exact sequence $0\to Y_{a_0}\to Z\to X/Y\to 0$, $Z$ belongs to $\mathcal {C}_{a_0}$. Then Lemma 5.3 tells us that $\operatorname {Ext}^1(Z,\bigoplus _{a\not =a_0}Y_a )=0$. Hence the exact sequence (5.1) splits, i.e. $X \simeq Z\oplus \bigoplus _{a\not =a_0}Y_a$.

Let $\approx$ be the equivalence relation on the set of isomorphism classes of simple objects of $\mathcal {C}$ generated by the relation $\approx '$ defined as follows: for simple objects $S,S' \in \mathcal {C}$,

\[ [S] \approx' [S'] \quad \text{if and only if } \quad \operatorname{Ext}^1_\mathcal{C}(S,S') \ne 0. \]

Theorem 5.5 Let $A$ be the set of $\approx$-equivalence classes. For $a\in A$, let $\mathcal {C}_a$ be the full subcategory of $\mathcal {C}$ consisting of objects $X$ such that any simple subquotient of $X$ belongs to $a$. Then $\mathcal {C}_a$ is a block, and the category $\mathcal {C}$ has a decomposition $\mathcal {C} = \bigoplus _{a \in A} \mathcal {C}_a$. Moreover, any block of $\mathcal {C}$ is equal to $\mathcal {C}_a$ for some $a$.

Proof. Lemma 5.4 implies the decomposition

\[ \mathcal{C} = \bigoplus_{a \in A} \mathcal{C}_a. \]

Moreover, since $a$ is a $\approx$-equivalence class, there is no non-trivial decomposition of $\mathcal {C}_a$ for any $a \in A$.

The next corollary follows directly from Theorem 5.5.

Corollary 5.6 Let $X$ be an indecomposable object of $\mathcal {C}$. Then $X$ belongs to some block. In particular, all the simple subquotients of $X$ belong to the same block.

5.2 Direct decomposition of $\mathscr {C}_{\mathfrak {g}}$

In this subsection, we shall prove that $\mathscr {C}_{\mathfrak {g}}$ has a decomposition parameterized by elements of $\mathcal {W}$.

Lemma 5.7 For modules $M,N \in \mathscr {C}_{\mathfrak {g}}$, there exists an isomorphism

(5.2)\begin{equation} \Psi\colon \mathbf{k}[z^{\pm1}] \otimes \operatorname{Hom}_{U'_q({\mathfrak{g}})}(N, \mathbf{1}) \otimes \operatorname{Hom}_{U'_q({\mathfrak{g}})}( \mathbf{1}, M) \buildrel \sim \over \longrightarrow \operatorname{Hom}_{ U'_q({\mathfrak{g}})}(N, M_z)\end{equation}

defined by $\Psi (a(z) \otimes f \otimes g ) = a(z) ( g \circ f )$ for $a(z) \in \mathbf {k}[z^{\pm 1}]$, $f \in \operatorname {Hom}_{U'_q({\mathfrak {g}})}(N, \mathbf {1})$ and $g \in \operatorname {Hom}_{U'_q({\mathfrak {g}})}(\mathbf {1}, M)$.

Proof. Note that $\mathbf {k}[z^{\pm 1}] \otimes \operatorname {Hom}_{U'_q({\mathfrak {g}})}( \mathbf {1}, M)\mathop {\xrightarrow {\sim}}\operatorname {Hom}_{U'_q({\mathfrak {g}})}( \mathbf {1}, M_z)$. There is a quotient $N'$ of $N$ which is a direct sum of copies of ${\bf {1}}$ and $\operatorname {Hom}(N',{\bf {1}})\mathop {\xrightarrow {\sim}}\operatorname {Hom}(N,{\bf {1}})$. Since (5.2) for $N'$ is obviously an isomorphism, $\Psi$ is injective.

To prove that $\Psi$ is surjective, we shall decompose a given non-zero $f\colon N \to M_z$ into $N \to {\bf {1}}^{\oplus \ell } \to M_z$ for some $\ell \in \mathbb {\mspace {1mu}Z\mspace {1mu}}_{>0}$. Here $\mathbf {1}^{\oplus \ell }$ is the direct sum of $\ell$ copies of the trivial module $\mathbf {1}$. Without loss of generality, we may assume that $f$ is injective. We set $\operatorname {wt}(N) \mathbin {:=} \{ \lambda \in \mathsf {P}_{\mathrm {cl}\mspace {1mu}} \mid N_\lambda \ne 0 \}$.

If $\operatorname {wt}(N) =\{0\}$, then $N$ should be isomorphic to $\mathbf {1} ^{\oplus \ell }$ for some $\ell \in \mathbb {\mspace {1mu}Z\mspace {1mu}}_{>0}$, which is the desired result.

Now suppose that $\operatorname {wt}(N)\not =\{0\}$. We choose a non-zero weight $\lambda \in \operatorname {wt}(N)$.

Note that the $U'_q({\mathfrak {g}})$-module structure on $M_z$ extends to a $U_q({\mathfrak {g}})$-module structure and we have a weight decomposition $M_z=\bigoplus _{\mu \in \mathsf {P}}(M_z)_\mu$. Then

\[ f(N_\lambda)\subset \bigoplus_{\mu\in\mathsf{P},\,{\mathrm{cl}\mspace{1mu}}(\mu)=\lambda}(M_z)_\mu, \]

where ${\mathrm {cl}\mspace {1mu}}\colon \mathsf {P}\to \mathsf {P}_{\mathrm {cl}\mspace {1mu}}$ is the classical projection. There exist $w \in \mathsf {W}$ and a non-zero integer $n$ such that $w( \mu ) = \mu +n\delta$ for any $\mu \in {\mathrm {cl}\mspace {1mu}}^{-1}(\lambda )$. We now consider the braid group action $T_w$ defined by $w$ on an integral module (see [Reference LusztigLus90, Reference SaitoSai94]). Then the $\mathbf {k}$-linear automorphism $T_w$ sends $(M_z)_\mu$ to $(M_z)_{w\mu }$. The space $f(N_\lambda )$ is invariant under the automorphism $T_w$, but any non-zero finite-dimensional subspace of $\bigoplus _{\mu \in \mathsf {P},\,{\mathrm {cl}\mspace {1mu}}(\mu )=\lambda } (M_z)_{\mu }$ cannot be invariant under $T_w$. This is a contradiction.

Proposition 5.8 For modules $M,N \in \mathscr {C}_{\mathfrak {g}}$ and a simple module $L \in \mathscr {C}_{\mathfrak {g}}$, we have the isomorphisms

\[ \mathbf{k}[z^{\pm1}] \otimes \operatorname{Hom}_{U_q'({\mathfrak{g}})}(M,N) \buildrel \sim \over \longrightarrow \operatorname{Hom}_{ \mathbf{k}[z^{\pm1}] \otimes U_q'({\mathfrak{g}})}(M\otimes L_z,N \otimes L_z). \]

Proof. By Lemma 5.7, we obtain that

\begin{align*} \operatorname{Hom}_{ \mathbf{k}[z^{\pm1}] \otimes U_q'({\mathfrak{g}}) }(M\otimes L_z,N \otimes L_z) & \simeq \operatorname{Hom}_{U_q'({\mathfrak{g}}) }(N^* \otimes M , (L \otimes L^*)_z) \\ & \simeq {{\mathbf{k}}}[z^{\pm1}] \otimes \operatorname{Hom}_{U_q'({\mathfrak{g}})}(N^* \otimes M ,{\bf{1}}) \mathop\otimes \operatorname{Hom}_{U_q'({\mathfrak{g}})}({\bf{1}}, L\otimes L^*)\\ & \simeq {{\mathbf{k}}}[z^{\pm1}] \otimes \operatorname{Hom}_{U_q'({\mathfrak{g}})}(M ,N). \end{align*}

Lemma 5.9 Let $M$ and $N$ be simple modules in $\mathscr {C}_{\mathfrak {g}}$. If

\[ \frac{c_{M,L}(z)}{c_{N,L}(z)} \notin\mathbf{k}(z) \quad\text{for some simple module } L \in \mathscr{C}_{\mathfrak{g}}, \]

then we have

\[ \operatorname{Ext}^1_{U_q'({\mathfrak{g}})}(M,N)=0. \]

Proof. Let $L \in \mathscr {C}_{\mathfrak {g}}$ be a simple module such that ${c_{M,L}(z)}/{c_{N,L}(z)} \notin \mathbf {k}(z)$.

We shall prove that any exact sequence

\[ 0 \rightarrow N \rightarrow X \rightarrow M \rightarrow 0 \]

splits. We set $\hat {L}_z \mathbin {:=} \mathbf {k}((z)) \otimes _{\mathbf {k} [z^{\pm 1}] } L_z$, where $L_z$ is the affinization of $L$. Then the following diagram commutes.

We set

\[ f(z) \mathbin{:=} \frac{ c_{M, L}(z) }{ c_{N, L}(z) } \notin \mathbf{k}(z) \quad\text{{and}}\quad R \mathbin{:=} c_{M, L}(z) R^{\mspace{1mu}\mathrm{univ}}_{X, \hat{L}_z}\colon X\otimes \hat{L}_z \rightarrow \hat{L}_z \otimes X. \]

It follows from

\[ c_{M,L}(z) R^{\mspace{1mu}\mathrm{univ}}_{M, \hat{L}_z}( M \otimes L_z) \subset L_z \otimes M \quad\text{{and}}\quad c_{N,L}(z) R^{\mspace{1mu}\mathrm{univ}}_{N, \hat{L}_z}( N \otimes L_z) \subset L_z \otimes N \]

that

\[ R( X \otimes L_z) \subset L_z \otimes X + \hat{L}_z \otimes N \quad\text{{and}}\quad R(N \otimes L_z ) \subset f(z) ( L_z \otimes N ). \]

Therefore $R$ induces the $\mathbf {k}[z^{\pm 1}]\mathop \otimes U'_q({\mathfrak {g}})$-linear homomorphism

\[ \mathcal{R} \colon M \otimes L_z\simeq \frac{X\mathop\otimes L_z}{N\mathop\otimes L_z} \longrightarrow \frac{\mathbf{k}(z) \otimes L_z \otimes X + \hat{L}_z \otimes N}{\mathbf{k}(z) \otimes L_z \otimes X + f(z) \mathbf{k}(z) \otimes L_z \otimes N}. \]

We set $\mathcal {P} \mathbin {:=} {\mathbf {k}((z))}/({\mathbf {k}(z) + f(z) \mathbf {k}(z)})$. Since

\begin{align*} \frac{ \mathbf{k}(z) \otimes L_z \otimes X + \hat{L}_z \otimes N}{ \mathbf{k}(z) \otimes L_z \otimes X + f(z) \mathbf{k}(z) \otimes L_z \otimes N } & \simeq \frac{\hat{L}_z \otimes N}{\mathbf{k}(z) \otimes L_z \otimes N + f(z) \mathbf{k}(z) \otimes L_z \otimes N }\\ & \simeq \mathcal{P} \otimes_{\mathbf{k}[z^{\pm1}]} L_z \otimes N, \end{align*}

we have the homomorphism of $\mathbf {k}[z^{\pm 1}] \mathop \otimes U'_q({\mathfrak {g}})$-modules

\[ \mathcal{R} \colon M \otimes L_z \longrightarrow \mathcal{P} \otimes_{\mathbf{k}[z^{\pm1}]} L_z \otimes N. \]

Let us show that $\mathcal {R}$ vanishes.

Assume that $\mathcal {R}\not =0$. Then

\[ \operatorname{Hom}_{\mathbf{k}[z^{\pm1}]\mathop \otimes U'_q({\mathfrak{g}})}(M\mathop\otimes L_z,\mathcal{P} \otimes_{\mathbf{k}[z^{\pm1}]} L_z \otimes N) \simeq \mathcal{P} \otimes_{\mathbf{k}[z^{\pm1}]} \operatorname{Hom}_{\mathbf{k}[z^{\pm1}]\mathop \otimes U'_q({\mathfrak{g}})} (M\mathop\otimes L_z,L_z \otimes N) \]

implies that $\operatorname {Hom}_{\mathbf {k}[z^{\pm 1}]\mathop \otimes U'_q({\mathfrak {g}})}(M\mathop \otimes L_z,L_z \otimes N)\not \simeq 0$.

