Monoidal categorification of cluster algebras

By Seok-Jin Kang, Masaki Kashiwara, Myungho Kim, and Se-jin Oh

Abstract

We prove that the quantum cluster algebra structure of a unipotent quantum coordinate ring , associated with a symmetric Kac–Moody algebra and its Weyl group element , admits a monoidal categorification via the representations of symmetric Khovanov–Lauda–Rouquier algebras. In order to achieve this goal, we give a formulation of monoidal categorifications of quantum cluster algebras and provide a criterion for a monoidal category of finite-dimensional graded -modules to become a monoidal categorification, where is a symmetric Khovanov–Lauda–Rouquier algebra. Roughly speaking, this criterion asserts that a quantum monoidal seed can be mutated successively in all the directions, once the first-step mutations are possible. Then, we show the existence of a quantum monoidal seed of which admits the first-step mutations in all the directions. As a consequence, we prove the conjecture that any cluster monomial is a member of the upper global basis up to a power of . In the course of our investigation, we also give a proof of a conjecture of Leclerc on the product of upper global basis elements.

Introduction

The purpose of this paper is to provide a monoidal categorification of the quantum cluster algebra structure on the unipotent quantum coordinate ring , which is associated with a symmetric Kac–Moody algebra and a Weyl group element .

The notion of cluster algebras was introduced by Fomin and Zelevinsky in Reference 6 for studying total positivity and upper global bases. Since their introduction, a lot of connections and applications have been discovered in various fields of mathematics including representation theory, Teichmüller theory, tropical geometry, integrable systems, and Poisson geometry.

A cluster algebra is a -subalgebra of a rational function field given by a set of generators, called the cluster variables. These generators are grouped into overlapping subsets, called the clusters, and the clusters are defined inductively by a procedure called mutation from the initial cluster , which is controlled by an exchange matrix . We call a monomial of cluster variables in each cluster a cluster monomial.

Fomin and Zelevinsky proved that every cluster variable is a Laurent polynomial of the initial cluster , and they conjectured that this Laurent polynomial has positive coefficients Reference 6. This positivity conjecture was proved by Lee and Schiffler in the skew-symmetric cluster algebra case in Reference 30. The linearly independence conjecture on cluster monomials was proved in the skew-symmetric cluster algebra case in Reference 4.

The notion of quantum cluster algebras, introduced by Berenstein and Zelevinsky in Reference 3, can be considered as a -analogue of cluster algebras. The commutation relation among the cluster variables is determined by a skew-symmetric matrix . As in the cluster algebra case, every cluster variable belongs to Reference 3 and is expected to be an element of , which is referred to as the quantum positivity conjecture (cf. Reference 5, Conjecture 4.7). In Reference 24, Kimura and Qin proved the quantum positivity conjecture for quantum cluster algebras containing acyclic seed and specific coefficients.

The unipotent quantum coordinate rings and are examples of quantum cluster algebras arising from Lie theory. The algebra is a -deformation of the coordinate ring of the unipotent subgroup and is isomorphic to the negative half of the quantum group as -algebras. The algebra is a -subalgebra of generated by a set of the dual Poincaré–Birkhoff–Witt (PBW) basis elements associated with a Weyl group element . The unipotent quantum coordinate ring has a very interesting basis, the so-called upper global basis (dual canonical basis) , which is dual to the lower global basis (canonical basis) Reference 16Reference 31. The upper global basis has been studied emphasizing its multiplicative structure. For example, Berenstein and Zelevinsky Reference 2 conjectured that, in the case is of type , the product of two elements and in is again an element of up to a multiple of a power of if and only if they are -commuting; i.e., for some . This conjecture turned out to be not true in general, because Leclerc Reference 29 found examples of an imaginary element such that does not belong to . Nevertheless, the idea of considering subsets of whose elements are -commuting with each other and studying the relations between those subsets has survived, and it became one of the motivations of the study of (quantum) cluster algebras.

