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

(ii)

If and commute, then

Proposition 3.2.14.

Let be a real simple module, and let be a module with a simple socle. If the following diagram

commutes up to a non-zero constant multiple, then is equal to the socle of . In particular, has a simple socle.

Proof.

Let be an arbitrary simple submodule of . Then we have the following commutative diagram:

By multiplying , where is the order of zero of , and specializing at , we have a commutative diagram (up to a constant multiple)

Here, we use the fact that from Lemma 3.1.5 and the fact that is equal to up to a non-zero constant multiple, because is a real simple module.

It follows that . Hence there exists a submodule of such that and by Lemma 2.2.3. Hence and by the assumption. Hence . Since is non-zero by the assumption, we have . Thus we obtain the desired result.

The following is a dual form of the preceding proposition.

Proposition 3.2.15.

Let be a real simple module. Let be a module with a simple head. If the following diagram

commutes up to a non-zero constant multiple, then is equal to the simple head of .

Proposition 3.2.16.

Let , , and be simple modules. We assume that is real and one of and is real.

(i)

If , then has a simple head and has a simple socle.

(ii)

If , then has a simple head and has a simple socle.

(iii)

If , then and have simple heads, and and have simple socles.

Proof.

(i) Denote and . Then the diagram

commutes. Hence Proposition 3.2.14 and Proposition 3.2.15 imply that has a simple head and has a simple socle. Item (ii) is proved similarly.

(iii) If , then we have and by Proposition 3.2.8. Thus the statements in (iii) follow from (i) and (ii).

Proposition 3.2.17.

Let and be simple modules. Assume that one of them is real and . Then we have an exact sequence

In particular, has length .

Proof.

In the course of the proof, we ignore the grading.

Set and . By let us regard as a submodule of . By the condition, we have up to a constant multiple (see Lemma 3.2.1 (ii)), and hence we have

We have an exact sequence

Since

we have by Proposition 3.2.9. Similarly,

implies that by Proposition 3.2.9.

Lemma 3.2.18.

Let and be simple modules. Assume that one of them is real. If there is an exact sequence

for self-dual simple modules , and integers , , then we have

Proof.

We may assume that and are self-dual without loss of generality. Then we have . Since

we have . Thus we obtain

Lemma 3.2.19.

Let and be simple modules. Assume that one of them is real. If the equation

holds in for self-dual simple modules , and integers , such that , then we have

(i)

,

(ii)

there exists an exact sequence

(iii)

is the socle of and is the head of .

Proof.

First note that since is not simple. By the assumption, there exists either an exact sequence

or

The second sequence cannot exist by Lemma 3.2.18 because . Hence the first sequence exists, and the assertion (iii) follows from Theorem 2.2.4.

Proposition 3.2.20.

Let , and be simple -modules. Assume that there is an exact sequence

and are simple, and are ungraded modules. Then is a real simple module.

Proof.

Assume that is not real. Then is reducible, and we have for any by Reference 15, Corollary 3.3. Note that is of length , because is of length .

Let be a simple submodule of . Consider an exact sequence

Then we have

Indeed, if , then there exists a submodule of such that and by Reference 15, Lemma 3.1. It contradicts the simplicity of . Thus Equation 3.2 holds.

Note that Equation 3.2 implies

since is simple.

(a)

Assume first that is semisimple so that for some simple submodule of . Then . Hence . Therefore we obtain which is a contradiction.

(b)

Assume that is not semisimple so that is a unique non-zero proper submodule of and is a unique non-zero proper quotient of . Without loss of generality, we may assume that is algebraically closed Reference 21, Corollary 3.19. Let be an eigenvalue of . Since , we have . It follows that

and hence we have an exact sequence

Since is of length , we have

which is a contradiction.

Corollary 3.2.21.

Let be simple -modules, and let be a real simple -module. If we have an exact sequence

and if and are simple, then is a real simple module.

Proof.

Since is real and is not simple, is not isomorphic to as an ungraded module by Lemma 3.2.3 (iv). It follows that is not isomorphic to , because is a domain so that for some implies . Now the assertion follows from Proposition 3.2.20.

Lemma 3.2.22.

Let and be a pair of commuting families of real simple modules. We assume that

(a)

is a commuting family of real simple modules,

(b)

commutes with for any .

Then we have

Proof.

Since is simple, it is enough to give an epimorphism . We shall show it by induction on . For , we have

as desired.

4. Leclerc’s conjecture

In this section, is assumed to be a symmetric KLR algebra over a base field .

4.1. Leclerc’s conjecture

The following theorem is a part of Leclerc’s conjecture stated in the Introduction.

Theorem 4.1.1.

Let and be simple modules. We assume that is real. Then we have the equalities in the Grothendieck group as follows:

(i)

with simple modules such that ,

(ii)

with simple modules such that ,

(iii)

with simple modules such that ,

(iv)

with simple modules such that .

In particular, as well as appears only once in the Jordan–Hölder series of in .

The following result is an immediate consequence of this theorem.

Corollary 4.1.2.

Let and be simple modules. We assume that one of them is real. Assume that and do not commute, Then we have the equality in the Grothendieck group

with simple modules . Moreover we have the following:

(i)

If is real, then we have , and , .

(ii)

If is real, then we have , and , .

Proof of Theorem 4.1.1.

We shall prove only (i). The other statements are proved similarly.