Since $\mathbf {k}(z)\mathop \otimes _{\mathbf {k}[z^{\pm 1}]}(M\mathop \otimes L_z)$ and $\mathbf {k}(z)\mathop \otimes _{\mathbf {k}[z^{\pm 1}]}(L_z \otimes N)$ are simple $\mathbf {k}(z)\mathop \otimes U'_q({\mathfrak {g}})$-modules, they are isomorphic. Since $\mathbf {k}(z)\mathop \otimes _{\mathbf {k}[z^{\pm 1}]}(L_z \otimes N)$ and $\mathbf {k}(z)\mathop \otimes _{\mathbf {k}[z^{\pm 1}]}(N\mathop \otimes L_z)$ are isomorphic, we conclude that $\mathbf {k}(z)\mathop \otimes _{\mathbf {k}[z^{\pm 1}]}(M\mathop \otimes L_z)\simeq \mathbf {k}(z)\mathop \otimes _{\mathbf {k}[z^{\pm 1}]}(N\mathop \otimes L_z)$. On the other hand, Proposition 5.8 implies that

\[ \mathbf{k}(z) \otimes \operatorname{Hom}_{U_q'({\mathfrak{g}})}(M,N) \mathop{\xrightarrow{\sim }} \operatorname{Hom}_{\mathbf{k}(z)\otimes U_q'({\mathfrak{g}})}(\mathbf{k}(z) \otimes_{\mathbf{k}[z^{\pm1}]}M\otimes L_z, \mathbf{k}(z) \otimes_{\mathbf{k}[z^{\pm1}]} N \otimes L_z). \]

Hence $\operatorname {Hom}_{U_q'({\mathfrak {g}})}(M,N) \not =0$, and we obtain that $M$ and $N$ are isomorphic, which is a contradiction. Therefore $\mathcal {R}=0$, which means that

\[ R\bigl(\mathbf{k}(z)\mathop\otimes (X\mathop\otimes L_z)\bigr)\subset \mathbf{k}(z) \otimes L_z \otimes X + f(z) \mathbf{k}(z) \otimes L_z \otimes N. \]

Let us consider the composition

We have

\[ R\bigl(\mathbf{k}(z)\mathop\otimes (X\mathop\otimes L_z)\bigr)\cap \hat{L}_z \mathop\otimes N =R\bigl( \mathbf{k}(z)\mathop\otimes (N\mathop\otimes L_z)\bigr)=f(z)\mathbf{k}(z)\mathop\otimes L_z\mathop\otimes N. \]

Hence $\ker (\Phi )=K\cap \bigl (\mathbf {k}(z) \otimes L_z \otimes N\bigr ) =\bigl ( f(z)\mathbf {k}(z) \otimes L_z \otimes N)\cap \bigl (\mathbf {k}(z) \otimes L_z \otimes N\bigr )$ vanishes, which means that $\Phi$ is a monomorphism.

Since $\mathbf {k}(z) \otimes L_z\mathop \otimes M$ and $\mathbf {k}(z) \otimes L_z\mathop \otimes N$ are simple $\mathbf {k}(z)\mathop \otimes U'_q({\mathfrak {g}})$-modules, $\mathbf {k}(z) \otimes L_z \otimes X$ has length $2$. Similarly, $R\bigl (\mathbf {k}(z)\mathop \otimes (X\mathop \otimes L_z)\bigr )$ also has length $2$. On the other hand, $\mathbf {k}(z) \otimes L_z \otimes X + f(z) \mathbf {k}(z) \otimes L_z \otimes N$ has length no more than $3$, which implies that $K$ does not vanish. Hence $\Phi$ is an isomorphism. Thus we conclude that the homomorphism

\begin{align*} \operatorname{Hom}(\mathbf{k}(z)\mathop\otimes L_z\mathop\otimes M,\mathbf{k}(z)\mathop\otimes L_z\mathop\otimes X)\to& \operatorname{Hom}(\mathbf{k}(z)\mathop\otimes L_z\mathop\otimes M,\mathbf{k}(z)\mathop\otimes L_z\mathop\otimes M)\\ &=\mathbf{k}(z)\operatorname{id}_{\mathbf{k}(z)\mathop\otimes L_z\mathop\otimes M} \end{align*}

is surjective. Then Proposition 5.8 implies that this homomorphism is isomorphic to

\[ \mathbf{k}(z)\mathop\otimes \operatorname{Hom}(M,X)\twoheadrightarrow \mathbf{k}(z)\mathop\otimes \operatorname{Hom}(M,M). \]

Thus we conclude that $\operatorname {Hom}(M,X)\to \operatorname {Hom}(M,M)$ is surjective, that is,

\[ 0\xrightarrow{\,{\hspace{2ex}}\,} N\xrightarrow{\,{\hspace{2ex}}\,} X\xrightarrow{\,{\hspace{2ex}}\,} M\xrightarrow{\,{\hspace{2ex}}\,}0 \]

splits.

For $\alpha \in \mathcal {W}$, let $\mathscr {C}_{{\mathfrak {g}}, \alpha }$ be the full subcategory of $\mathscr {C}_{\mathfrak {g}}$ consisting of objects $X$ such that $\mathsf {E} (S)=\alpha$ for any simple subquotient $S$ of $X$.

Theorem 5.10 There exist the decompositions

\[ \mathscr{C}_{\mathfrak{g}} = \bigoplus_{\alpha \in \mathcal{W} } \mathscr{C}_{{\mathfrak{g}}, \alpha} \quad \text{and} \quad \mathscr{C}_{\mathfrak{g}}^0 = \bigoplus_{\alpha \in \mathcal{W} _0} \mathscr{C}_{{\mathfrak{g}}, \alpha}. \]

Proof. Let $\alpha, \beta \in \mathcal {W}$ with $\alpha \ne \beta$. For simple modules $M \in \mathscr {C}_{{\mathfrak {g}}, \alpha }$ and $N \in \mathscr {C}_{{\mathfrak {g}}, \beta }$, Lemma 4.1(iii) says that ${a_{M,V(\varpi _i)}(z)}/{a_{N,V(\varpi _i)}(z)} \notin \mathbf {k}(z)$ for some $i\in I_0$. Hence Lemma 5.9 implies that $\operatorname {Ext}^1_{U_q'({\mathfrak {g}})}(M,N)=0.$ The desired result then follows from Lemma 5.4.

5.3 The block $\mathscr {C}_{{\mathfrak {g}}, \alpha }$

Recall the automorphism $\tau _t$ on ${\sigma ({\mathfrak {g}})}$ defined in (4.4). For $(i,a) \in {\sigma ({\mathfrak {g}})}$ we write

\[ V(i,a) \mathbin{:=} V(\varpi_i)_a. \]

Note that $V(\tau _t \alpha ) = V(\alpha )_t$ for $\alpha \in {\sigma ({\mathfrak {g}})}$ and $t\in \mathbf {k}^\times$. For $\alpha \in {\sigma ({\mathfrak {g}})}$, we define $\alpha ^* \in {\sigma ({\mathfrak {g}})}$ by

\[ V(\alpha^*) \simeq V(\alpha)^*. \]

Thus we have

\[ \alpha^{**} = \tau_{\tilde{p}^{{-}1}} (\alpha) \quad \text{for } \alpha \in {\sigma({\mathfrak{g}})}. \]

Lemma 5.11 Let $\alpha _1, \ldots, \alpha _k \in {\sigma ({\mathfrak {g}})}$ for $k\in \mathbb {\mspace {1mu}Z\mspace {1mu}}_{>0}$. Then all the simple subquotients of $V(\alpha _1) \otimes V(\alpha _2) \otimes \cdots \otimes V(\alpha _k)$ are contained in the same block of $\mathscr {C}_{\mathfrak {g}}$.

Proof. There exists a permutation $\sigma \in \mathfrak {S}_k$ such that the tensor product $V(\alpha _{\sigma (1)}) \otimes V(\alpha _{\sigma (2)}) \otimes \cdots \otimes V(\alpha _{\sigma (k)})$ has a simple head by Theorem 2.2, and hence it is indecomposable. Thus, all the simple subquotients of $V(\alpha _{\sigma (1)}) \otimes V(\alpha _{\sigma (2)}) \otimes \cdots \otimes V(\alpha _{\sigma (k)})$ are contained in the same block by Corollary 5.6. Since any simple subquotient of $V(\alpha _1) \otimes V(\alpha _2) \otimes \cdots \otimes V(\alpha _k)$ is isomorphic to some simple subquotient of $V(\alpha _{\sigma (1)}) \otimes V(\alpha _{\sigma (2)}) \otimes \cdots \otimes V(\alpha _{\sigma (k)})$, we obtain the desired result.

We set

\[ \mathcal{P} \mathbin{:=} \bigoplus_{\alpha \in {\sigma({\mathfrak{g}})}} \mathbb{\mspace{1mu}Z\mspace{1mu}} \mathsf{e}_\alpha, \quad \mathcal{P}_0 \mathbin{:=} \bigoplus_{\alpha \in {\sigma_0({\mathfrak{g}})}} \mathbb{\mspace{1mu}Z\mspace{1mu}} \mathsf{e}_\alpha \]

and

\[ \mathcal{P}^+ \mathbin{:=} \sum_{\alpha \in {\sigma({\mathfrak{g}})}} \mathbb{\mspace{1mu}Z\mspace{1mu}}_{\geqslant0} \mathsf{e}_\alpha\subset\mathcal{P}, \]

where $\mathsf {e}_\alpha$ is a symbol. Define a group homomorphism

\[ p\colon \mathcal{P} \twoheadrightarrow \mathcal{W} , \quad \mathsf{e}_{(i,a)} \mapsto \mathsf{s} _{i,a}, \]

and set

\[ p_0 \mathbin{:=} p|_{\mathcal{P}_0}\colon \mathcal{P}_0 \twoheadrightarrow \mathcal{W} _0. \]

By Proposition 4.5, we have

(5.3)\begin{equation} \mathcal{P} = \bigoplus_{k \in \mathbf{k}^\times{/} {\mathbf{k}_0} } (\mathcal{P}_0)_k. \end{equation}

Let $\mathcal {Q}_0$ be the subgroup of $\mathcal {P}_0$ generated by elements of the form $\sum _{k=1}^m \mathsf {e}_{\alpha _k}$ ($\alpha _k \in {\sigma _0({\mathfrak {g}})}$) such that the trivial module $\mathbf {1}$ appears in $V(\alpha _1) \otimes V(\alpha _2) \otimes \cdots \otimes V(\alpha _{ m })$ as a simple subquotient. We then have $p_0 ( \mathcal {Q}_0 ) = 0$.

We set

(5.4)\begin{equation} \mathcal{Q} \mathbin{:=} \bigoplus_{k \in \mathbf{k}^\times{/} {\mathbf{k}_0} } (\mathcal{Q}_0)_k. \end{equation}

Recall $\phi _Q\colon {\Delta ^+_{Q}} \mathop {\xrightarrow {\sim}}{\sigma _Q({\mathfrak {g}})}$ in (2.11). Let ${\Pi _{Q}}\subset {\Delta ^+_{Q}}$ be the set of simple roots of the positive root system ${\Delta ^+_{Q}}$ and $\mathsf {Q}_Q$ the corresponding root lattice. Hence we have ${\Pi _{Q}}\subset {\Delta ^+_{Q}}\subset \mathsf {Q}_Q$.

In the proof of the following lemma, we do not use Theorem 4.6.

Lemma 5.12 For $\alpha \in {\sigma _0({\mathfrak {g}})}$, denote by $\bar {\mathsf {e}}_\alpha \in \mathcal {P}_0/\mathcal {Q}_0$ the image of $\mathsf {e}_\alpha$ under the projection $\mathcal {P}_0\to \mathcal {P}_0/\mathcal {Q}_0$.

  1. (i) The map ${\Delta ^+_{Q}}\ni \alpha \mapsto \bar {\mathsf {e}}_{\phi _Q(\alpha )}\in \mathcal {P}_0/\mathcal {Q}_0$ extends to an additive map $\psi '_Q\colon \mathsf {Q}_Q\to \mathcal {P}_0/\mathcal {Q}_0$.

  2. (ii) We have that $\psi '_Q$ is surjective, i.e.

    \[ \mathcal{P}_0/\mathcal{Q}_0=\sum_{\beta\in{\Pi_{Q}}}\mathbb{\mspace{1mu}Z\mspace{1mu}}\bar{\mathsf{e}}_{\phi_Q(\beta)}. \]
  3. (iii) Let $\psi _Q\colon \mathsf {Q}_Q\to \mathcal {W} _0$ be the composition $\mathsf {Q}_Q\xrightarrow {\,{\psi '_Q}\,}\mathcal {P}_0/\mathcal {Q}_0\xrightarrow {\,{\hspace {2ex}}\,} \mathcal {W} _0$. Then

    \[ \psi_Q(\beta)= \mathsf{E} \bigl( V_Q(\beta)\bigr). \]
  4. (iv) We have that $\psi _Q$ is surjective, i.e. $\mathcal {W} _0=\sum _{\alpha \in \phi _Q({\Pi _{Q}})}\mathbb {\mspace {1mu}Z\mspace {1mu}} p_0( \mathsf {e}_\alpha )$.