In a series of papers Reference 8Reference 9Reference 11, Geiß, Leclerc, and Schröer showed that the unipotent quantum coordinate ring has a skew-symmetric quantum cluster algebra structure whose initial cluster consists of the so-called unipotent quantum minors. In Reference 23, Kimura proved that is compatible with the upper global basis of ; i.e., the set is a basis of . Thus, with a result of Reference 4, one can expect that every cluster monomial of is contained in the upper global basis , which is named the quantization conjecture by Kimura Reference 23.

Conjecture (Reference 11, Conjecture 12.9, Reference 23, Conjecture 1.1(2)).

When is a symmetric Kac–Moody algebra, every quantum cluster monomial in belongs to the upper global basis up to a power of .

It can be regarded as a reformulation of Berenstein–Zelevinsky’s ideas on the multiplicative properties of . There are some partial results of this conjecture. It is proved for , , and in Reference 2 and Reference 7, Section 12. When , and is a square of a Coxeter element, it is shown in Reference 26 and Reference 27 that the cluster variables belong to the upper global basis. When is symmetric and is a square of a Coxeter element, the conjecture is proved in Reference 24. Notably, Qin provided recently a proof of the conjecture for a large class with a condition on the Weyl group element Reference 37. Note that Nakajima proposed a geometric approach of this conjecture via quiver varieties Reference 35.

In this paper, we prove the above conjecture completely by showing that there exists a monoidal categorification of .

In Reference 12, Hernandez and Leclerc introduced the notion of monoidal categorification of cluster algebras. A simple object of a monoidal category is real if is simple, and it is prime if there exists no nontrivial factorization . They say that is a monoidal categorification of a cluster algebra if the Grothendieck ring of is isomorphic to and if

(M1)

the cluster monomials of are the classes of real simple objects of ,

(M2)

the cluster variables of are the classes of real simple prime objects of .

(Note that the above version is weaker than the original definition of the monoidal categorification in Reference 12.) They proved that certain categories of modules over symmetric quantum affine algebras give monoidal categorifications of some cluster algebras. Nakajima extended this result to the cases of the cluster algebras of types Reference 36 (see also Reference 13). It is worthwhile to remark that once a cluster algebra has a monoidal categorification, the positivity of cluster variables of and the linear independence of cluster monomials of follow (see Reference 12, Proposition 2.2).

In this paper, we refine Hernandez–Leclerc’s notion of monoidal categorifications including the quantum cluster algebra case. Let us briefly explain it. Let be an abelian monoidal category equipped with an auto-equivalence and a tensor product which is compatible with a decomposition . Fix a finite index set with a decomposition into the exchangeable part and the frozen part. Let be a quadruple of a family of simple objects in , an integer-valued skew-symmetric -matrix , an integer-valued -matrix with a skew-symmetric principal part, and a family of elements in . If this datum satisfies the conditions in Definition 6.2.1 below, then it is called a quantum monoidal seed in . For each , we have mutations , and of , , and , respectively. We say that a quantum monoidal seed admits a mutation in direction if there exists a simple object which fits into two short exact sequences Equation 0.2 below in reflecting the mutation rule in quantum cluster algebras, and thus obtained quadruple is again a quantum monoidal seed in . We call the mutation of in direction .

Now the category is called a monoidal categorification of a quantum cluster algebra over if

(0.1)

(i)

the Grothendieck ring is isomorphic to ,

(ii)

there exists a quantum monoidal seed in such that is a quantum seed of for some ,

(iii)

admits successive mutations in all directions in .

The existence of monoidal category which provides a monoidal categorification of quantum cluster algebra implies the following:

(QM1)

Every quantum cluster monomial corresponds to the isomorphism class of a real simple object of . In particular, the set of quantum cluster monomials is -linearly independent.

(QM2)

The quantum positivity conjecture holds for .