Then we have . Let us consider the renormalized R-matrix

Then sends to for any . Hence evaluating the above diagram at , we obtain

Since , we have . Hence, sends to . Thus is well defined. Then it sends to for . Thus we obtain an R-matrix

Hence we have

for some . Since the homogeneous degree of is , we obtain

Recall that the isomorphism classes of self-dual simple modules in are parametrized by the crystal basis Reference 28. The following theorem is an application of the above theorem.

Theorem 4.1.3.

Let be an element of the Grothendieck group given by

where is the self-dual simple module corresponding to and . Let be a real simple module in . Assume that we have an equality

in for some . Then commutes with and

for every such that .

Proof.

Note that we have

for some simple modules and satisfying

by Theorem 4.1.1.

We may assume that . Set

By taking the classes of self-dual simple modules with in the expansions of and , we obtain

In particular, we have .

Set

Then, by a similar argument we have .

It follows that

for every such that . Hence and commute.

Since

we have

for any such that , as desired.

Corollary 4.1.4.

Let and be simple modules. Assume that one of them is real. If and q-commute (i.e., for some ), then and commute. In particular, is simple.

The following corollary is an immediate consequence of the corollary above and Theorem 2.1.4.

Corollary 4.1.5.

Assume that the generalized Cartan matrix is symmetric and that satisfy the following conditions:

(i)

one of and is a member of the upper global basis up to a power of ,

(ii)

and q-commute.

Then their product is a member of the upper global basis of up to a power of .

4.2. Geometric results

The result of this subsection (Theorem 4.2.1) was explained to us by Peter McNamara. It will be used in the proof of the crucial result Theorem 10.3.1. In this subsection, we assume further that the base field is of characteristic .

Theorem 4.2.1 (Reference 34, Lemma 7.5).

Assume that the base field is of characteristic . Assume that has a head with a self-dual simple module and . Then we have the equality in the Grothendieck group

with self-dual simple modules and .

By duality, we obtain the following corollary.

Corollary 4.2.2.

Assume that the base field is a field of characteristic . Assume that has a socle with a self-dual simple module and . Then we have the equality in

with self-dual simple modules and .

Applying this theorem to convolution products, we obtain the following corollary.

Corollary 4.2.3.

Assume that the base field is of characteristic . Let and be simple modules. We assume that one of them is real. Then we have the equalities in as follows:

(i)

with self-dual simple modules and

(ii)

with self-dual simple modules and .

Note that is self-dual by Lemma 3.1.4.

Theorem 4.1.1 and Theorem 4.2.1 solve affirmatively Conjecture 1 of Leclerc Reference 29 in the symmetric generalized Cartan matrix case, as stated in the Introduction. More precisely, let be a symmetric KLR algebra over a base field of characteristic , and let and be simple modules over . Assume further that is real. Then by Theorem 4.1.1 and appear exactly once in a Jordan–Hölder series of . Write and for some self-dual simple modules , , and . By Theorem 4.2.1, we have

where are self-dual simple modules, and for all . Collecting the terms, we obtain

with

which proves Leclerc’s first conjecture via Theorem 2.1.4.

We obtain the following result which is a generalization of Lemma 3.2.18 in the characteristic-zero case.

Corollary 4.2.4.

Assume that the base field is of characteristic . Let and be simple modules. We assume that one of them is real. Write

with self-dual simple modules and . Then we have

4.3. Proof of Theorem 4.2.1

Recall that the graded algebra () is geometrically realized as follows Reference 40. There exist a reductive group and a -equivariant projective morphism from a smooth algebraic -variety to an affine -variety defined over the complex number field such that

Here, denotes the -equivariant derived category of the -variety with coefficient , and with

We denote by the direct sum of the constant sheaves on each connected component of , all of which are shifted by their dimensions. By the decomposition theorem Reference 1, we have a decomposition

where is a finite family of simple perverse sheaves on and is a non-zero finite-dimensional graded -vector space such that

The last fact Equation 4.1 follows from the hard Lefschetz theorem Reference 1.

Set . Then we have the multiplication morphisms

so that

has a structure of -graded algebra such that

Hence the family of the isomorphism classes of simple objects (up to a grading shift) in is . Here, is the module obtained by the algebra homomorphism , where the last arrow is the th projection. Hence we have

On the other hand, we have

Set

Then, is endowed with a natural structure of -bimodule. It is well known that the functor gives a graded Morita-equivalence

Note that and is the set of isomorphism classes of self-dual simple graded -modules by Equation 4.1.

By the above observation, in order to prove the theorem, it is enough to show the corresponding statement for the graded ring , which is obvious.

5. Quantum cluster algebras

In this section we recall the definition of skew-symmetric quantum cluster algebras following Reference 3 and Reference 11, Section 8.

5.1. Quantum seeds

Fix a finite index set with the decomposition into the set of exchangeable indices and the set of frozen indices. Let be a skew-symmetric integer-valued -matrix.

Definition 5.1.1.

We define as the -algebra generated by a family of elements with the defining relations

We denote by the skew field of fractions of .

For , we define the element of as

Here we take a total order on the set and where with . Note that does not depend on the choice of a total order of .

We have

If , then belongs to .

It is well known that is a basis of as a -module.

Let be a -algebra. We say that a family of elements of is -commuting if it satisfies for any . In such a case we can define for any . We say that an -commuting family is algebraically independent if the algebra map given by is injective.

Let be an integer-valued -matrix. We assume that the principal part of is skew-symmetric.

To the matrix we can associate the quiver without loops, -cycles, and arrows between frozen vertices such that its vertices are labeled by and the arrows are given by

Here we extend the -matrix to the skew-symmetric -matrix by setting for .