Proof. (i) The map ${\Pi _{Q}}\ni \alpha \mapsto \bar {\mathsf {e}}_{\phi _Q(\alpha )}\in \mathcal {P}_0/\mathcal {Q}_0$ extends to a linear map $\psi '_Q\colon \mathsf {Q}_Q\to \mathcal {P}_0/\mathcal {Q}_0$. It is enough to show that $\bar {\mathsf {e}}_{\phi _Q(\gamma )}=\psi '_Q(\gamma )$ for any $\gamma \in {\Delta ^+_{Q}}$. Let us show this by induction on the length of $\gamma$. If $\gamma$ is not a simple root, take a minimal pair $(\beta,\beta ')$ of $\gamma$ (see Proposition 2.4). Since $V_Q(\gamma )$ appears as a composition factor of $V_Q(\beta ) \otimes V_Q(\beta ')$ by Proposition 2.4, we have

\[ \bar{\mathsf{e}}_{\phi_Q(\gamma)}=\bar{\mathsf{e}}_{\phi_Q(\beta)}+\bar{\mathsf{e}}_{\phi_Q(\beta')}=\psi'_Q(\beta)+\psi'_Q(\beta') =\psi'_Q(\gamma). \]

Assertion (ii) follows from (i), and (iii) follows from (ii) and a surjective map $\mathcal {P}_0/\mathcal {Q}_0\twoheadrightarrow \mathcal {W} _0$.

In the proof of the following lemma, we use the fact that the rank of $\mathcal {W} _0$ is at least the rank of ${\Delta ^+_{Q}}$ (stated in Theorem 4.6, whose proof is postponed to § 6; see (6.3)).

Lemma 5.13 We have the isomorphisms

\[ \mathcal{P}_0/\mathcal{Q}_0\mathop{\xrightarrow{\sim }} \mathcal{W} _0\quad\text{{and}}\quad\mathcal{P}/\mathcal{Q}\mathop{\xrightarrow{\sim }} \mathcal{W} . \]

Proof. The second isomorphism easily follows from the first isomorphism together with (5.3) and (5.4). So we need only show that $\mathcal {P}_0/\mathcal {Q}_0\twoheadrightarrow \mathcal {W} _0$ is an isomorphism.

Let $r$ be the rank of ${\Delta ^+_{Q}}$. By (6.3), the rank of $\mathcal {W} _0$ is at least $r$. Let us consider a surjective homomorphism

(5.5)\begin{equation} \mathcal{P}_0/\mathcal{Q}_0\twoheadrightarrow \mathcal{W} _0.\end{equation}

By Lemma 5.12, $\mathcal {P}_0/\mathcal {Q}_0$ is generated by $r$ elements. Hence (5.5) is an isomorphism.

For $\lambda = \sum _{t=1}^k \mathsf {e}_{\alpha _t} \in \mathcal {P}^+$, we set

\[ \bar{V}(\lambda) \mathbin{:=}[ V(\alpha_1) \otimes V(\alpha_2) \otimes \cdots \otimes V(\alpha_k) ] \in K(\mathscr{C}_{\mathfrak{g}}). \]

Note that for $\lambda, \mu \in \mathcal {P}^+$, if $\mathbf {1}$ appears in $\bar {V}(\lambda )$ and $\bar {V}(\mu )$, then $\mathbf {1}$ also appears in $\bar {V}(\lambda ) \otimes \bar {V}(\mu )$. Hence any element of $\mathcal {Q}$ can be written as $\lambda - \mu$ with $\lambda, \mu \in \mathcal {P}^+$ such that ${\bf {1}}$ appears in both $\bar {V}(\lambda )$ and $\bar {V}(\mu )$.

Theorem 5.14 For any $\alpha \in \mathcal {W}$, the subcategory $\mathscr {C}_{{\mathfrak {g}}, \alpha }$ is a block of $\mathscr {C}_{\mathfrak {g}}$.

Proof. Let $\alpha \in \mathcal {W}$, and let $S$ and $S'$ be simple modules in $\mathscr {C}_{{\mathfrak {g}}, \alpha }$. We shall show that $S$ and $S'$ belong to the same block.

Thanks to Theorem 2.2(iv), there exist $\lambda, \lambda ' \in \mathcal {P}^+$ such that $S$ appears in $\bar {V}(\lambda )$ and $S'$ appears in $\bar {V}(\lambda ')$. By Lemma 5.13, we have $\lambda - \lambda ' \in \ker p = \mathcal {Q}$. Then there exist $\mu, \mu ' \in \mathcal {P}^+$ that satisfy the following:

  • $\lambda - \lambda ' = \mu ' - \mu$;

  • $\mathbf {1}$ appears in $\bar {V}(\mu )$ and $\bar {V}(\mu ')$.

Thus the following hold:

  1. (a) $\lambda + \mu = \lambda ' + \mu '$, i.e. $\bar {V}(\lambda +\mu ) = \bar {V}(\lambda '+\mu ')$;

  2. (b) $S$ appears in $\bar {V}(\lambda ) \otimes \bar {V}(\mu ) = \bar {V}(\lambda +\mu )$;

  3. (c) $S'$ appears in $\bar {V}(\lambda ') \otimes \bar {V}(\mu ') = \bar {V}(\lambda '+\mu ')$.

This tells us that $S$ and $S'$ belong to the same block by Lemma 5.11.

Combining Theorem 5.10 with Theorem 5.14, we have the following block decomposition.

Corollary 5.15 There exist the block decompositions

\[ \mathscr{C}_{\mathfrak{g}} = \bigoplus_{\beta \in \mathcal{W} } \mathscr{C}_{{\mathfrak{g}}, \beta} \quad \text{and} \quad \mathscr{C}_{\mathfrak{g}}^0 = \bigoplus_{\beta \in \mathcal{W} _0} \mathscr{C}_{{\mathfrak{g}}, \beta}. \]

Remark 5.16 Lemma 5.13 gives a group presentation of $\mathcal {W}$ which parameterizes the block decomposition of $\mathscr {C}_{\mathfrak {g}}$. When ${\mathfrak {g}}$ is of untwisted type, the block decomposition of $\mathscr {C}_{\mathfrak {g}}$ was given in [Reference Chari and MouraCM05] and [Reference Etingof and MouraEM03]. Considering [Reference Chari and MouraCM05] and [Reference Etingof and MouraEM03] in our setting, their results give another group presentation of $\mathcal {W}$. Let us explain more precisely what this means in our setting.

Suppose that ${\mathfrak {g}}$ is of untwisted type. We define

\[ \mathcal{P}_S \mathbin{:=} \bigoplus_{(i,a)\in {\sigma({\mathfrak{g}})},\, i\in S} \mathbb{\mspace{1mu}Z\mspace{1mu}} \mathsf{e}_{(i,a)}, \]

where

\[ S = \begin{cases} \{1\} & \text{if ${\mathfrak{g}}$ is of type $A_n^{(1)}$, $C_n^{(1)}$ or $E_6^{(1)}$,} \\ \{n\} & \text{if ${\mathfrak{g}}$ is of type $B_n^{(1)}$ or $D_n^{(1)}$ ($n$ odd)}, \\ \{ n-1, n \} & \text{if ${\mathfrak{g}}$ is of type $D_n^{(1)}$ ($n$ even)}, \end{cases} \]

and $S$ is $\{2\}$, $\{4\}$, $\{7\}$ or $\{8\}$ if ${\mathfrak {g}}$ is of type $G_2^{(1)}$, $F_4^{(1)}$, $E_7^{(1)}$ or $E_8^{(1)}$, respectively.

One can show that $p( \mathcal {P}_S ) = \mathcal {W}$. Thus we have the surjective homomorphism

\[ p_S \mathbin{:=} p|_{ \mathcal{P}_S}\colon \mathcal{P}_S \twoheadrightarrow \mathcal{W} . \]

Then the results in [Reference Chari and MouraCM05, Proposition 4.1 and Appendix A] and [Reference Etingof and MouraEM03, Lemma 4.6 and § 6] explain that the kernel $\ker (p_S)$ is generated by the subset $G$ described as follows:

  1. (a) if ${\mathfrak {g}}$ is of type $A_n^{(1)}$, then $G = \{ \sum _{k=0}^n \mathsf {e}_{(1, tq^{2k})} \mid t \in \mathbf {k}^\times \}$;

  2. (b) if ${\mathfrak {g}}$ is of type $B_n^{(1)}$, then $G = \{ \mathsf {e}_{(n,t)} + \mathsf {e}_{(n,tq^{2n-1})} \mid t \in \mathbf {k}^\times \}$;

  3. (c) if ${\mathfrak {g}}$ is of type $C_n^{(1)}$, then $G = \{ \mathsf {e}_{(1,t)} + \mathsf {e}_{(1,tq^{ n+1})} \mid t \in \mathbf {k}^\times \}$;

  4. (d) if ${\mathfrak {g}}$ is of type $D_n^{(1)}$ and $n$ is odd, then $G = \{ \mathsf {e}_{n,t} + \mathsf {e}_{n,tq^2} + \mathsf {e}_{n,tq^{2n-2}} + \mathsf {e}_{n,tq^{2n}} \mid t \in \mathbf {k}^\times \}$;

  5. (e) if ${\mathfrak {g}}$ is of type $D_n^{(1)}$ and $n$ is even, then $G = \{ \mathsf {e}_{(n-1,t)} + \mathsf {e}_{(n-1,tq^2)} + \mathsf {e}_{(n,tq^{2n-2})} + \mathsf {e}_{(n,tq^{2n})}, \mathsf {e}_{(n-1,t)} + \mathsf {e}_{(n-1,tq^{2n-2})}, \mathsf {e}_{(n ,t)} + \mathsf {e}_{(n ,tq^{2n-2})} \mid t \in \mathbf {k}^\times \}$;

  6. (f) if ${\mathfrak {g}}$ is of type $E_6^{(1)}$, then $G = \{ \mathsf {e}_{(1,t)} + \mathsf {e}_{(1,tq^{ 8})} + \mathsf {e}_{(1,tq^{ 16})} , \mathsf {e}_{(1,t)} + \mathsf {e}_{(1,tq^{ 2})} + \mathsf {e}_{(1,tq^{ 4})} + \mathsf {e}_{(1,tq^{ 12})} + \mathsf {e}_{(1,tq^{ 14})} + \mathsf {e}_{(1,tq^{ 16})} \mid t \in \mathbf {k}^\times \}$;

  7. (g) if ${\mathfrak {g}}$ is of type $E_7^{(1)}$, then $G = \{ \mathsf {e}_{(7,t)} + \mathsf {e}_{(7,tq^{ 18})}, \mathsf {e}_{(7,t)} + \mathsf {e}_{(7,tq^{ 2})} + \mathsf {e}_{(7,tq^{ 12})} + \mathsf {e}_{(7,tq^{ 14})} + \mathsf {e}_{(7,tq^{ 24})} + \mathsf {e}_{(7,tq^{ 26})} \mid t \in \mathbf {k}^\times \}$;

  8. (h) if ${\mathfrak {g}}$ is of type $E_8^{(1)}$, then $G = \{ \mathsf {e}_{(8,t)} + \mathsf {e}_{(8,tq^{ 30})}, \mathsf {e}_{(8,t)} + \mathsf {e}_{(8,tq^{ 20})} + \mathsf {e}_{(8,tq^{ 40})} ,\ \mathsf {e}_{(8,t)} + \mathsf {e}_{(8,tq^{ 12})} + \mathsf {e}_{(8,tq^{ 24})} + \mathsf {e}_{(8,tq^{ 36})} + \mathsf {e}_{(8,tq^{ 48})} \mid t \in \mathbf {k}^\times \}$;

  9. (i) if ${\mathfrak {g}}$ is of type $F_4^{(1)}$, then $G = \{ \mathsf {e}_{(4,t)} + \mathsf {e}_{(4,tq^{ 9})} , \mathsf {e}_{(4,t)} + \mathsf {e}_{(4,tq^{6})} + \mathsf {e}_{(4,tq^{ 12})} \mid t \in \mathbf {k}^\times \}$;

  10. (j) if ${\mathfrak {g}}$ is of type $G_2^{(1)}$, then $G = \{\mathsf {e}_{(2,t)} + \mathsf {e}_{(2,tq^{ 4})} , \mathsf {e}_{(2,t)} + \mathsf {e}_{(2,{t(-q_t)}^{8})} + \mathsf {e}_{(2,{t(-q_t)}^{ 16}}) \mid t \in \mathbf {k}^\times \}$.