In the case of unipotent quantum coordinate ring , there is a natural candidate for monoidal categorification, the category of finite-dimensional graded modules over a Khovanov–Lauda–Rouquier algebras (Reference 21Reference 22, Reference 38). The Khovanov–Lauda–Rouquier algebras (abbreviated by KLR algebras), introduced by Khovanov–Lauda Reference 21Reference 22 and Rouquier Reference 38 independently, are a family of -graded algebras which categorifies the negative half of a symmetrizable quantum group . More precisely, there exists a family of algebras such that the Grothendieck ring of , the direct sum of the categories of finite-dimensional graded -modules, is isomorphic to the integral form of . Here the tensor functor of the monoidal category is given by the convolution product , and the action of is given by the grading shift functor. In Reference 39Reference 40, Varagnolo–Vasserot and Rouquier proved that the upper global basis of corresponds to the set of the isomorphism classes of all self-dual simple modules of under the assumption that is associated with a symmetric quantum group and the base field is of characteristic .

Combining works of Reference 11Reference 23Reference 40, the unipotent quantum coordinate ring associated with a symmetric quantum group and a Weyl group element is isomorphic to the Grothendieck group of a monoidal abelian full subcategory of whose base field is of characteristic , satisfying the following properties: (i) is stable under extensions and grading shift functor, (ii) the composition factors of are contained in (see Definition 11.2.1). In particular, the first condition in 0.1 holds. However, it is not evident that the second and the third conditions in 0.1 on quantum monoidal seeds are satisfied. The purpose of this paper is to ensure that those conditions hold in .

In order to establish it, in the first part of the paper, we start with a continuation of the work of Reference 15 about the convolution products, heads, and socles of graded modules over symmetric KLR algebras. One of the main results in Reference 15 is that the convolution product of a real simple -module and a simple -module has a unique simple quotient and a unique simple submodule. Moreover, if up to a grading shift, then is simple. In such a case we say that and commute. The main tool of Reference 15 was the R-matrix , constructed in Reference 14, which is a homogeneous homomorphism from to of degree . In this work, we define some integers encoding necessary information on ,

and study the representation theoretic meaning of the integers , , and .

We then prove Leclerc’s first conjecture Reference 29 on the multiplicative structure of elements in , when the generalized Cartan matrix is symmetric (Theorem 4.1.1 and Theorem 4.2.1). Theorem 4.2.1 is due to McNamara Reference 34, Lemma 7.5, and the authors thank him for informing us of his result.

We say that is real if .

Theorem (Reference 29, Conjecture 1).

Let and be elements in such that one of them is real and . Then the expansion of with respect to is of the form

where , , , and

More precisely, we prove that and correspond to the simple head and the simple socle of , respectively, when corresponds to a simple module and corresponds to a simple module .

Next, we move to provide an algebraic framework for monoidal categorification of quantum cluster algebras. In order to simplify the conditions of quantum monoidal seeds and their mutations, we introduce the notion of admissible pairs in . A pair is called admissible in if (i) is a commuting family of self-dual real simple objects of , (ii) is an integer-valued -matrix with a skew-symmetric principal part, and (iii) for each , there exists a self-dual simple object in such that commutes with for all and there are exact sequences in

where and are prescribed integers and is a convolution product up to a power of .

For an admissible pair , let be the skew-symmetric matrix where is the homogeneous degree of , the R-matrix between and , and let be the family of elements in given by .

Then, together with the result of Reference 11, our main theorem in the first part of the paper reads as follows.

Main Theorem 1 (Theorem 7.1.3 and Corollary 7.1.4).

If there exists an admissible pair in such that is an initial seed of , then is a monoidal categorification of .

The second part of this paper (Sections 8–11) is mainly devoted to showing that there exists an admissible pair in for every symmetric Kac–Moody algebra and its Weyl group element . In Reference 11, Geiß, Leclerc, and Schröer provided an initial quantum seed in whose quantum cluster variables are unipotent quantum minors. The unipotent quantum minors are elements in , which are regarded as a -analogue of a generalization of the minors of upper triangular matrices. In particular, they are elements in . We define the determinantial module to be the simple module in corresponding to the unipotent quantum minor under the isomorphism . Here is a pair of elements in the weight lattice of satisfying certain conditions.