Conversely, whenever we have a quiver with vertices labeled by and without loops, -cycles, and arrows between frozen vertices, we can associate a -matrix by Equation 5.2.

We say that the pair is compatible if there exists a positive integer such that

Let be a compatible pair and a -algebra. We say that is a quantum seed in if is an algebraically independent -commuting family of elements of .

The set is called the cluster of and its elements the cluster variables. The cluster variables () are called the frozen variables. The elements () are called the quantum cluster monomials.

5.2. Mutation

For , we define a -matrix and a -matrix as follows:

The mutation of a compatible pair in direction is given by

Then the pair is also compatible with the same integer as in the case of Reference 3.

Note that for each , we have

and

Note also that we have

for with , since is compatible.

We define

and set and .

Let be a -algebra contained in a skew field . Let be a quantum seed in . Define the elements of by

Then is an algebraically independent -commuting family in . We call

the mutation of in direction . It becomes a new quantum seed in .

Definition 5.2.1.

Let be a quantum seed in . The quantum cluster algebra associated to the quantum seed is the -subalgebra of the skew field generated by all the quantum cluster variables in the quantum seeds obtained from by any sequence of mutations.

We call the initial quantum seed of the quantum cluster algebra .

6. Monoidal categorification of cluster algebras

Throughout this section, fix and a base field .

Let be a -linear abelian monoidal category. For the definition of monoidal category, see, for example, Reference 14, Appendix A.1. Note that in Reference 14, it was called the tensor category. A -linear abelian monoidal category is a -linear monoidal category such that it is abelian and the tensor functor is -bilinear and exact.

We assume further the following conditions on :

A simple object in is called real if is simple.

6.1. Ungraded cases

Definition 6.1.1.

Let be a pair of a family of simple objects in and an integer-valued -matrix whose principal part is skew-symmetric. We call a monoidal seed in if

(i)

for any ,

(ii)

is simple for any .

Definition 6.1.2.

For , we say that a monoidal seed admits a mutation in direction if there exists a simple object such that

(i)

there exist exact sequences in ,

(ii)

the pair is a monoidal seed in .

Recall that a cluster algebra with an initial seed is the -subalgebra of generated by all the cluster variables in the seeds obtained from by any sequence of mutations. Here, the mutation of a cluster variable is given by

and the mutation of is given in Equation 5.4.

Definition 6.1.3.

A -linear abelian monoidal category satisfying Equation 6.1 is called a monoidal categorification of a cluster algebra if

(i)

the Grothendieck ring is isomorphic to ,

(ii)

there exists a monoidal seed in such that is the initial seed of and admits successive mutations in all directions.

Note that if is a monoidal categorification of , then every seed in is of the form for some monoidal seed in . In particular, all the cluster monomials in are the classes of real simple objects in .

6.2. Graded cases

Let be a free abelian group equipped with a symmetric bilinear form

We consider a -linear abelian monoidal category satisfying Equation 6.1 and the following conditions:

(6.2)

(i)

We have a direct sum decomposition such that the tensor product sends to for every .

(ii)

There exists an object satisfying

(a)

there is an isomorphism

functorial in such that

commutes for any ;

(b)

the functor is an equivalence of categories.

(iii)

for any , , we have except finitely many integers .

We denote by the auto-equivalence of , and call it the grading shift functor.

In such a case the Grothendieck group is a -graded -algebra: where . Moreover, we have

where ranges over equivalence classes of simple modules. Here, two simple modules and are equivalent if for some .

For , we sometimes write and call it the weight of . Similarly, for , we write and call it the weight of .

Definition 6.2.1.

We call a quadruple a quantum monoidal seed in if it satisfies the following conditions:

(i)

is an integer-valued -matrix whose principal part is skew-symmetric,

(ii)

is an integer-valued skew-symmetric -matrix,

(iii)

is a family of elements in ,

(iv)

is a family of simple objects such that for any ,

(v)

for all ,

(vi)

is simple for any sequence in ,

(vii)

The pair is compatible in the sense of Equation 5.3 with ,

(viii)

for all ,

(ix)

for all .

Let be a quantum monoidal seed. For any and such that and , we set

and

Then . For any sequence in , we define

Then we have

For any , we have

Hence for any subset of and any set of non-negative integers , we can define .

For and , we have

Let be a quantum monoidal seed. When the -commuting family of elements of is algebraically independent, we shall define a quantum seed in by

Set

Then for any , we have

where .

For a given , we define the mutation of in direction with respect to by

Note that

Note also that satisfies conditions (viii) and (ix) in Definition 6.2.1 for any .

We have the following lemma.

Lemma 6.2.2.

Set , the mutation of as in Equation 5.6. Set . Then we have

where

Proof.

By Equation 5.1, we have

Let and be as in Equation 5.5. Because

we have

Note that It follows that

One can calculate in a similar way.

Definition 6.2.3.

We say that a quantum monoidal seed admits a mutation in direction if there exists a simple object such that

(i)

there exist exact sequences in ,

where and are as in Equation 6.3.

(ii)

is a quantum monoidal seed in .

We call the mutation of in direction .

By Lemma 6.2.2, the following lemma is obvious.

Lemma 6.2.4.

Let be a quantum monoidal seed which admits a mutation in direction . Then we have

Definition 6.2.5.

Assume that a -linear abelian monoidal category satisfies Equation 6.1 and 6.2. The category is called a monoidal categorification of a quantum cluster algebra over if

(i)

the Grothendieck ring is isomorphic to ,

(ii)

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

(iii)

admits successive mutations in all the directions.