We remark that there are typos in the descriptions for types $E_8$ and $F_4$ in [Reference Chari and MouraCM05, Appendix A].

6. Proof of Theorem 4.6

6.1 Strategy of the proof

We now start to prove Theorem 4.6. We shall use the same notation as in §§ 2.3 and 2.4. Recall the explicit descriptions for ${\sigma _0({\mathfrak {g}})}$ and ${\sigma _Q({\mathfrak {g}})}$. Let ${\Pi _{Q}}=\{{\alpha _i}\}_{i\in {\mathrm {I_{fin}}}}$ be the set of simple roots of ${\Delta ^+_{Q}}$, and let $\mathsf {Q}_Q$ be the root lattice of ${{\mathfrak {g}}_{\mathrm {fin}}}$. Hence

\[ {\Pi_{Q}} \subset{\Delta^+_{Q}}\subset\mathsf{Q}_Q. \]

Then, by Lemma 5.12, we have

(6.1)\begin{equation} \mathcal{W} _0 = \sum_{i\in {\mathrm{I_{fin}}}}\mathbb{\mspace{1mu}Z\mspace{1mu}} \mathsf{s} _{\phi_Q(\alpha_i)}, \end{equation}

where $\phi _Q\colon {\Delta ^+_{Q}}\mathop {\xrightarrow {\sim}}{\sigma _Q({\mathfrak {g}})}$ is the bijection given in (2.11).

Let ${ \mathsf {M}_Q } \mathbin {:=} ( { \mathsf {m}^Q_{i,j} } )_{i,j\in {\mathrm {I_{fin}}} }$ be the square matrix given by

\[ { \mathsf{m}^Q_{i,j} } \mathbin{:=} ( \mathsf{s} _{\phi_Q(\alpha_i)}, \mathsf{s} _{\phi_Q(\alpha_j)} ). \]

Thanks to Lemma 4.3, we know that

\[ { \mathsf{m}^Q_{i,i} } = 2 \quad \text{for any } i\in {\mathrm{I_{fin}}}. \]

To prove Theorem 4.6, it suffices to show that ${ \mathsf {M}_Q }$ is the Cartan matrix of the finite simple Lie algebra ${{\mathfrak {g}}_{\mathrm {fin}}}$, i.e.

(6.2)\begin{equation} ( \mathsf{s} _{\phi_Q(\alpha_i)}, \mathsf{s} _{\phi_Q(\alpha_j)} )=(\alpha_i,\alpha_j).\end{equation}

Indeed, (6.2) implies the following lemma, and Theorem 4.6 is its immediate consequence.

Lemma 6.1 Assume ( 6.2). Then the map ${\Delta ^+_{Q}}\ni \beta \mapsto \mathsf {E} \bigl ( V_Q(\beta )\bigr )\in \Delta _0\subset \mathcal {W} _0$ extends uniquely to an additive isomorphism

\[ \psi_{Q}\colon\mathsf{Q}_Q\mathop{\xrightarrow{\sim }} \mathcal{W} _0. \]

Moreover, it preserves the inner products of $\mathsf {Q}_Q$ and $\mathcal {W} _0$.

Proof. Since the Cartan matrix is a symmetric positive-definite matrix, $\{{ \mathsf {s} _{\phi _Q(\alpha _i)}}\}_{i\in {\mathrm {I_{fin}}}}$ is linearly independent. Hence we obtain that

(6.3)\begin{equation} \text{the rank of $ \mathcal{W} _0$ is at least the rank $r$ of ${{\mathfrak{g}}_{\mathrm{fin}}}$.} \end{equation}

On the other hand, Lemma 5.12 implies that $\psi _Q\colon \mathsf {Q}_Q\to \mathcal {W} _0$ is surjective. Hence $\psi _Q$ is an isomorphism. Moreover, (6.2) shows that $\psi _Q$ preserves the inner products of $\mathsf {Q}_Q$ and $\mathcal {W} _0$. The other assertions then easily follow.

6.2 Calculation of the inner products

In this subsection, we give a type-by-type proof of (6.2).

Lemma 6.2 Suppose that ${\mathfrak {g}}$ is of affine ADE type. Let $i,j \in I_0$.

  1. (i) For $t\in \mathbb {\mspace {1mu}Z\mspace {1mu}}$, we have

    \[ \operatorname{\mathfrak{d}}( V(\varpi_i), V(\varpi_{j})_{ ({-}q)^t} ) = \delta( 2 \leqslant |t| \leqslant h ) \, \tilde{c}_{i,j}(|t|-1), \]
    where $h$ is the Coxeter number of ${\mathfrak {g}}$ and $\tilde {c}_{i,j}(k)$ is the integer defined in (A.1) in Appendix A.
  2. (ii) If $0 < t < 2h$, then we have

    \[ \Lambda^\infty(V(\varpi_i), V(\varpi_j)_{({-}q)^{t}} ) = \tilde{c}_{i,j}(t-1)-\tilde{c}_{i,j}(t+1) \]
    and $\Lambda ^\infty (V(\varpi _i), V(\varpi _j)) = -2 \delta _{i,j}$.

Proof. (i) For $i,j \in I$, we write $d_{i,j}(z) \mathbin {:=} d_{V(\varpi _i), V(\varpi _j)} (z)$. Combining Proposition 3.9 with the denominator formula

\[ d_{i,j}(z) = \prod_{k=1}^{h-1} (z - ({-}q)^{k+1})^{\tilde{c}_{i,j}(k)} \]

given in (A.2), we compute

\begin{align*} \operatorname{\mathfrak{d}}( V(\varpi_i), V(\varpi_{j})_{ ({-}q)^t} ) & = \delta( 2 \leqslant t \leqslant h ) \,\tilde{c}_{i,j}(t-1) + \delta( 2 \leqslant -t \leqslant h ) \,\tilde{c}_{i,j}({-}t-1) \\ &= \delta( 2 \leqslant |t| \leqslant h ) \,\tilde{c}_{i,j}(|t|-1). \end{align*}

(ii) For $a \in \mathbb {\mspace {1mu}Z\mspace {1mu}}$, let $[a]\mathbin {:=}\prod _{n=0}^\infty (1 - (-q)^{a} {\tilde {p}}^n z)$. Combining the equation (A.13) in [Reference Akasaka and KashiwaraAK97] with the denominator formula (A.2), we have

\begin{align*} a_{i,j} (({-}q)^t z) &=\prod_{1\leqslant k \leqslant h-1} \frac{([h+k+1+t]^{\tilde{c}_{j,i^*}(k)}) ( [h-k-1+t]^{\tilde{c}_{j,i^*}(k)} )} { ([k+1+t]^{\tilde{c}_{i,j}(k)}) ([2h-k-1+t]^{\tilde{c}_{i,j}(k)}) } \\ &=\prod_{1\leqslant k \leqslant h-1} \frac{([h+k+1+t]^{-\tilde{c}_{i,j}(h+k)}) ( [h-k-1+t]^{-\tilde{c}_{i,j}(h+k)} )} { ([k+1+t]^{\tilde{c}_{i,j}(k)}) ([2h-k-1+t]^{\tilde{c}_{i,j}(k)}) } \\ &=\prod_{1\leqslant k \leqslant 2h-1} \frac{1}{ ([k+1+t]^{\tilde{c}_{i,j}(k)}) ([2h-k-1+t]^{\tilde{c}_{i,j}(k)}) } \end{align*}

for any $t\in \mathbb {\mspace {1mu}Z\mspace {1mu}}$, up to a constant multiple. For the second equality, we used

\[ \tilde{c}_{i,j}(h+k) ={-}\tilde{c}_{i,j}(h-k)={-}\tilde{c}_{j^*,i}(k) \quad \text{for} \ 1\leqslant k \leqslant h-1, \]

which comes from [Reference FujitaFuj22, Lemma 3.7 (4) and (5)]. Hence we have

\begin{align*} & \Lambda^\infty(V(\varpi_i), V(\varpi_j)_{({-}q)^{t}} ) ={-} \mathrm{Deg}^{\infty} (a_{i,j} (({-}q)^t z) )\\ &\quad =\sum_{1\leqslant k \leqslant 2h-1} \bigl(\, \tilde{c}_{i,j}(k)(\delta(k+1+t\equiv 0 \mspace{3mu}\mathbin{\mathrm{mod}}\mspace{1mu} 2h) + \delta(2h-k-1+t\equiv 0 \mspace{3mu}\mathbin{\mathrm{mod}}\mspace{1mu} 2h) ) \bigr) \\ &\quad = \tilde{c}_{ij}(2h-t-1) +\tilde{c}_{ij}(t-1) \\ &\quad ={-} \tilde{c}_{ij}( t+1) + \tilde{c}_{ij}(t-1) \end{align*}

for $1\leqslant t\leqslant 2h-1$. If $t=0$, then we have

\[ \Lambda^\infty(V(\varpi_i), V(\varpi_j) ) = 2 \tilde{c}_{i,j}(2h-1) ={-} 2\tilde{c}_{i,j}(1) ={-} 2\delta_{i,j}, \]

as desired.

Type $A_n^{(1)}$. If $n=1$, then it is obvious that ${ \mathsf {M}_Q }$ is a Cartan matrix, so we may assume that $n\geqslant 2$. Recall the explicit description of ${\sigma _Q({\mathfrak {g}})}$ for type $A_n^{(1)}$. Note that the Dynkin quiver corresponding to ${\sigma _Q({\mathfrak {g}})}$ is given in (A.3). In this case, $h=n+1$ and

\[ \phi_Q(\alpha_i) = (1, ({-}q)^{ 2 -2i}) \in {\sigma_Q({\mathfrak{g}})} \quad \text{for } i\in {\mathrm{I_{fin}}}= \{{1,\ldots,n}\} \]

by [Reference Kang, Kashiwara and KimKKK15, Lemma 3.2.3]. For example, if it is of type $A_4^{(1)}$, then elements $(i,(-q)^{k})$ of ${\sigma _Q({\mathfrak {g}})}$ with the values of $\phi _Q^{-1}$ can be drawn as follows.

Here $(a_1,a_2,a_3,a_4) \mathbin {:=} \sum _{k=1}^4 a_k\alpha _k \in {\Delta ^+_{Q}}$ is placed at the position $\phi _Q( a_1, a_2, a_3, a_4 )$, and the underlined ones are simple roots. Using the formula given in Appendix A.1, one can compute that $\tilde {c}_{1,1}(2k)=0$ and

\[ \tilde{c}_{1,1}(2k+1) = (\tau^k \alpha_1,\varpi_1) = (\alpha_{k+1},\varpi_1)=\delta_{k,0} \quad \text{for } 0 \leqslant k < n. \]

Lemma 6.2 implies that

\[ \Lambda^\infty(V(\varpi_1), V(\varpi_1)_{({-}q)^{2k}} ) = \delta_{k,1} \quad \text{for $k\in \mathbb{\mspace{1mu}Z\mspace{1mu}}$ with $ 1 \leqslant k \leqslant n-1 $.} \]

Therefore, for $i > j$ we have

\[ { \mathsf{m}^Q_{i,j} } ={-} \Lambda^\infty(V(\varpi_1), V(\varpi_1)_{({-}q)^{2(i-j)}} ) ={-} \delta_{i-j, 1}, \]

which tells us that ${ \mathsf {M}_Q }$ is a Cartan matrix of type $A_n$.

Type $B_n^{(1)}$. Recall the explicit description of ${\sigma _Q({\mathfrak {g}})}$ for type $B_n^{(1)}$ ($n \geqslant 2$), which can be obtained from [Reference Kashiwara and OhKO19]. Note that the Dynkin diagram of $B_2^{(1)}$ is given in (2.1). In this case ${{\mathfrak {g}}_{\mathrm {fin}}}$ is of type $A_{2n-1}$, and for $i\in {\mathrm {I_{fin}}}=\{{1,\ldots, 2n-1}\}$ we have

\[ \phi_Q(\alpha_i) = \begin{cases} (1, ({-}1)^{n+1} q_s^{2n+1-4i}) & \text{if } 1 \leqslant i \leqslant n-1, \\ (n, q^{ {-}2n+2 }) & \text{if } i=n, \\ (n, q^{ {-}2n+3 }) & \text{if } i=n+1, \\ (1, ({-}1)^{n+1} q_s ^{{-}6n+4i-1}) & \text{if } n+2 \leqslant i \leqslant 2n-1. \end{cases} \]

For example, if it is of type $B_3^{(1)}$, then elements of ${\sigma _Q({\mathfrak {g}})}$ with the values of $\phi _Q^{-1}$ can be drawn as follows.