Our main theorem of the second part is as follows.

Main Theorem 2 (Theorem 11.2.2).

Let be the initial quantum seed of in Reference 11 with respect to a reduced expression of . Let be the determinantial module corresponding to the unipotent quantum minor . Then the pair

is admissible in .

Combining these theorems, the category gives a monoidal categorification of the quantum cluster algebra . If we take the base field of the symmetric KLR algebra to be of characteristic , these theorems, along with Theorem 2.1.4 due to Reference 39Reference 40, imply the quantization conjecture.

The most essential condition for an admissible pair is that there exists the first mutation in the exact sequences Equation 0.2 for each . To establish this, we investigate the properties of determinantial modules and those of their convolution products. Note that a unipotent quantum minor is the image of a global basis element of the quantum coordinate ring under a natural projection . Since there exists a bicrystal embedding from the crystal basis of to the crystal basis of the modified quantum groups , this investigation amounts to the study of the interplay among the crystal and global bases of , , and . Hence we start the second part of the paper with the studies of those algebras and their crystal/global bases along the line of the works in Reference 17Reference 18Reference 19.

Next, we recall the (unipotent) quantum minors and the T-system, an equation consisting of three terms in products of unipotent quantum minors studied in Reference 3Reference 11. A detailed study of the relation between , , and and their global bases enables us to establish several equations involving unipotent quantum minors in the algebra . The upshot is that those equations can be translated into exact sequences in the category involving convolution products of determinantial modules via the categorification of . It enables us to show that the pair is admissible.

The paper is organized as follows. In Section 1, we briefly review basic materials on quantum group and KLR algebra . In Section 2, we continue the study in Reference 15 of the R-matrices between -modules. In Section 3, we derive certain properties of and . In Section 4, we prove the first conjecture of Leclerc in Reference 29. In Section 5, we recall the definition of quantum cluster algebras. In Section 6, we give the definitions of a monoidal seed, a quantum monoidal seed, a monoidal categorification of a cluster algebra, and a monoidal categorification of a quantum cluster algebra. In Section 7, we prove Main Theorem 1. In Section 8, we review the algebras , , and , and study the relations among them. In Section 9, we study the properties of quantum minors including -systems and generalized -systems. In Section 10, we study the determinantial modules over KLR algebras. Finally, in Section 11, we establish Main Theorem 2.

1. Quantum groups and global bases

In this section, we briefly recall the quantum groups and the crystal and global bases theory for . We refer to Reference 16Reference 17Reference 20 for materials in this subsection.

1.1. Quantum groups

Let be an index set. A Cartan datum is a quintuple consisting of

(i)

an integer-valued matrix , called the symmetrizable generalized Cartan matrix, which satisfies

(a)

,

(b)

,

(c)

there exists a diagonal matrix such that is symmetric, and are relatively prime positive integers,

(ii)

a free abelian group , called the weight lattice,

(iii)

, called the set of simple roots,

(iv)

, called the co-weight lattice,

(v)

, called the set of simple coroots, satisfying the following properties:

(1)

for all ,

(2)

is linearly independent over ,

(3)

for each , there exists such that for all .

We call the fundamental weights.

The free abelian group is called the root lattice. Set and . For , we set .

Set . Then there exists a symmetric bilinear form on satisfying

The Weyl group of is the group of linear transformations on generated by , where

Let be an indeterminate. For each , set .

Definition 1.1.1.

The quantum group associated with a Cartan datum is the algebra over generated by and satisfying the following relations:

Here, we set , and for and such that .

Let (resp. ) be the subalgebra of generated by ’s (resp. ’s), and let be the subalgebra of generated by . Then we have the triangular decomposition

and the weight space decomposition

where .