Note that if is a monoidal categorification of a quantum cluster algebra , then any quantum seed in obtained by a sequence of mutations from the initial quantum seed is of the form for some quantum monoidal seed . In particular, all the quantum cluster monomials in are the classes of real simple objects in up to a power of .

7. Monoidal categorification via modules over KLR algebras

7.1. Admissible pair

In this section, we assume that is a symmetric KLR algebra over a base field .

From now on, we focus on the case when is a full subcategory of stable under taking convolution products, subquotients, extensions, and grading shift. In particular, we have

and we have the grading shift functor on . Hence we have

and has a -basis consisting of the isomorphism classes of self-dual simple modules.

Definition 7.1.1.

A pair is called admissible if

(i)

is a family of real simple self-dual objects of which commute with each other,

(ii)

is an integer-valued -matrix with a skew-symmetric principal part,

(iii)

for each , there exists a self-dual simple object of such that there is an exact sequence in

and commutes with for any .

Note that is uniquely determined by and . Indeed, it follows from and Reference 15, Corollary 3.7. Note also that if there is an epimorphism for some , then should coincide with by Lemma 3.1.4 and Lemma 3.2.7.

For an admissible pair , let be the skew-symmetric matrix given by . and let be the family of elements of given by .

Now we can simplify the conditions in Definition 6.2.1 and Definition 6.2.3 as follows.

Proposition 7.1.2.

Let be an admissible pair in , and let be as in Definition 7.1.1. Then we have the following properties:

(a)

The quadruple is a quantum monoidal seed in .

(b)

The self-dual simple object is real for every .

(c)

The quantum monoidal seed admits a mutation in each direction .

(d)

and are simply linked for any (i.e., ).

(e)

For any and , we have

Proof.

Item d follows from the exact sequence Equation 7.1 and Lemma 3.2.18.

Item b follows from the exact sequence Equation 7.1 by applying Corollary 3.2.21 to the case

Item e follows from

and

Let us show a. The conditions (i)–(v) in Definition 6.2.1 are satisfied by the construction. The condition (vi) follows from Proposition 3.2.5 and the fact that is real simple for every . The condition (viii) is nothing but Lemma 3.1.2. The condition (ix) follows easily from the fact that the weights of the first and the last terms in the exact sequence Equation 7.1 coincide.

Let us show the condition (vii) in Definition 6.2.1. By Equation 7.2 and d of this proposition, we have

for and . Thus we have shown that is a quantum monoidal seed in .

Let us show c. Let . The exact sequence Equation 6.4 follows from Equation 7.1 and the equality

which is an immediate consequence of Equation 7.2.

Similarly, taking the dual of the exact sequence Equation 7.1, we obtain an exact sequence

which gives the exact sequence Equation 6.5.

It remains to prove that is a quantum monoidal seed in for any .

We see easily that satisfies the conditions (i)–(iv) and (vii)–(ix) in Definition 6.2.1.

For the condition (v), it is enough to show that for we have

where for and . In the case and , we have

The other cases follow from Equation 7.2.

The condition (vi) in Definition 6.2.1 for follows from Proposition 3.2.5 and the fact that is a commuting family of real simple modules.

Now we are ready to give one of our main theorems.

Theorem 7.1.3.

Let be an admissible pair in and set

as in Proposition 7.1.2. We set . We assume further that

Then, for each , the pair is admissible in .

Proof.

In Proposition 7.1.2 (b), we have already shown that the condition (i) in Definition 7.1.1 holds for . The condition (ii) is clear from the definition. Let us show (iii). Set and for and . It is enough to show that, for any , there exists a self-dual simple module such that there is a short exact sequence

and

If , then , and hence satisfies the desired condition.

Assume that and . Then for any and for any . Hence satisfies the desired condition.

We will show the assertion in the case . We omit the proof of the case because it can be shown in a similar way.

Recall that we have

for different from and .

Set

Then using Equation 7.6 repeatedly, we have

Set

and set

Then Equation 7.5 is read as

Note that we have

Taking the convolution products of and Equation 7.9, we obtain

Since commutes with , we have

On the other hand, we have

Note that we used the compatibility of the pair when we derive the equality (a).

Since , the equality implies

Hence the following diagram is commutative by Proposition 3.2.8 (i):

where . Note that since commutes with and , is an isomorphism. Hence we have

Therefore we obtain an exact sequence

On the other hand, decomposes (up to a grading shift) by Lemma 3.1.5 as follows:

Since commutes with , the homomorphisms is an isomorphism, and hence we have

Similarly, decomposes (up to a grading shift) as follows:

Since commutes with , the homomorphism is an isomorphism, and hence we have

On the other hand, Lemma 3.2.22 implies that

and hence we obtain

Thus the exact sequence Equation 7.10 becomes the exact sequence in ,

for some . Since is self-dual, . On the other hand, by Proposition 3.2.13 (i) and Proposition 7.1.2 (d), we have

By the exact sequence Equation 7.11, is not simple, and we conclude

Then Lemma 3.2.18 implies that . Thus we obtain an exact sequence in ,

Now we shall rewrite Equation 7.12 by using instead of . We have

On the other hand, the exact sequence Equation 7.9 gives

It follows that

Hence we have

where by Lemma 3.1.4.

Thus we obtain the identity in ,

On the other hand, the hypothesis Equation 7.4 implies that there exists corresponding to so that it satisfies

and

where .