Here we set $(a_1,a_2,a_3,a_4,a_5) \mathbin {:=} \sum _{k=1}^5 a_k\alpha _k \in {\Delta ^+_{Q}}$, and the underlined ones are simple roots. Combining Propositions 3.11 and 3.9 with the denominator formula given in Appendix A, we compute that $\operatorname {\mathfrak {d}} (V_Q(\alpha _i), \mathscr {D}^k V_Q(\alpha _j) ) = 0$ for $i\ne j$ and $k \ne 0$ and that

\begin{align*} \Lambda^\infty(V(\varpi_1), V(\varpi_1)_{q^k}) &= \operatorname{\mathfrak{d}} ( V(\varpi_1), V(\varpi_1)_{q^k} ) \\ & = \delta_{k,2} \quad \text{for } k=1, 2, \ldots, 2n-4,\\ \Lambda^\infty(V(\varpi_n), V(\varpi_1)_{({-}1)^{n+1}q_s^t}) &= \operatorname{\mathfrak{d}} ( V(\varpi_n), V(\varpi_1)_{({-}1)^{n+1}q_s^t} ) \\ & = \delta_{t, 2n+1} \quad \text{for } t=2n-1, 2n+1 , \ldots, 6n-7,\\ \Lambda^\infty(V(\varpi_n), V(\varpi_n)_{q}) &= 1. \end{align*}

Therefore, for $i > j$ we obtain

\[ { \mathsf{m}^Q_{i,j} } ={-} \delta_{i-j, 1}, \]

which tells us that ${ \mathsf {M}_Q }$ is a Cartan matrix of type $A_{2n-1}$.

Type $C_n^{(1)}$. Recall the explicit description of ${\sigma _Q({\mathfrak {g}})}$ for type $C_n^{(1)}$ ($n \geqslant 3$), which can be obtained from [Reference Kashiwara and OhKO19]. In this case ${{\mathfrak {g}}_{\mathrm {fin}}}$ is of type $D_{n+1}$, and for $1 \leqslant i \leqslant n+1$ we have

\[ \phi_Q (\alpha_i) = \begin{cases} (1, ({-}q_s)^{2-2i}) & \text{if } 1 \leqslant i \leqslant n, \\ (n, ({-}q_s)^{{-}3n+1}) & \text{if } i=n+1. \end{cases} \]

For example, if it is of type $C_4^{(1)}$, then elements of ${\sigma _Q({\mathfrak {g}})}$ with the values of $\phi _Q^{-1}$ can be drawn as follows.

Here we set ${\bigl (\begin {smallmatrix}{a_5}\\[3pt]{a_1a_2a_3a_4}\end {smallmatrix}\bigr )} \mathbin {:=} \sum _{k=1}^5 a_k\alpha _k \in {\Delta ^+_{Q}}$, and the underlined ones are simple roots. Combining Propositions 3.11 and 3.9 with the denominator formula given in Appendix A, we compute that $\operatorname {\mathfrak {d}} (V_Q(\alpha _i), \mathscr {D}^k V_Q(\alpha _j) ) = 0$ for $i\ne j$ and $k \ne 0$ and that

\begin{align*} \Lambda^\infty(V(\varpi_1), V(\varpi_1)_{({-}q_s)^k}) & = \operatorname{\mathfrak{d}} ( V(\varpi_1), V(\varpi_1)_{({-}q_s)^k} ) \\ & = \delta_{k,2} \quad \text{for } k=2, 4, \ldots, 2n-2,\\ \Lambda^\infty(V(\varpi_n), V(\varpi_1)_{({-}q_s)^t}) & = \operatorname{\mathfrak{d}} ( V(\varpi_n), V(\varpi_1)_{({-}q_s)^t} ) \\ & = \delta_{t, n+3} \quad \text{for } t=n+1, n+3, \ldots, 3n-1. \end{align*}

Therefore, for $i > j$ we have

\[ { \mathsf{m}^Q_{i,j} } = \begin{cases} -1 & \text{if ($i \leqslant n$ and $i-j=1$) or $(i,j)= (n+1,n-1) $}, \\ 0 & \text{otherwise,} \end{cases} \]

which says that ${ \mathsf {M}_Q }$ is a Cartan matrix of type $D_{ n+1}$.

Type $D_n^{(1)}$. Recall the explicit description of ${\sigma _Q({\mathfrak {g}})}$ for type $D_n^{(1)}$ ($n \geqslant 4$). Note that the Dynkin quiver corresponding to ${\sigma ({\mathfrak {g}})}_Q$ is given in (A.3). In this case $h=2n-2$, and for $1 \leqslant i \leqslant n$ we have

\[ \phi_Q(\alpha_i) = \begin{cases} (1, ({-}q)^{{-}2(i-1)}) & \text{if } i \leqslant n-2, \\ (n-1, ({-}q)^{ {-}3n+6 }) & \text{if ($i = n-1$ and $n$ is even) or ($i = n$ and $n$ is odd)}, \\ (n, ({-}q)^{ {-}3n+6 }) & \text{if ($i = n $ and $n$ is even) or ($i = n-1$ and $n$ is odd)} \end{cases} \]

by [Reference Kang, Kashiwara and KimKKK15, Lemma 3.2.3]. For example, if it is of type $D_5^{(1)}$, then elements $(i, (-q)^k)$ of ${\sigma _Q({\mathfrak {g}})}$ with the values of $\phi _Q^{-1}$ can be drawn as follows.

Here we set ${\bigl (\begin {smallmatrix}{a_5}\\[3pt]{a_1a_2a_3a_4}\end {smallmatrix}\bigr )} \mathbin {:=} \sum _{k=1}^5 a_k\alpha _k \in {\Delta ^+_{Q}}$, and the underlined ones are simple roots. Using the formula given in Appendix A.1, one can compute that for $1 \leqslant k < h$,

\begin{align*} & \tilde{c}_{1,1}(k) = \delta_{k,1} + \delta_{k,2n-3}, \quad \tilde{c}_{n,1}(k)=\tilde{c}_{n-1,1}(k) = \delta_{k,n-1} , \\ & \tilde{c}_{n,n}(k) = \tilde{c}_{n-1,n-1}(k) = \delta( k\equiv 1 \mspace{3mu}\mathbin{\mathrm{mod}}\mspace{1mu} 4 ), \\ & \tilde{c}_{n,n-1}(k) = \tilde{c}_{n-1,n}(k) = \delta( k\equiv 3 \mspace{3mu}\mathbin{\mathrm{mod}}\mspace{1mu} 4 ). \end{align*}

Combining this with Lemma 6.2, we compute that

\begin{align*} \Lambda^\infty(V(\varpi_1), V(\varpi_1)_{({-}q)^{k}} ) & = \delta_{k,2} \quad \text{for $ 2 \leqslant k \leqslant h-4 $}, \\ \Lambda^\infty(V(\varpi_n), V(\varpi_1)_{({-}q)^{k}} ) & = \delta_{k,n} \quad \text{for $ n \leqslant k \leqslant 3n-6 $}, \\ \Lambda^\infty(V(\varpi_n), V(\varpi_{n-1})) & = 0. \end{align*}

Therefore, for $i > j$ we have

\[ { \mathsf{m}^Q_{i,j} } = \begin{cases} -1 & \text{if ($i \leqslant n-1$ and $i-j=1$) or $(i,j)= (n,n-2) $}, \\ 0 & \text{otherwise,} \end{cases} \]

which says that ${ \mathsf {M}_Q }$ is a Cartan matrix of type $D_n$.

Type $A_{2n}^{(2)}$. Recall the explicit description of ${\sigma _Q({\mathfrak {g}})}$ for type $A_{2n}^{(2)}$ ($n \geqslant 1$), which can be obtained from [Reference Kang, Kashiwara, Kim and OhKKKO16]. In this case ${{\mathfrak {g}}_{\mathrm {fin}}}$ is of type $A_{2n}$, and for $1 \leqslant i \leqslant 2n$ we have

\[ \phi_Q(\alpha_i) = (1, ({-}q)^{2-2i}). \]

For example, if it is of type $A_4^{(2)}$, then elements of ${\sigma _Q({\mathfrak {g}})}$ with the values of $\phi _Q^{-1}$ can be drawn as follows.

Here $(a_1,a_2,a_3,a_4) \mathbin {:=} \sum _{k=1}^4 a_k\alpha _k \in {\Delta ^+_{Q}}$, and the underlined ones are simple roots. It follows from Propositions 3.11 and 3.9 and the denominator formula in Appendix A that $\operatorname {\mathfrak {d}} (V_Q(\alpha _i), \mathscr {D}^k V_Q(\alpha _j) ) = 0$ for $i\ne j$ and $k \ne 0$ and that

\begin{align*} \Lambda^\infty(V(\varpi_1), V(\varpi_1)_{({-}q)^k}) & = \operatorname{\mathfrak{d}} ( V(\varpi_1), V(\varpi_1)_{({-}q )^k} ) \\ & = \delta_{k,2} \quad \text{for } k=2, 4, \ldots, 4n-2. \end{align*}

Therefore, for $i > j$ we have

\[ { \mathsf{m}^Q_{i,j} } ={-} \Lambda^\infty(V(\varpi_1), V(\varpi_1)_{({-}q)^{2(i-j)}} ) ={-} \delta_{i-j, 1}, \]

which tells us that ${ \mathsf {M}_Q }$ is a Cartan matrix of type $A_{2n}$.

Type $A_{2n-1}^{(2)}$. Recall the explicit description of ${\sigma _Q({\mathfrak {g}})}$ for type $A_{2n-1}^{(2)}$ ($n \geqslant 2$), which can be obtained from [Reference Kang, Kashiwara, Kim and OhKKKO16]. In this case ${{\mathfrak {g}}_{\mathrm {fin}}}$ is of type $A_{2n-1}$, and for $1 \leqslant i \leqslant 2n-1$ we have

\[ \phi_Q (\alpha_i) = (1, ({-}q)^{2-2i}). \]

For example, if it is of type $A_5^{(2)}$, then elements of ${\sigma _Q({\mathfrak {g}})}$ with the values of $\phi _Q ^{-1}$ can be drawn as follows.

Here $(a_1,a_2,a_3,a_4,a_5) \mathbin {:=} \sum _{k=1}^5 a_k\alpha _k \in {\Delta ^+_{Q}}$, and the underlined ones are simple roots. Note that $V(\varpi _n)_{a} \simeq V(\varpi _n)_{-a}$. It follows from Propositions 3.11 and 3.9 and the denominator formula in Appendix A that $\operatorname {\mathfrak {d}} (V_Q(\alpha _i), \mathscr {D}^k V_Q(\alpha _j) ) = 0$ for $i\ne j$ and $k \ne 0$ and that

\begin{align*} \Lambda^\infty(V(\varpi_1), V(\varpi_1)_{({-}q)^k}) & = \operatorname{\mathfrak{d}} ( V(\varpi_1), V(\varpi_1)_{({-}q )^k} ) \\ & = \delta_{k,2} \quad \text{for } k=2, 4, \ldots, 4n-4. \end{align*}

Thus we obtain

\[ { \mathsf{m}^Q_{i,j} } ={-} \Lambda^\infty(V(\varpi_1), V(\varpi_1)_{({-}q)^{2(i-j)}} ) ={-} \delta_{i-j, 1}, \quad \text{for }i > j, \]

which implies that ${ \mathsf {M}_Q }$ is a Cartan matrix of type $A_{2n-1}$.

Type $D_{n+1}^{(2)}$. Recall the explicit description of ${\sigma _Q({\mathfrak {g}})}$ for type $D_{n+1}^{(2)}$ ($n \geqslant 3$), which can be obtained from [Reference Kang, Kashiwara, Kim and OhKKKO16]. In this case ${{\mathfrak {g}}_{\mathrm {fin}}}$ is of type $D_{ n+1}$, and for $1 \leqslant i \leqslant n+1$ we have

\[ \phi_Q (\alpha_i) = \begin{cases} (1, (\sqrt{-1})^{n} ({-}q)^{{-}2(i-1)}) & \text{if } i \leqslant n-1, \\ ( n, ({-}1)^i ({-}q)^{ {-}3n+3 }) & \text{if } i=n, n+1. \end{cases} \]

For example, if it is of type $D_5^{(2)}$, then elements of ${\sigma _Q({\mathfrak {g}})}$ with the values of $\phi _Q^{-1}$ can be drawn as follows.