There are -algebra antiautomorphisms and of given as follows:

There is also a -algebra automorphism of given by

We define the divided powers by

Let us denote by the -subalgebra of generated by , , , and (), where . Let us also denote by the -subalgebra of generated by (, ), and by the -subalgebra of generated by (, ).

1.2. Integrable representations

A -module is called integrable if where , , and the actions of and on are locally nilpotent for all . We denote by the category of integrable left -modules satisfying that there exist finitely many weights , …, such that . The category is semisimple with its simple objects being isomorphic to the highest weight modules with highest weight vector of highest weight , the set of dominant integral weights.

For , let us denote by the left -module with the action of given by

Then belongs to .

For a left -module , we denote by the right -module with the right action of given by

We denote by the category of right integrable -modules such that .

There are two comultiplications and on defined as follows:

For two -modules and , let us denote by and the vector space endowed with -module structure induced by the comultiplications and , respectively. Then we have

For any , there exists a unique -linear endomorphism of such that

The quantum boson algebra is defined as the subalgebra of generated by and . Then has a -algebra anti-automorphism which sends to and to . As a -module, is simple.

The simple -module and the simple -module have a unique non-degenerate symmetric bilinear form such that

Note that induces the non-degenerate bilinear form

given by , by which is canonically isomorphic to .

1.3. Crystal bases and global bases

For a subring of , we say that is an -lattice of a -vector space if is a free -submodule of such that .

Let us denote by (resp. ) the ring of rational functions in which are regular at (resp. ). Set .

Let be a -module in . Then, for each , any can be uniquely written as

We define the lower Kashiwara operators by

and the upper Kashiwara operators by

Similarly, for each , any element can be written uniquely as

We define the Kashiwara operators on by

We say that an -lattice of is a lower (resp. upper) crystal lattice of if , where and it is invariant by the lower (resp. upper) Kashiwara operators.

Lemma 1.3.1.

Let be a lower crystal lattice of . Then we have

(i)

is an upper crystal lattice of .

(ii)

is an upper crystal lattice of .

Proof.

(i) Let be the endomorphism of given by . Then we have and .

Item (ii) follows from , in Reference 17. Note that the definition of upper Kashiwara operators are slightly different from the ones in Reference 17, but similar properties hold.

Definition 1.3.2.

A lower (resp. upper) crystal basis of consists of a pair satisfying the following conditions:

(i)

is a lower (resp. upper) crystal lattice of ,

(ii)

is a basis of the -vector space , where ,

(iii)

the induced maps and on satisfy

Here and denote the lower (resp. upper) Kashiwara operators.

For , let be the highest weight vector of . Let be the -submodule of generated by , and let be the subset of given by

It is shown in Reference 16 that is a lower crystal basis of . Using the non-degenerate symmetric bilinear form , has the upper crystal basis where

and is the dual basis of with respect to the induced non-degenerate pairing between and .

An (abstract) crystal is a set together with maps

such that

(C1)

for any ,

(C2)

if satisfies , then

(C3)

if satisfies , then

(C4)

for , if and only if ,

(C5)

if , then .

Recall that, with the notions of morphism and tensor product rule of crystals, the category of crystals becomes a monoidal category Reference 19. If is a crystal basis of , then is an abstract crystal. Since , we drop the superscripts for simplicity.

Let be a -vector space, and let be an -lattice of , an -lattice of , and an -lattice of . We say that the triple is balanced if the following canonical map is a -linear isomorphism:

The inverse of the above isomorphism is called the globalizing map. If is balanced, then we have

Hence, if is a basis of , then is a basis of , , , and . We call a global basis.

We define the two -lattices of by

Recall that there is a -linear automorphism—on defined by

Then and are balanced. Let us denote by and the associated globalizing maps, respectively. (If there is no danger of confusion, we simply denote them and , respectively.) Then the sets

form -bases of

respectively. They are called the lower global basis and the upper global basis of .