Hence, in , we have

Since is a domain, we conclude that

On the other hand, Equation 7.14 implies

Hence, Theorem 4.1.3 implies that, when we write

we have

In particular, each module with is simple because is a real simple module. Thus we obtain

Since is simple, there exists such that is isomorphic to up to a grading shift, and for . Set . Then we conclude that so that

We emphasize that is a self-dual simple module in which satisfies that up to a grading shift.

Now Equation 7.13 implies

Hence there exists an exact sequence

where and or and . By Lemma 3.2.18, the second case does not occur, and we have an exact sequence

Since , , and are self-dual, we have , and we obtain the desired short exact sequence Equation 7.7.

Since commutes with up to a power of in , and is real simple, commutes with for , by Corollary 4.1.4.

Corollary 7.1.4.

Let be an admissible pair in . Under the assumption Equation 7.4, is a monoidal categorification of the quantum cluster algebra . Furthermore, the following statements hold:

(i)

The quantum monoidal seed admits successive mutations in all directions.

(ii)

Any cluster monomial in is the isomorphism class of a real simple object in up to a power of .

(iii)

Any cluster monomial in is a Laurent polynomial of the initial cluster variables with a coefficient in .

Proof.

Items (i) and (ii) are straightforward.

Let us show (iii). Let be a cluster monomial. By the Laurent phenomenon Reference 3, we can write

where is the initial cluster, , and . Since and are the isomorphism classes of simple modules up to a power of , their product can be written as a linear combination of the isomorphism classes of simple modules with coefficients in . Since every is the isomorphism class of a simple module up to a power of , we have .

8. Quantum coordinate rings and modified quantized enveloping algebras

8.1. Quantum coordinate ring

Let be . Then the comultiplication (see Equation 1.1) induces the multiplication on as follows:

Later on, it will be convenient to use Sweedler’s notation . With this notation,

The -bimodule structure on induces a -bimodule structure on . Namely,

Then the multiplication is a morphism of a -bimodule, where has the structure of a -bimodule via . That is, for and , we have

where and .

Definition 8.1.1.

We define the quantum coordinate ring as follows:

Then, is a subring of because (i) is -bilinear, and (ii) and are closed under the tensor product.

We have the weight decomposition , where

For , we write

For any , we have the -bilinear homomorphism

given by

Proposition 8.1.2 (Reference 17, Proposition 7.2.2).

We have an isomorphism of -bimodules

given by . Namely,

We introduce the crystal basis of as the images by of

Hence it is a crystal base with respect to the left action of and also the right action of . We sometimes write by and the operators of obtained by the right actions of and .

We define the -form of by

We define the bar-involution of by

Note that the bar-involution is not a ring homomorphism but it satisfies

Since we do not use this formula and it is proved similarly to Proposition 8.1.4 below, we omit its proof.

The triple is balanced Reference 17, Theorem 1, and hence there exists an upper global basis of ,

For and , we denote by the unique member of the upper global basis of with weight . It is also a member of the lower global basis.

Proposition 8.1.3.

Let , , and . Then, is a member of the upper global basis of .

Proof.

The element is bar-invariant and a member of crystal basis modulo . For any ,

belongs to because and . Hence belongs to .

The -algebra anti-automorphism of induces a -linear automorphism of by

We have

and

It is obvious that preserves , , and .

Proposition 8.1.4.

In order to prove this proposition, we prepare a sublemma.

Let be the -algebra automorphism of given by

We can easily see

Let be the automorphism of given by

Sublemma 8.1.5.

We have

where .

Proof.

Let us show that, for each , the following equality,

holds for any .

The equality Equation 8.2 is obviously true for . If Equation 8.2 is true for , then

Since , Equation 8.2 holds for . Similarly if Equation 8.2 holds for , then it holds for .

Proof of Proposition 8.1.4.

We have

Hence, we have

It follows that

Therefore we obtain

with

8.2. Unipotent quantum coordinate ring

Let us endow with the algebra structure defined by

Let be the algebra homomorphism given by

Set

Defining the bilinear form by

we get an algebra structure on given by

where .

Since has a -bimodule structure, so does .

We define the -form of by

and define the bar-involution on by

Note that the bar-involution is not a ring homomorphism but it satisfies

For , we denote by the right action of on .

Lemma 8.2.1.

For , we have -boson relations

Proof.

If we set , then we have

Hence, we have

The second identity follows in a similar way.

We define the map by

Since is a non-degenerate bilinear form on , is injective. The relation

implies that

Lemma 8.2.2.

is an algebra isomorphism.

Proof.

The map is an algebra homomorphism because and both satisfy the same -boson relation.

Hence, the algebra has an upper crystal basis such that . Furthermore, has an upper global basis

induced by the balanced triple (see Equation 1.3).

There exists an injective map

induced by the -linear homomorphism given by

The map commutes with . We have

Remark 8.2.3.

Note that the multiplication on given in Reference 11 is different from ours. Indeed, by denoting the product of and in Reference 11, Section 4.2 by , for , we have

where for . By Lemma 8.5.3 below, we have

for , where . In particular, we have a -algebra isomorphism from to given by

Note also that the bar-involution is a ring anti-isomorphism between and .

8.3. Modified quantum enveloping algebra

For the materials in this subsection we refer the reader to Reference 19Reference 32. We denote by the category of left -modules with the weight space decomposition. Let be the functor from to the category of vector spaces over , forgetting the -module structure.

Let us denote by the endomorphism ring of . Note that contains . For , let denote the projector to the weight space of weight . Then the defining relation of (as a left -module) is

We have

Then is isomorphic to . We set

Then is a subalgebra of . We call it the modified quantum enveloping algebra. Note that any -module in has a natural -module structure.