Here we set ${\bigl (\begin {smallmatrix}{a_5}\\[3pt]{a_1a_2a_3a_4}\end {smallmatrix}\bigr )} \mathbin {:=} \sum _{k=1}^5 a_k\alpha _k \in {\Delta ^+_{Q}}$, and the underlined ones are simple roots. Note that $V(\varpi _i)_{a} \simeq V(\varpi _i)_{-a}$ for $i < n$. It follows from Propositions 3.11 and 3.9 and the denominator formula in Appendix A that $\operatorname {\mathfrak {d}} (V_Q(\alpha _i), \mathscr {D}^k V_Q(\alpha _j) ) = 0$ for $i\ne j$ and $k \ne 0$ and that

\begin{align*} \Lambda^\infty(V(\varpi_1), V(\varpi_1)_{({-}q)^k}) & = \operatorname{\mathfrak{d}} ( V(\varpi_1), V(\varpi_1)_{({-}q)^k} ) \\ & = \delta_{k,2} \quad \text{for } k=2, 4, \ldots, 2n-4, \\ \Lambda^\infty(V(\varpi_n), V(\varpi_1)_{{\pm} \sqrt{-1}^{n } ({-}q)^{k}}) & = \operatorname{\mathfrak{d}} ( V(\varpi_{n }), V(\varpi_1)_{{\pm} \sqrt{-1}^{n } ({-}q)^{k}} ) \\ & = \delta_{k,n+1} \quad \text{for } k=n+1, n+3, \ldots, 3n-3,\\ \Lambda^\infty(V(\varpi_n), V(\varpi_n)_{{-}1}) & = 0, \end{align*}

which give the values of ${ \mathsf {m}^Q_{i,j} }$. Thus, one can check that the matrix ${ \mathsf {M}_Q }$ is a Cartan matrix of type $D_{n+1}$.

Type $E_6^{(1)}$. Recall the explicit description of ${\sigma _Q({\mathfrak {g}})}$ for type $E_6^{(1)}$. The Dynkin quiver corresponding to ${\sigma ({\mathfrak {g}})}_Q$ is given in (A.3). In this case, $h=12$ and elements $(i, (-q)^k)$ of ${\sigma _Q({\mathfrak {g}})}$ with the values of $\phi _Q^{-1}$ can be drawn as follows.

Here we set ${\bigl (\begin {smallmatrix}{a_1a_2a_3}\\[3pt]{a_4a_5a_6}\end {smallmatrix}\bigr )} \mathbin {:=} \sum _{i=1}^6 a_i \alpha _i \in {\Delta ^+_{Q}}$, and the underlined ones are simple roots. Using the formula given in Appendix A.1, one can compute that for $1 \leqslant k < h$,

\[ \tilde{c}_{1,1}(k) = \delta_{k,1} + \delta_{k,7}, \quad \tilde{c}_{1,2}(k) = \delta_{k,4} + \delta_{k,8}. \]

By Lemma 6.2, we compute

\begin{align*} \Lambda^\infty(V(\varpi_1), V(\varpi_1)_{({-}q)^{k}} ) & = \delta_{k,2} + \delta_{k,8} \quad \text{for $ k=2,4,8,10,12,14 $}, \\ \Lambda^\infty(V(\varpi_1), V(\varpi_2)_{({-}q)^{k}} ) & = \delta_{k,9} \quad \text{for $ k={-}1, 1,9,11,13$}, \end{align*}

which give the values of ${ \mathsf {m}^Q_{i,j} }$. Therefore, one can check that the matrix ${ \mathsf {M}_Q }$ is a Cartan matrix of type $E_6$.

Type $E_7^{(1)}$. Recall the explicit description of ${\sigma _Q({\mathfrak {g}})}$ for type $E_7^{(1)}$. The Dynkin quiver corresponding to ${\sigma ({\mathfrak {g}})}_Q$ is given in (A.3). In this case, $h=18$ and elements $(i, (-q)^k)$ of ${\sigma _Q({\mathfrak {g}})}$ with the values of $\phi _Q^{-1}$ can be drawn as follows.

Here we set ${\bigl (\begin {smallmatrix}{a_1a_2a_3a_4}\\[3pt]{a_5a_6a_7}\end {smallmatrix}\bigr )} \mathbin {:=} \sum _{i=1}^7 a_i \alpha _i \in {\Delta ^+_{Q}}$, and the underlined ones are simple roots. Using the formula given in Appendix A.1, one can compute that for $1 \leqslant k < h$,

\begin{align*} & \tilde{c}_{1,1}(k) = \delta_{k,1} + \delta_{k,7} + \delta_{k,11} + \delta_{k,17}, \quad \tilde{c}_{1,2}(k) = \delta_{k,4} + \delta_{k,8} + \delta_{k,10} + \delta_{k,14}, \\ & \tilde{c}_{7,1}(k) = \delta_{k,6} + \delta_{k,12}, \quad \tilde{c}_{7,2}(k) = \delta_{k,5} + \delta_{k,9} + \delta_{k,13}, \quad \tilde{c}_{7,7}(k) = \delta_{k,1} + \delta_{k,9} + \delta_{k,17}. \end{align*}

By Lemma 6.2, we compute

\begin{align*} \Lambda^\infty(V(\varpi_1), V(\varpi_1)_{({-}q)^{2}} ) & = 1, \quad \Lambda^\infty(V(\varpi_1), V(\varpi_2)_{({-}q)} ) = \Lambda^\infty(V(\varpi_2), V(\varpi_1)_{({-}q)} )= 0, \\ \Lambda^\infty(V(\varpi_7), V(\varpi_1)_{({-}q)^{k}} ) & = \delta_{k,13} \quad \text{for $ k=13,15,17,19,21$}, \\ \Lambda^\infty(V(\varpi_7), V(\varpi_{2})_{{({-}q)}^k}) & = \delta_{k,14} \quad \text{for $ k=14,16,18,20$}, \\ \Lambda^\infty(V(\varpi_7), V(\varpi_{7})_{{({-}q)}^k} ) & = \delta_{k,2} \quad \text{for $ k=2,4,6$}, \end{align*}

which give the values of ${ \mathsf {m}^Q_{i,j} }$. Therefore, one can check that the matrix ${ \mathsf {M}_Q }$ is a Cartan matrix of type $E_7$.

Type $E_8^{(1)}$. Recall the explicit description of ${\sigma _Q({\mathfrak {g}})}$ for type $E_8^{(1)}$. The Dynkin quiver corresponding to ${\sigma ({\mathfrak {g}})}_Q$ is given in (A.3). In this case, $h=30$ and elements $(i, (-q)^k)$ of ${\sigma _Q({\mathfrak {g}})}$ with the values of $\phi _Q^{-1}$ can be drawn as follows.

Here we set ${\Bigl (\begin {smallmatrix}{a_1a_2}\\ {a_3a_4}\\ {a_5a_6}\\ {a_7a_8}\end {smallmatrix}\Bigr )} \mathbin {:=} \sum _{i=1}^8 a_i \alpha _i \in {\Delta ^+_{Q}}$, and the underlined ones are simple roots. Using the formula given in Appendix A.1, one can compute that for $1 \leqslant k < h$,

\begin{align*} \tilde{c}_{1,1}(k) & = \delta(k=1,7,11,13,17,19,23,29), \\ \tilde{c}_{1,2}(k) & = \delta(k=4,8,10,12,14,16,18,20,22,26), \\ \tilde{c}_{8,1}(k) & = \delta(k=7,13,17,23), \\ \tilde{c}_{8,2}(k) & = \delta(k=6,10,14,16,20,24), \\ \tilde{c}_{8,8}(k) & = \delta(k=1,11,19,29). \end{align*}

By Lemma 6.2, we compute

\begin{align*} \Lambda^\infty(V(\varpi_1), V(\varpi_1)_{({-}q)^{2}} ) & = 1, \quad \Lambda^\infty(V(\varpi_1), V(\varpi_2)_{({-}q)} ) = \Lambda^\infty(V(\varpi_2), V(\varpi_1)_{({-}q)} )= 0, \\ \Lambda^\infty(V(\varpi_8), V(\varpi_1)_{({-}q)^{k}} ) & = \delta_{k,24} \quad \text{for $ k=24,26,28,30,32,34$}, \\ \Lambda^\infty(V(\varpi_8), V(\varpi_{2})_{({-}q)^k} ) & = \delta_{k,25} \quad \text{for $ k=25,27,29,31,33$}, \\ \Lambda^\infty(V(\varpi_8), V(\varpi_{8})_{({-}q)^k} ) & = \delta_{k,2} \quad \text{for $ k=2,4,6,8$}, \end{align*}

which give the values of ${ \mathsf {m}^Q_{i,j} }$. Therefore, one can check that the matrix ${ \mathsf {M}_Q }$ is a Cartan matrix of type $E_8$.

Type $F_4^{(1)}$. Recall the explicit description of ${\sigma _Q({\mathfrak {g}})}$ for type $F_4^{(1)}$, which can be obtained from [Reference Oh and ScrimshawOS19a]. In this case ${{\mathfrak {g}}_{\mathrm {fin}}}$ is of type $E_{6}$, and elements of ${\sigma _Q({\mathfrak {g}})}$ with the values of $\phi _Q^{-1}$ can be drawn as follows.

Here we set ${\bigl (\begin {smallmatrix}{a_1a_2a_3}\\[3pt]{a_4a_5a_6}\end {smallmatrix}\bigr )} \mathbin {:=} \sum _{i=1}^6 a_i \alpha _i \in {\Delta ^+_{Q}}$, and the underlined ones are simple roots. It follows from Propositions 3.11 and 3.9 and the denominator formula in Appendix A that $\operatorname {\mathfrak {d}} (V_Q(\alpha _i), \mathscr {D}^k V_Q(\alpha _j) ) = 0$ for $i\ne j$ and $k \ne 0$ and that

\begin{align*} \Lambda^\infty(V(\varpi_1), V(\varpi_1)_{ q_s ^k}) & = \operatorname{\mathfrak{d}} ( V(\varpi_1), V(\varpi_1)_{ q_s^k} ) = \delta_{k,4} \quad \text{for } k=2,4, \\ \Lambda^\infty(V(\varpi_3), V(\varpi_1)_{ q_s ^k}) & = \operatorname{\mathfrak{d}} ( V(\varpi_3), V(\varpi_1)_{ q_s^k} ) = \delta_{k, {15} } \quad \text{for } k= {15}, {17}, {19}, \\ \Lambda^\infty(V(\varpi_4), V(\varpi_1)_{ {-}q_s ^k}) & = \operatorname{\mathfrak{d}} ( V(\varpi_4), V(\varpi_1)_{ - q_s^k} ) \\ & = \delta_{k,14} \quad \text{for } k={-}2,0,2,12,14,16, \\ \Lambda^\infty(V(\varpi_3), V(\varpi_4)_{ - q_s^k}) & = \operatorname{\mathfrak{d}} ( V(\varpi_3), V(\varpi_4)_{- q_s^k} ) = 1 \quad \text{for } k= {3}, {17}, \\ \Lambda^\infty(V(\varpi_4), V(\varpi_4)_{ q_s^{14}}) & = \operatorname{\mathfrak{d}} ( V(\varpi_4), V(\varpi_4)_{ q_s^{14}} ) = 0, \end{align*}

which give the values of ${ \mathsf {m}^Q_{i,j} }$. Thus, one can check that the matrix ${ \mathsf {M}_Q }$ is a Cartan matrix of type $E_6$.

Type $G_2^{(1)}$. Recall the explicit description of ${\sigma _Q({\mathfrak {g}})}$ for type $G_2^{(1)}$, which can be obtained from [Reference Oh and ScrimshawOS19a]. In this case ${{\mathfrak {g}}_{\mathrm {fin}}}$ is of type $D_{4}$, and elements of ${\sigma _Q({\mathfrak {g}})}$ with the values of $\phi _Q^{-1}$ can be drawn as follows.