Set

Then is a lower crystal basis of the simple -module and the triple is balanced. Let us denote the globalizing map by . Then the set

forms a -basis of and is called the lower global basis of .

Let us denote by

the dual basis of with respect to . Then it is a -basis of

and called the upper global basis of . Note that has a -algebra structure as a subalgebra of (see also Section 8.2).

2. KLR algebras and R-matrices

2.1. KLR algebras

We recall the definition of Khovanov–Lauda–Rouquier algebra or quiver Hecke algebra (hereafter, we abbreviate it as KLR algebra) associated with a given Cartan datum .

Let be a base field. For such that , set

Let us take a family of polynomials in which are of the form

We denote by the symmetric group on letters, where is the transposition of and . Then acts on by place permutations.

For and such that , we set

Definition 2.1.1.

For with , the KLR algebra at associated with a Cartan datum and a matrix is the algebra over generated by the elements , , satisfying the following defining relations:

The above relations are homogeneous provided that

and hence is a -graded algebra.

For a graded -module , we define , where

We call the grading shift functor on the category of graded -modules.

If is an -module, then we set and call it the weight of .

We denote by the category of -modules, and by the full subcategory of consisting of modules such that are finite-dimensional over , and the actions of the ’s on are nilpotent.

Similarly, we denote by and by the category of graded -modules and the category of graded -modules which are finite-dimensional over , respectively. We set

For with , , set

Then is an idempotent. Let

be the -algebra homomorphism given by ( and ) (), (), (), and (). Here is the concatenation of and ; i.e., .

For an -module and an -module , we define the convolution product by

For , the dual space

admits an -module structure via

where denotes the -algebra anti-involution on which fixes the generators , , and for , and .

It is known that (see Reference 28, Theorem 2.2 (2))

for any and .

A simple module in is called self-dual if . Every simple module is isomorphic to a grading shift of a self-dual simple module Reference 21, Section 3.2. Note also that we have for every simple module in Reference 21, Corollary 3.19.

Let us denote by the Grothendieck group of . Then, is an algebra over with the multiplication induced by the convolution product and the -action induced by the grading shift functor .

In Reference 21Reference 38, it is shown that a KLR algebra categorifies the negative half of the corresponding quantum group. More precisely, we have the following theorem.

Theorem 2.1.2 (Reference 21Reference 38).

For a given Cartan datum , we take a parameter matrix satisfying the conditions in Equation 2.1, and let and be the associated quantum group and the KLR algebras, respectively. Then there exists a -algebra isomorphism

KLR algebras also categorify the upper global bases.

Definition 2.1.3.

We say that a KLR algebra is symmetric if is a polynomial in for all .

In particular, the corresponding generalized Cartan matrix is symmetric. In symmetric case, we assume for .

Theorem 2.1.4 (Reference 39Reference 40).

Assume that the KLR algebra is symmetric and the base field is of characteristic . Then under the isomorphism Equation 2.2 in Theorem 2.1.2, the upper global basis corresponds to the set of the isomorphism classes of self-dual simple -modules.

2.2. R-matrices for KLR algebras

For and , we define by

They are called the intertwiners. Since satisfies the braid relation, does not depend on the choice of reduced expression .

For , let us denote by the element of defined by

Let with , , and let be an -module and an -module. Then the map given by is -linear, and hence it extends to an -module homomorphism

Assume that the KLR algebra is symmetric. Let be an indeterminate which is homogeneous of degree , and let be the graded algebra homomorphism

given by

For an -module , we denote by the -module with the action of twisted by . Namely,

for , , and . Note that the multiplication by on induces an -module endomorphism on . For , we sometimes denote by the corresponding element of the -module .

For a non-zero and a non-zero ,

(2.3)

let be the order of zero of ; i.e., the largest non-negative integer such that the image of is contained in .