The (anti-)automorphisms , , and of extend to the ones of by

For a dominant integral weight , let us denote by (resp. ) the irreducible module with highest (resp. lowest) weight (resp. ). Let (resp. ) be the highest (resp. lowest) weight vector.

For , , we set

Then is generated by as a -module, and the defining relation of is

Let us define the -linear automorphism of by

We set

(i)

,

(ii)

,

(iii)

.

Proposition 8.3.1 (Reference 32).

is a lower crystal basis of . Furthermore, is balanced, and there exists a lower global basis obtained from the lower crystal basis .

Theorem 8.3.2 (Reference 32).

The algebra has a lower crystal basis satisfying the following properties:

(i)

and , where

and

.

(ii)

Set . Then is balanced, and has the lower global basis .

(iii)

For any and , let

be the -linear map . Then we have

(iv)

Let be the induced homomorphism

Then we have

(a)

,

(b)

for any .

(v)

has a structure of crystal such that the injective map induced by (iv) (a)

is a strict embedding of crystals for any and .

For , take any and such that . Then is embedded into .

For , let be the crystal with

Since we have

is embedded into the crystal . Taking and , we have

Lemma 8.3.3 (Reference 19).

For any , we have a canonical crystal isomorphism

Hence we identify

For and , we shall denote by

Then for any , , and , we have

Reference 19, (3.1.1). In particular, we have

Theorem 8.3.4 (Reference 19).
(i)

is invariant under the anti-automorphisms and .

(ii)

.

(iii)

and for .

Corollary 8.3.5 (Reference 19).

For , , we have

(1)

.

(2)

.

We define, for with , or ,

This defines another crystal structure on : For , , and , we have

In particular, we have

8.4. Relationship of and

There exists a canonical pairing by

Theorem 8.4.1 (Reference 19).

There exists a bi-crystal embedding

which satisfies

for any and .

8.5. Relationship of and

Definition 8.5.1.

Let be the homomorphism induced by ,

Then we have

It is obvious that sends all ( and ) to . Note that .

Proposition 8.5.2.

For , set

. Then we have

Proof.

Set . Then for any , we have

Hence the map sends the upper global basis of to the upper global basis of or zero. Thus we have a map

Although the map is not an algebra homomorphism, it preserves the multiplications up to a power of , as we will see below.

Lemma 8.5.3.

For , if , then

Proof.

Assume that Equation 8.5 holds for . Note that

On the other hand, we have

Hence Equation 8.5 holds for .

Proposition 8.5.4.

For , we have

Proof.

For , set . Then, we have

Here, we used in (a).

8.6. Global basis of and tensor products of -modules in

Let be an integrable -module with a bar-involution ; that is, there is a -linear automorphism satisfying for all and for all . Then, for any , there exists a unique bar-involution on satisfying

Indeed, there exists , which defines a bar-involution by setting

(see Reference 33, Chapter 4). Assume that has a lower crystal basis and an -form such that is balanced. Then we have

Proposition 8.6.1.

The triple in is balanced.

Note that is a lower global basis for any , i.e., .

In particular, it applies to . Moreover, we have the following proposition.

Proposition 8.6.2.

Let and . Then for any , vanishes or is a member of the lower global basis of .

Hence we have a crystal morphism

by .

Similarly, we have a bar-involution on such that

Hence if has an upper crystal basis and an -form such that is balanced, then has an upper global basis. Note that is a member of the upper global basis for .

In particular for , has a lower global basis and has an upper global basis.

The bilinear form

defined by , , satisfies

With respect to this bilinear form, the lower global basis of and the upper global basis of are the dual bases of each other.

9. Quantum minors and -systems

9.1. Quantum minors

Using the isomorphism in Equation 8.1, for each and , we define the elements

and

The element is called a (generalized) quantum minor and is called a unipotent quantum minor.

Lemma 9.1.1.

is a member of the upper global basis of . Moreover, is either a member of the upper global basis of or zero.

Proof.

Our assertions follow from Proposition 8.1.3 and Proposition 8.5.2.

Lemma 9.1.2 (Reference 3, (9.13)).

For and , we have

By Proposition 8.5.4, we have the following corollary.

Corollary 9.1.3.

For and , we have

Note that

Then if and only if . Recall that for in the same -orbit, we say that if there exists a sequence of positive real roots such that, defining , , we have and .

More precisely, we have the following lemma.

Lemma 9.1.4.

Let and . Then the following conditions are equivalent:

(i)

is an element of the upper global basis of ,

(ii)

,

(iii)

,

(iv)

,

(v)

,

(vi)

for any such that , there exists (in the Bruhat order) such that ,

(vii)

there exist such that , , and .

Proof.

(i) and (ii) are equivalent by Lemma 9.1.1. The equivalence of (ii), (iii), and (iv) is obvious. The equivalence of (v), (vi), and (vii) is well known. The equivalence of (iv) and (vi) is proved in Reference 18.

For any and , we set

Then for any , we have

Lemma 9.1.5.

Let , such that and .

(i)

If , then

(ii)

If and , then .

(iii)

If , then

(iv)

If and , then .

Proof.

We have and . Moreover, commutes with and .

Let us show (ii). Set . Then we have , which implies . Hence . We have

Hence we have . By the assumption , does not vanish. Hence we have .

The other statements can be proved similarly.

Proposition 9.1.6 (Reference 3, (10.2)).