Here we set ${ \bigl (\begin {smallmatrix} a_4 \\[3pt] a_1 a_2 a_3 \end {smallmatrix}\bigr ) } \mathbin {:=} \sum _{k=1}^4 a_k\alpha _k \in {\Delta ^+_{Q}}$, and the underlined ones are simple roots. It follows from Propositions 3.11 and 3.9 and the denominator formula in Appendix A that $\operatorname {\mathfrak {d}} (V_Q(\alpha _i), \mathscr {D}^k V_Q(\alpha _j) ) = 0$ for $i\ne j$ and $k \ne 0$ and that

\begin{align*} \Lambda^\infty(V(\varpi_1), V(\varpi_2)_{ ({-}q_t)^k}) & = \operatorname{\mathfrak{d}} ( V(\varpi_1), V(\varpi_2)_{ ({-}q_t)^k} ) = \delta_{k,11} \quad \text{for } k=3,9,11, \\ \Lambda^\infty(V(\varpi_2), V(\varpi_2)_{ ({-}q_t)^k}) & = \operatorname{\mathfrak{d}} ( V(\varpi_2), V(\varpi_2)_{ ({-}q_t)^k} ) = \delta_{k,2} + \delta_{k,8} \quad \text{for } k=2,6,8, \end{align*}

which give the values of ${ \mathsf {m}^Q_{i,j} }$. Thus, one can check that the matrix ${ \mathsf {M}_Q }$ is a Cartan matrix of type $D_4$.

Type $E_6^{(2)}$. Recall the explicit description of ${\sigma _Q({\mathfrak {g}})}$ for type $E_6^{(2)}$, which can be obtained from [Reference Oh and ScrimshawOS19a]. In this case ${{\mathfrak {g}}_{\mathrm {fin}}}$ is of type $E_{6}$, and elements of ${\sigma _Q({\mathfrak {g}})}$ with the values of $\phi _Q^{-1}$ can be drawn as follows.

Here we set ${\bigl (\begin {smallmatrix}{a_1a_2a_3}\\[3pt]{a_4a_5a_6}\end {smallmatrix}\bigr )}\mathbin {:=} \sum _{i=1}^6 a_i \alpha _i \in {\Delta ^+_{Q}}$, and the underlined ones are simple roots. Note that $V(\varpi _i)_{a} \simeq V(\varpi _i)_{-a}$ for $i =3,4$. It follows from Propositions 3.11 and 3.9 and the denominator formula in Appendix A that $\operatorname {\mathfrak {d}} (V_Q(\alpha _i), \mathscr {D}^k V_Q(\alpha _j) ) = 0$ for $i\ne j$ and $k \ne 0$ and that

\begin{align*} \Lambda^\infty(V(\varpi_1), V(\varpi_1)_{ q ^k}) & = \operatorname{\mathfrak{d}} ( V(\varpi_1), V(\varpi_1)_{ q^k} ) = \delta_{k,2} + \delta_{k,8} \quad \text{for } k=2,4,8,10,12,14, \\ \Lambda^\infty(V(\varpi_1), V(\varpi_4)_{ \sqrt{-1}q ^k}) & = \operatorname{\mathfrak{d}} ( V(\varpi_1), V(\varpi_4)_{ \sqrt{-1} q^k} ) \\ & = \delta_{k,9} \quad \text{for } k={-}1,0,1,9,11,13, \end{align*}

which give the values of ${ \mathsf {m}^Q_{i,j} }$. Thus, one can check that the matrix ${ \mathsf {M}_Q }$ is a Cartan matrix of type $E_6$.

Type $D_4^{(3)}$. Recall the explicit description of ${\sigma _Q({\mathfrak {g}})}$ for type $D_4^{(3)}$, which can be obtained from [Reference Oh and ScrimshawOS19a]. In this case ${{\mathfrak {g}}_{\mathrm {fin}}}$ is of type $D_{4}$, and elements of ${\sigma _Q({\mathfrak {g}})}$ with the values of $\phi _Q^{-1}$ can be drawn as follows.

Here we set ${\bigl (\begin {smallmatrix} a_4 \\[3pt] a_1 a_2 a_3 \end {smallmatrix}\bigr ) } \mathbin {:=} \sum _{k=1}^4 a_k\alpha _k \in {\Delta ^+_{Q}}$, and the underlined ones are simple roots. Note that $V(\varpi _2)_{a} \simeq V(\varpi _2)_{\omega ^t a}$ for $t=1,2$. It follows from Propositions 3.11 and 3.9 and the denominator formula in Appendix A that $\operatorname {\mathfrak {d}} (V_Q(\alpha _i), \mathscr {D}^k V_Q(\alpha _j) ) = 0$ for $i\ne j$ and $k \ne 0$ and that

\begin{align*} \Lambda^\infty(V(\varpi_1), V(\varpi_1)_{ \omega^t q^k}) = \begin{cases} 1 & \text{if } (t,k)= (0,2), (1,4), (2,4), \\ 0 & \text{if } (t,k)= (1,0), (2,0), (1,6), (2,6), \end{cases} \end{align*}

which give the values of ${ \mathsf {m}^Q_{i,j} }$. Thus, one can check that the matrix ${ \mathsf {M}_Q }$ is a Cartan matrix of type $D_4$.

Acknowledgements

The second, third and fourth authors gratefully acknowledge the hospitality of RIMS (Kyoto University) during their visit in 2020. The authors would like to thank the anonymous referee for valuable comments and suggestions.

Appendix A. Denominator formulas

The denominator formulas were studied and computed in [Reference Akasaka and KashiwaraAK97, Reference Date and OkadoDO94, Reference FujitaFuj22, Reference Kang, Kashiwara and KimKKK15, Reference OhOh15, Reference Oh and ScrimshawOS19a]. In this appendix we present the denominator formulas for all types.

Let $q_s, q_t \in \mathbf {k}^\times$ be such that $q = q_s^2= q_t^3$, and let $\omega \in \mathbf {k}$ be such that $\omega ^2+\omega +1=0$. For $i,j \in I$, we set

\[ d_{i,j}(z) \mathbin{:=} d_{V(\varpi_i), V(\varpi_j)} (z). \]

A.1 Simply laced affine ADE types

Suppose that the Cartan matrix $\mathsf {C} = (c_{i,j})_{i,j\in I_0}$ is of type $A_n$, $D_n$ or $E_k$ ($k=6,7,8$). The quantum Cartan matrix $\mathsf {C}(z)= (c_{i,j}(z))_{i,j\in I_0}$ is defined by

\[ c_{i,j}(z) \mathbin{:=} \delta(i=j) (z + z^{{-}1}) + \delta(i\ne j) c_{i,j}. \]

We denote by $\tilde {\mathsf {C}} (z) = ( \tilde {c}_{i,j}(z) )_{i,j\in I_0}$ the inverse of $\mathsf {C}(z)$, and write

(A.1)\begin{equation} \tilde{c}_{i,j}(z) = \sum_{k\in \mathbb{\mspace{1mu}Z\mspace{1mu}}_{\geqslant0}} \tilde{c}_{i,j}(k) z^k \quad \text{for } i,j \in I_0. \end{equation}

Then the following beautiful formula is given in [Reference FujitaFuj22, Theorem 2.10]:

(A.2)\begin{equation} d_{i,j}(z) = \prod_{k=1}^{h -1} (z - ({-}q)^{k+1})^{\tilde{c}_{i,j}(k)}, \end{equation}

where $h$ is the Coxeter number. Note that the dual Coxeter number is equal to the Coxeter number in this case.

Let ${\mathfrak {g}}_0$ be a simple Lie algebra of type ADE with index set $I_0$, and let $Q$ be a Dynkin quiver of ${\mathfrak {g}}$. Let $\xi \colon I_0 \rightarrow \mathbb {\mspace {1mu}Z\mspace {1mu}}$ be a height function such that $\xi _j = \xi _i - 1$ for $i \rightarrow j$ in $Q$. Choose a total order > on $I$ such that $i > j$ for $\xi _i > \xi _j$ and write $I_0 = \{ i_1 > i_2 > \cdots > i_n \}$. We set $\tau \mathbin {:=} s_{i_1} \cdots s_{i_n}$, which is a Coxeter element. For $i\in I_0$ we set $\gamma _i \mathbin {:=} \sum _{j\in B(i)} \alpha _j,$ where $B(i)$ is the subset of $I_0$ consisting of all elements $j$ such that there is a path from $j$ to $i$ in $Q$. Then we have the following.

Proposition A.1 [Reference Hernandez and LeclercHL15, Proposition 2.1]

For $i,j\in I$ and $k \in \mathbb {\mspace {1mu}Z\mspace {1mu}}_{>0}$, we have

\begin{align*} \tilde{c}_{i,j}(k) = \begin{cases} ( \tau^{(k +\xi_i - \xi_j -1)/2} (\gamma_i), \varpi_j) & \text{if $k +\xi_i - \xi_j -1$ is even}, \\ 0 & \text{otherwise.} \end{cases} \end{align*}

In this paper, we make the following choice of Dynkin quivers:

(A.3)

In this case we have the following data, which allow us to compute $\tilde {c}_{i,j}(k)$ explicitly.

  1. (a) (Type $A_n$) $\tau = s_1s_2 \cdots s_n$, $\xi _i =1-i$ and $\gamma _i = \sum _{j=1}^i \alpha _j$.

  2. (b) (Type $D_n$) $\tau = s_1s_2 \cdots s_{n-1} s_n$ and

    \begin{align*} \xi_i = \begin{cases} 1-i & \text{if } i < n-1, \\ -n+2 & \text{if } i=n-1,n, \end{cases} \qquad \gamma_i = \begin{cases} \displaystyle \sum_{j=1}^i \alpha_j & \text{if } i < n, \\ \displaystyle \sum_{j=1}^{n-2} \alpha_j + \alpha_n & \text{if } i=n. \end{cases} \end{align*}
  3. (c) (Type $E_n, \: n=6,7,8$) $\tau = s_1s_2\ldots s_n$, $\xi _1= 0$, $\xi _2=-1$ and $\xi _k =2-k$ for $k=3,4,\ldots,n$, and $\gamma _1 = \alpha _1$, $\gamma _2 = \alpha _2$, $\gamma _3 = \alpha _1+\alpha _3$ and $\gamma _t = \sum _{k=1}^t \alpha _k$ for $t=4,\ldots,n$.

Indeed, in the figures of § 6.2, the root $\gamma _i$ is the rightmost one in the row labeled by $i$, and $\tau$ corresponds to horizontal translation by $-2$. Hence one can read such values of $\tilde {c}_{i,j}(k)$ easily from the figures.

A.2 Other classical affine types

The denominator formulas for other classical affine types can be found in [Reference Akasaka and KashiwaraAK97, Appendix C.4] for type $C_n^{(1)}$ and in [Reference OhOh15, Appendix] for types $B_n^{(1)}$, $D_{n+1}^{(2)}$ and $A_N^{(2)}$ ($N=2n, 2n-1$).

  1. (i) Type $B_{n}^{(1)}$ ($n \geqslant 2$):

    1. (a) $d_{k,l}(z) = \prod _{s=1}^{\min (k,l)} \bigl (z-(-q)^{|k-l|+2s}\bigr )\bigl (z+(-q)^{2n-k-l-1+2s}\bigr )$ for $1 \leqslant k,l \leqslant n-1$;

    2. (b) $d_{k,n}(z) = \prod _{s=1}^{k}\bigl (z-(-1)^{n+k}q_s^{2n-2k-1+4s}\bigr )$ for $1 \leqslant k \leqslant n-1$;

    3. (c) $d_{n,n}(z)= \prod _{s=1}^{n} \bigl (z-(q_s)^{4s-2}\bigr )$.

  2. (ii) Type $C_{n}^{(1)}$ ($n \geqslant 2$):

    1. (a) $d_{k,l}(z)= \prod _{s=1}^{ \min (k,l,n-k,n-l)} \bigl (z-(-q_s)^{|k-l|+2s}\bigr ) \prod _{s=1}^{ \min (k,l)} \bigl (z-(-q_s)^{2n+2-k-l+2s}\bigr )$ for $1 \leqslant k,l \leqslant n$.

  3. (iii) Type $A_{2n-1}^{(2)}$ ($n\geqslant 2$):

    1. (a) $d_{k,l}(z)= \prod _{s=1}^{\min (k,l)} \bigl (z-(-q)^{|k-l|+2s}\bigr )\bigl (z+(-q)^{2n-k-l+2s}\bigr )$ for $1 \leqslant k,l \leqslant n$.

  4. (iv) Type $A_{2n}^{(2)}$ ($n \geqslant 1$):

    1. (a) $d_{k,l}(z) = \prod _{s=1}^{\min (k,l)} \bigl (z-(-q)^{|k-l|+2s}\bigr )\bigl (z-(-q)^{2n+1-k-l+2s}\bigr )$ for $1 \leqslant k,l \leqslant n$.