Note that such an exists because does not vanish Reference 14, Proposition 1.4.4 (iii). We denote by the morphism .

Definition 2.2.1.

Assume that is symmetric. For a non-zero and a non-zero , let be an integer as in 2.3. We define

by

and call it the renormalized R-matrix.

By the definition, the renormalized R-matrix never vanishes.

We define also

by

where is the order of zero of .

If and are symmetric, then coincides with the order of zero of , and (see Reference 15, (1.11)).

By the construction, if the composition for does not vanish, then it is equal to .

Definition 2.2.2.

A simple -module is called real if is simple.

The following lemma was used significantly in Reference 15.

Lemma 2.2.3 (Reference 15, Lemma 3.1).

Let and . Let be an -submodule of and an -submodule of such that as submodules of . Then there exists an -submodule of such that and .

One of the main results in Reference 15 is the following theorem.

Theorem 2.2.4 (Reference 15, Theorem 3.2).

Let and assume that is symmetric. Let be a real simple module in and a simple module in . Then

(i)

and have simple socles and simple heads.

(ii)

Moreover, is equal to the head of and socle of . Similarly, is equal to the head of and socle of .

We will use the following convention frequently.

Definition 2.2.5.

For simple -modules and , we denote by the head of and by the socle of .

3. Simplicity of heads and socles of convolution products

In this section, we assume that is symmetric for any ; i.e., is a function in for any .

We also work always in the category of graded modules. For the sake of simplicity, we simply say that is an -module instead of saying that is a graded -module for . We also sometimes ignore grading shifts if there is no danger of confusion. Hence, for -modules and , we sometimes say that is a homomorphism if is a morphism in for some . If we want to emphasize that is a morphism in , we say so.

3.1. Homogeneous degrees of R-matrices

Definition 3.1.1.

For non-zero , we denote by the homogeneous degree of the R-matrix .

Hence

are morphisms in and in , respectively.

Lemma 3.1.2.

For non-zero -modules and , we have

Proof.

Set and . By Reference 14, (1.3.3), the homogeneous degree of is , where is the symmetric bilinear form on given by . Hence has degree .

Definition 3.1.3.

For non-zero -modules and , we set

Lemma 3.1.4.

Let and be self-dual simple modules. If one of them is real, then

is a self-dual simple module.

Proof.

Set and . Set for some self-dual simple module and some . Then we have

since . Taking dual, we obtain

In particular, is a simple quotient of . Hence we have , which implies .

Lemma 3.1.5.
(i)

Let be non-zero modules , and let and be non-zero homomorphisms. Assume further that is a simple module. Then the composition

does not vanish.

(ii)

Let be a simple module, and let be non-zero modules. Then the composition

coincides with , and the composition

coincides with .

In particular, we have

and

Proof.
(i)

Assume that the composition vanishes. Then we have . By Lemma 2.2.3, there is a submodule of such that and . The first inclusion implies that since is non-zero, and the second implies since is non-zero. It contradicts the simplicity of .

(ii)

It is enough to show that the compositions and do not vanish, but these immediately follow from (i).

3.2. Properties of and

Lemma 3.2.1.

Let and be simple -modules. Then we have

(i)

(ii)

If for some , then

up to constant multiples.

Proof.

By Reference 14, Proposition 1.6.2, the morphism

is equal to for some . Since is homogeneous of degree , we have for some .

Definition 3.2.2.

For non-zero modules and , we set

Note that if and are simple modules, then we have . Note also that if are simple modules, then we have by Lemma 3.1.5 (ii).

Lemma 3.2.3 (Reference 15).

Let be simple modules and assume that one of them is real. Then the following conditions are equivalent:

(i)

.

(ii)

and are inverse to each other up to a constant multiple.

(iii)

and are isomorphic up to a grading shift.

(iv)

and are isomorphic up to a grading shift.

(v)

is simple.

Proof.

By specializing the equations in Lemma 3.2.1 (ii) at , we obtain that if and only if and up to non-zero constant multiples. Hence the conditions (i) and (ii) are equivalent.