Let and such that and . Then we have

(i)

.

(ii)

If we assume further that and , then we have

or equivalently

Note that (ii) follows from Proposition 8.5.4 and (i). Note also that both sides of Equation 9.2 are bar-invariant, and hence they are members of the upper global basis as seen by Corollary 4.1.5.

Proposition 9.1.7.

For and , set and with . Then we have

Proof.

Recall that there is a pairing defined by . It satisfies

For and , we have

Hence for , we have

If for , then we have

The element vanishes or is a global basis of by Proposition 8.6.2. Since is a member of the upper global basis of , we have

Here is the crystal morphism given in Equation 8.6.

Hence we obtain

where is a unique element such that

. On the other hand, we have and . The last equality implies because

As seen in Equation 8.4, we have

Hence we obtain

In the last equality, we used and .

Hence we conclude that .

Let

be the canonical embedding and

the induced crystal embedding.

Lemma 9.1.8.

For and such that , we have

Proof.

Let us show by induction on the length of in . We may assume that . Then there exists such that . If , then and . If , then and . In both cases, is connected with an element of .

Lemma 9.1.9.

For and , we have

Proof.

We have

where . Hence Proposition 9.1.7 implies that

Then, gives the desired result.

9.2. -system

In this subsection, we recall the -system among the (unipotent) quantum minors for later use (see Reference 25 for -system).

Proposition 9.2.1 (Reference 11, Proposition 3.2).

Assume that the Kac–Moody algebra is of symmetric type. Assume that and satisfy and . Then

and

where .

Note that the difference of and are -invariant. Hence we have from Corollary 9.1.3, by disregarding a power of .

9.3. Revisit of crystal bases and global bases

In order to prove Theorem 9.3.3 below, we first investigate the upper crystal lattice of induced by an upper crystal lattice of .

Let be a -module in . Let be an upper crystal lattice of . Then we have (see Lemma 1.3.1)

Recall that, for , the upper crystal lattice and the lower crystal lattice of are related by

Write

with finite-dimensional -vector spaces . Accordingly, we have a canonical decomposition

where is an -lattice of .

On the other hand, we have

Note that we have

We define the induced upper crystal lattice of by

where . Then we have

Indeed, we have

Since , we have

The properties and imply the following lemma.

Lemma 9.3.1.

is the largest upper crystal lattice of contained in the lower crystal lattice .

Let . Then is a lower crystal lattice of . Let be the -module isomorphism defined by

Then

is an upper crystal lattice of . Since we have for any and , Lemma 9.3.1 implies that

Lemma 9.3.2.

Let and such that . Then we have

Proof.

By the definition, we have

Hence we have

The right-hand side of Equation 9.5 can be calculated as follows. Let us take such that for . Here denotes the canonical -module homomorphism and such a exists by Lemma 9.1.8.

Then we have

The last equality follows from for all .

On the other hand, we have

and

Hence

by Equation 9.4, as desired.

Theorem 9.3.3.

Let and such that . Then we have

Proof.

Applying to Equation 9.5, we have

Hence the desired result follows from Proposition 8.5.2, Proposition 8.5.4, and Lemma 9.3.4 below.

Lemma 9.3.4.

Let and such that . Then we have

Proof.

We shall argue by induction on . We set . Since the case is obvious, assume that . Take such that .

(a)

First assume that . Then we have . Hence by the induction hypothesis,

We have and . Hence, applying to Equation 9.6, we obtain

(b)

Assume that . Then we have , and the induction hypothesis implies that

Apply to both sides. Then the left-hand side yields

Since , the right-hand side yields

Since and , we have

9.4. Generalized -system

The -system in Section 9.2 can be interpreted as a system of equations among the three products of elements in or . In this subsection, we introduce another among the three products of elements in , called a generalized -system.

Proposition 9.4.1.

Let , and set . Then we have

Note that if , then and the last term in Equation 9.7 vanishes. If , then and , .

Proof.

In the sequel, we omit for the sake of simplicity. Set

Then .

It is obvious that we have for . Since , we have

Here the second equality follows from Lemma 9.1.9 and the third follows from Proposition 8.1.3. On the other hand, we have

Hence we have . Thus, is a lowest weight vector of weight with respect to the right action of . Therefore there exists some such that

Hence we have . On the other hand, we have

Note that since and for , we have .

Hence in order to prove our assertion, it is enough to show that

This follows from Proposition 8.5.2 and

Let us prove Equation 9.8. Since

the left-hand side of Equation 9.8 is equal to

Since , we obtain

10. KLR algebras and their modules

10.1. Chevalley and Kashiwara operators

Let us recall the definition of several functors on modules over KLR algebras which are used to categorify .

Definition 10.1.1.

Let .

(i)

For and , set

(ii)

We take conventions

which are functors from to .

(iii)

For a simple module , we set

(iv)

For and , we set

Here denotes the -module . Then is a self-dual real simple -module.

Note that, under the isomorphism in Theorem 2.1.2, the functors and correspond to the linear operators and on , respectively. Note also that, for a simple -module , we have , and if .

In the course of proving the following propositions, we use the following notations:

Then we have

Proposition 10.1.2.

Let with . Assume that an -module satisfies . Then the left -module homomorphism given by

extends uniquely to an -bilinear homomorphism

Proof.
(i)

First note that, for ,

since .

(ii)

In order to see that Equation 10.3 is a well-defined -linear homomorphism, it is enough to show that Equation 10.2 is -linear.

(a)

Commutation with : We have

by Equation 10.4.