  5. (v) Type $D_{n+1}^{(2)}$ ($n \geqslant 3$):

    1. (a) $d_{k,l}(z) = \prod _{s=1}^{\min (k,l)} \bigl (z^2 - (-q^2)^{|k-l|+2s}\bigr )\bigl (z^2 - (-q^2)^{2n-k-l+2s}\bigr )$ for $1 \leqslant k,l \leqslant n-1$;

    2. (b) $d_{k,n}(z) = \prod _{s=1}^{k}\bigl (z^2+(-q^{2})^{n-k+2s}\bigr )$ for $1 \leqslant k \leqslant n-1$;

    3. (c) $d_{n,n}(z)= \prod _{s=1}^{n} \bigl (z+(-q^2)^{s}\bigr )$ for $k=l=n$.

A.3 Other exceptional affine types

The denominator formulas for exceptional affine type can be found in [Reference Oh and ScrimshawOS19a, §§ 4 and 7].

  1. (i) Type $G_2^{(1)}$:

    1. (a) $d_{1,1}(z) = (z - q_t^6) (z - q_t^{8}) (z - q_t^{10}) (z - q_t^{12})$;

    2. (b) $d_{1,2}(z) = (z + q_t^7) (z + q_t^{11})$;

    3. (c) $d_{2,2}(z) = (z -q_t^2) (z -q_t^{8}) (z -q_t^{12})$.

  2. (ii) Type $F_4^{(1)}$:

    1. (a) $d_{1,1}(z) = (z-q_s^4) (z-q_s^{10}) (z-q_s^{12}) (z-q_s^{18})$;

    2. (b) $d_{1,2}(z) = (z+q_s^{6}) (z+q_s^{8}) (z+q_s^{10}) (z+q_s^{12}) (z+q_s^{14}) (z+q_s^{16})$;

    3. (c) $d_{1,3}(z) = (z-q_s^{7}) (z-q_s^{9}) (z-q_s^{13}) (z-q_s^{15})$;

    4. (d) $d_{1,4}(z) = (z+q_s^{8}) (z+q_s^{14})$;

    5. (e) $d_{2,2}(z) = (z-q_s^{4}) (z-q_s^{6}) (z-q_s^{8})^2 (z-q_s^{10})^2 (z-q_s^{12})^2 (z-q_s^{14})^2 (z-q_s^{16}) (z-q_s^{18})$;

    6. (f) $d_{2,3}(z) = (z+q_s^{5}) (z+q_s^{7}) (z+q_s^{9}) (z+q_s^{11})^2 (z+q_s^{13}) (z+q_s^{15}) (z+q_s^{17})$;

    7. (g) $d_{2,4}(z) = (z-q_s^{6}) (z-q_s^{10}) (z-q_s^{12}) (z-q_s^{16})$;

    8. (h) $d_{3,3}(z) = (z - q_s^{2}) (z - q_s^{6}) (z - q_s^{8}) (z - q_s^{10}) (z - q_s^{12})^2 (z - q_s^{16}) (z - q_s^{18})$;

    9. (i) $d_{3,4}(z) = (z+q_s^{3}) (z+q_s^{7}) (z+q_s^{11}) (z+q_s^{13}) (z+q_s^{17})$;

    10. (j) $d_{4,4}(z) = (z - q_s^{2}) (z - q_s^{8}) (z - q_s^{12}) (z - q_s^{18})$.

  3. (iii) Type $D_4^{(3)}$:

    1. (a) $d_{1,1}(z) = (z-q^2) (z-q^6) (z-\omega q^4) (z - \omega ^2 q^4)$;

    2. (b) $d_{1,2}(z) = (z^3 + q^9) (z^3 + q^{15})$;

    3. (c) $d_{2,2}(z) = (z^3-q^6) (z^3-q^{12})^{2} (z^3-q^{18})$.

  4. (iv) Type $E_6^{(2)}$:

    1. (a) $d_{1,1}(z) = (z-q^{2}) (z+q^{6}) (z-q^{8}) (z+q^{12})$;

    2. (b) $d_{1,2} (z) = ( z+q^{3}) ( z-q^{5}) ( z-q^{7}) ( z+q^{7}) ( z+q^{9}) ( z-q^{11})$;

    3. (c) $d_{1,3}(z) = (z^2 + q^{8}) (z^2 + q^{12}) (z^2 + q^{16}) (z^2 + q^{20})$;

    4. (d) $d_{1,4}(z) = (z^2 + q^{10}) (z^2 + q^{18})$;

    5. (e) $d_{2,2}(z) = (z-q^{2}) (z-q^{4}) (z-q^{6}) (z-q^{8})^2 (z-q^{10}) (z+q^{4}) (z+q^{6})^2 (z+q^{8}) (z+q^{10}) (z+q^{12})$;

    6. (f) $d_{2,3}(z) = (z^2 + q^{6}) (z^2 + q^{10})^2 (z^2 + q^{14})^2 (z^2 + q^{18})^2 (z^2 + q^{22})$;

    7. (g) $d_{2,4}(z) = (z^2 + q^8) (z^2 + q^{12}) (z^2 + q^{16}) (z^2 + q^{20})$;

    8. (h) $d_{3,3}(z) = (z^2 - q^{4}) (z^2 - q^{8})^2 (z^2 - q^{12})^3 (z^2 - q^{16})^3 (z^2 - q^{20})^2 (z^2 - q^{24})$;

    9. (i) $d_{3,4}(z) = ( z^2 - q^{6}) ( z^2 - q^{10}) ( z^2 - q^{14})^2 ( z^2 - q^{18})^{2} ( z^2 - q^{22})$;

    10. (j) $d_{4,4}(z) = (z^2 - q^4) (z^2 - q^{12}) (z^2 - q^{16}) (z^2 - q^{24})$.

Footnotes

The research of M. Kashiwara was supported by Grant-in-Aid for Scientific Research (B) 20H01795 from the Japan Society for the Promotion of Science.

The research of M. Kim was supported by a National Research Foundation (NRF) grant funded by the government of Korea (MSIP) (NRF-2017R1C1B2007824 and NRF-2020R1A5A1016126).

The research of S.-J.O. was supported by the Ministry of Education of the Republic of Korea and the National Research Foundation of Korea (NRF-2019R1A2C4069647).

The research of E.P. was supported by a National Research Foundation (NRF) grant funded by the government of Korea (MSIP)(NRF-2020R1F1A1A01065992 and NRF-2020R1A5A1016126).

References

Akasaka, T. and Kashiwara, M., Finite-dimensional representations of quantum affine algebras, Publ. RIMS Kyoto Univ. 33 (1997), 839867.CrossRefGoogle Scholar
Chari, V., Braid group actions and tensor products, Int. Math. Res. Not. IMRN 2002 (2010), 357382.CrossRefGoogle Scholar
Chari, V. and Moura, A. A., Spectral characters of finite-dimensional representations of affine algebras, J. Algebra 279 (2004), 820839.CrossRefGoogle Scholar
Chari, V. and Moura, A. A., Characters and blocks for finite-dimensional representations of quantum affine algebras, Int. Math. Res. Not. IMRN 5 (2005), 257298.CrossRefGoogle Scholar
Chari, V. and Pressley, A., A guide to quantum groups (Cambridge University Press, Cambridge, 1994).Google Scholar
Date, E. and Okado, M., Calculation of excitation spectra of the spin model related with the vector representation of the quantized affine algebra of type $A^{(1)}_n$, Internat. J. Modern Phys. A 9 (1994), 399417.CrossRefGoogle Scholar
Drinfel'd, V. G., Quantum groups, in Proc. int. cong. math., vol. 1 and 2, Berkeley, CA, 1986 (American Mathematical Society, Providence, RI, 1987), 798–820.Google Scholar
Etingof, P. and Moura, A. A., Elliptic central characters and blocks of finite dimensional representations of quantum affine algebras, Represent. Theory 7 (2003), 346373.CrossRefGoogle Scholar
Frenkel, E. and Hernandez, D., Langlands duality for representations of quantum groups, Math. Ann. 349 (2011), 705746.CrossRefGoogle Scholar
Frenkel, E. and Hernandez, D., Langlands duality for finite-dimensional representations of quantum affine algebras, Lett. Math. Phys. 96 (2011), 217261.CrossRefGoogle Scholar
Frenkel, E. and Reshetikhin, N., Deformations of $W$-algebras associated to simple Lie algebras, Comm. Math. Phys. 197 (1998), 132.Google Scholar
Frenkel, E. and Reshetikhin, N., The q-characters of representations of quantum affine algebras and deformations of W-algebras, recent developments in quantum affine algebras and related topics, Contemp. Math. 248 (1999), 163205.CrossRefGoogle Scholar
Fujita, R., Geometric realization of Dynkin quiver type quantum affine Schur–Weyl duality, Int. Math. Res. Not. IMRN 22 (2020), 83538386.Google Scholar
Fujita, R., Graded quiver varieties and singularities of normalized R-matrices for fundamental modules, Selecta Math. (N.S.) 28(1) (2022), Paper no. 2.CrossRefGoogle Scholar
Fujita, R., Hernandez, D., Oh, S.-J. and Oya, H., Isomorphisms among quantum Grothendieck rings and propagation of positivity, Preprint (2021), arXiv:2101.07489v2.Google Scholar
Fujita, R. and Oh, S.-J., $Q$-data and representation theory of untwisted quantum affine algebras, Comm. Math. Phys. 384 (2021), 13511407.CrossRefGoogle Scholar
Hernandez, D. and Leclerc, B., Cluster algebras and quantum affine algebras, Duke Math. J. 154 (2010), 265341.CrossRefGoogle Scholar
Hernandez, D. and Leclerc, B., Quantum Grothendieck ring and derived Hall algebras, J. Reine Angew. Math. 701 (2015), 77126.CrossRefGoogle Scholar
Hernandez, D. and Leclerc, B., A cluster algebra approach to q-characters of Kirillov-Reshetikhin modules, J. Eur. Math. Soc. 18 (2016), 11131159.CrossRefGoogle Scholar
Jakelić, D. and Moura, A. A., Tensor products, characters, and blocks of finite-dimensional representations of quantum affine algebras at roots of unity, Int. Math. Res. Not. IMRN 18 (2011), 41474199.Google Scholar
Kac, V., Infinite dimensional lie algebras, third edition (Cambridge University Press, Cambridge, 1990).CrossRefGoogle Scholar
Kang, S.-J., Kashiwara, M. and Kim, M., Symmetric quiver Hecke algebras and $R$-matrices of quantum affine algebras II, Duke Math. J. 164 (2015), 15491602.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., Symmetric quiver Hecke algebras and R-matrices of quantum affine algebras IV, Selecta Math. (N.S.) 22 (2016), 19872015.CrossRefGoogle Scholar
Kashiwara, M., On level zero representations of quantum affine algebras, Duke Math. J. 112 (2002), 117175.CrossRefGoogle Scholar
Kashiwara, M., Kim, M., Oh, S.-J. and Park, E., Monoidal categorification and quantum affine algebras, Compos. Math. 156 (2020), 10391077.CrossRefGoogle Scholar
Kashiwara, M., Kim, M., Oh, S.-J. and Park, E., Monoidal categorification and quantum affine algebras II, Preprint (2021), arXiv:2103.10067v1.Google Scholar
Kashiwara, M. and Oh, S.-J., Categorical relations between Langlands dual quantum affine algebras: doubly laced types, J. Algebraic Combin. 49 (2019), 401435.CrossRefGoogle Scholar
Lusztig, G., Quantum groups at roots of 1, Geom. Dedicata 35 (1990), 89–11.CrossRefGoogle Scholar
Nakajima, H., Quiver varieties and finite-dimensional representations of quantum affine algebras, J. Amer. Math. Soc. 14 (2001), 145238.CrossRefGoogle Scholar
Oh, S.-J., The denominators of normalized R-matrices of types $A^{(2)}_{2n-1}$, $A^{(2)}_{2n}$, $B^{(1)}_n$ and $D^{(2)}_{n+1}$, Publ. RIMS Kyoto Univ. 51 (2015), 709744.CrossRefGoogle Scholar
Oh, S.-J. and Scrimshaw, T., Categorical relations between Langlands dual quantum affine algebras: exceptional cases, Comm. Math. Phys. 368 (2019), 295367.CrossRefGoogle Scholar
Oh, S.-J. and Suh, U., Twisted and folded Auslander-Reiten quiver and applications to the representation theory of quantum affine algebras, J. Algebra 535 (2019), 53132.Google Scholar
Saito, Y., PBW basis of quantized universal enveloping algebras, Publ. Res. Inst. Math. Sci. 30 (1994), 209232.CrossRefGoogle Scholar
Senesi, P., The block decomposition of finite-dimensional representations of twisted loop algebras, Pacific J. Math. 244 (2010), 335357.CrossRefGoogle Scholar