The conditions (ii), (iii), (iv), and (v) are equivalent by Reference 15, Theorem 3.2, Proposition 3.8, and Corollary 3.9.

Definition 3.2.4.

Let be simple modules.

(i)

We say that and commute if .

(ii)

We say that and are simply linked if .

Proposition 3.2.5.

Let be a commuting family of real simple modules. Then the convolution product

is a real simple module.

Proof.

We shall first show the simplicity of the convolutions. By induction on , we may assume that is a simple module. Then we have

so that is simple by Lemma 3.2.3.

Since is also simple, is real.

Definition 3.2.6.

Let be real simple modules. Assume that they commute with each other. We set

It is invariant under the permutations of .

Lemma 3.2.7.

Let be real simple modules commuting with each other. Then for any , we have

Moreover, if the ’s are self-dual, then so is .

Proof.

It follows from Lemma 3.1.4 and .

Proposition 3.2.8.

Let be a morphism between non-zero -modules , and let be a non-zero -module.

(i)

If , then the following diagram is commutative:

(ii)

If , then the composition

vanishes.

(iii)

If , then the composition

vanishes.

(iv)

If is surjective, then we have

If is injective, then we have

Proof.

Let be the order of zero of for . Then we have .

Set . Then the following diagram is commutative:

(i) If , then by specializing in the above diagram, we obtain the commutativity of the diagram in (i).

(ii) If , then we have

so that vanishes. Hence we have

as desired. In particular, is not surjective.

(iii) Similarly, if , then we have and is not injective.

(iv) The statements for and follow from (ii) and (iii). The other statements can be shown in a similar way.

Proposition 3.2.9.

Let and be simple modules. We assume that one of them is real. Then we have

Proof.

Since the other case can be proved similarly, we assume that is real. Let be a morphism. Note that we have and by Lemma 3.1.5 (ii) and by the fact that up to a constant multiple. Thus, by Proposition 3.2.8, we have a commutative diagram (up to a constant multiple)

Hence we have

Hence there exists a submodule of such that and by Lemma 2.2.3. Since , we have . Hence , which means that factors as . It remains to remark that .

Proposition 3.2.10.

Let , , and be simple modules. Then we have

for any subquotient of . Moreover, when is real, the following conditions are equivalent:

(i)

commutes with and .

(ii)

Any simple subquotient of commutes with and satisfies .

(iii)

Any simple subquotient of commutes with and satisfies .

Proof.

The inequalities Equation 3.1 are consequences of Proposition 3.2.8. Let us show the equivalence of (i)–(iii).

Let be a Jordan–Hölder series of . Then the renormalized R-matrix is homogeneous of degree , and it sends to for any . Hence sends to .

First assume (i). Then is an isomorphism. Hence is injective. By comparing their dimension, is an isomorphism, Hence is an isomorphism of homogeneous degree . Hence we obtain (ii).

Conversely, assume (ii). Then, and have the same homogeneous degree, and hence they should coincide. It implies that is an isomorphism for any . Therefore is an isomorphism, which implies that and are isomorphisms. Thus we obtain (i).

Similarly, (i) and (iii) are equivalent.

Lemma 3.2.11.

Let , , and be simple modules. We assume that is real and commutes with . Then the diagram

commutes.

Proof.

Otherwise the composition

vanishes by Proposition 3.2.8. Hence we have

Hence, by Lemma 2.2.3, there exists a submodule of such that

The first inclusion implies and the second implies , which contradicts the simplicity of .

The following lemma can be proved similarly.

Lemma 3.2.12.

Let , , and be simple modules. We assume that is real and commutes with . Then the diagram

commutes.

The following proposition follows from Lemma 3.2.11 and Lemma 3.2.12.

Proposition 3.2.13.

Let , , and be simple modules. Assume that is real. Then we have the following:

(i)

If and commute, then