(b)

Commutation with : We have

The last term vanishes because implies

for any polynomial and .

(iii)

Now let us show that Equation 10.3 is right -linear. By Equation 10.4, we have

for . Therefore we have

Recall that for , we denote by the element of defined by

Set , where is a reduced expression of . Note that does not depend on the choice of reduced expression Reference 14, Corollary 1.4.3.

Proposition 10.1.3.

Let and , and set and . If for any , then

gives a well-defined -linear homomorphism .

Proof.

The proceeding proposition implies that

gives a well-defined -linear homomorphism . Hence it is enough to show that it is right -linear. Since we know that it commutes with the right multiplication of , it is enough to show that it commutes with the right multiplication of . For this, we may assume that and . Set .

Thus we have reduced the problem to the equality

for , which is a consequence of

for . Note that

and

Hence it is enough to show

This follows from

for and , .

Let be a projective cover of . Define the functor

by

where denotes the right -module obtained from the left -module via the anti-automorphism . We define the functor in a similar way. Note that

Corollary 10.1.4.

Let be a symmetric KLR algebra. Let and a simple module. Then we have

Proof.

Set and . Then the preceding proposition implies . Hence we have , which implies

Proposition 10.1.5.

Let , be modules and .

(i)

If and , then we have

(ii)

If and , then we have

Proof.

Our assertions follow from the shuffle lemma Reference 21, Lemma 2.20.

The following corollaries are immediate consequences of Proposition 10.1.5.

Corollary 10.1.6.

Let , and let be a real simple module. Then is also real simple.

Corollary 10.1.7.

Let , and let be a simple module with . Then we have .

Proposition 10.1.8.

Let and be simple modules. We assume that one of them is real. If , then we have an isomorphism in

Similarly, if , then we have

Proof.

Set and . Then or is real. Now we have

which induces a non-zero map . Hence we have a surjective map

Since or is real by Corollary 10.1.6, has a simple head and we obtain the desired result. A similar proof works for the second statement.

10.2. Determinantial modules and -system

We will use the materials in Section 9 to obtain properties on the determinantial modules.

In the rest of this paper, we assume that is symmetric and the base field is of characteristic . Under this condition, the family of self-dual simple -modules corresponds to the upper global basis of by Theorem 2.1.4.

Let be the map from to obtained by composing and the isomorphism Equation 2.2 in Theorem 2.1.2.

Definition 10.2.1.

For and such that , let be a simple -module such that .

Since is a member of the upper global basis, such a module exists uniquely due to Theorem 2.1.4. The module is self-dual, and we call it the determinantial module.

Lemma 10.2.2.

is a real simple module.

Proof.

It follows from which is a member of the upper global basis up to a power of . Here the last equality follows from Corollary 9.1.3.

Proposition 10.2.3.

Let , and such that , , , and . Then

(i)

and commute,

(ii)

,

(iii)

,.

Proof.

It is a consequence of Proposition 9.1.6 (ii) and Corollary 4.1.4.

Proposition 10.2.4.

Let , such that and .

(i)

If , then

(ii)

If and , then .

(iii)

If , then

(iv)

If and , then .

Proof.

It is a consequence of Lemma 9.1.5.

Proposition 10.2.5.

Assume that and satisfy and .

(i)

We have exact sequences

and

where .

(ii)

.

Proof.

Since the proof of Equation 10.6 is similar, let us only prove Equation 10.7. (Indeed, they are dual to each other.)

Set

Then Proposition 9.2.1 implies that

Since and are simple and self-dual, our assertion follows from Lemma 3.2.19.

10.3. Generalized -system on determinantial module

Theorem 10.3.1.

Let and such that . Then there exists a canonical epimorphism

which is equivalent to saying that .

In particular, we have

Proof.
(a)

Our assertion follows from Theorem 9.3.3 and Theorem 4.2.1 when .

(b)

We shall prove the general case by induction on . By (a), we may assume that . Then there exists such that . The induction hypothesis yields that

Since , Proposition 10.2.4 (iv) gives

Then Proposition 10.1.8 implies that

from which we obtain

By Lemma 3.1.4, we have . Hence we obtain

Proposition 10.3.2.

Let and .

(i)

If , , , and , then we have

(ii)

If , , , and , then we have

Proof.

In the course of proof, we omit for the sake of simplicity. If , then the assertion follows from Proposition 10.2.3 (i). Hence we may assume that .

Let us show (i). By Proposition 9.4.1, we have

where and . Let be the operator on given by the application of from the right, where is a reduced expression of and . Then applying to Equation 10.8, we obtain

Recall that is called -dominant if . Here is a reduced expression of and (). Recall that an element with is called -highest if is -dominant and

If is -highest, then is a linear combination of -highest ’s. Moreover, is either a member of the upper global basis or zero. Since is -highest of weight , we obtain

Applying , we obtain

for some integer . Hence we obtain (i) by Lemma 3.2.19 (i).

(ii) is proved similarly. By applying to Equation 10.8, we obtain

Here we used Proposition 8.1.4. Then the similar arguments as above show (ii).

Proposition 10.3.3.

Let such that and . Then we have

Proof.

By Proposition 10.3.2 (ii), we have . Assuming , let us derive a contradiction.

By Theorem 10.3.1 and the assumption, we have

Hence we have

for any . Since , Proposition 10.2.4 implies that

It implies that

It is a contradiction since does not vanish.

11. Monoidal categorification of

11.1. Quantum cluster algebra structure on

In this subsection, we shall consider the Kac–Moody algebra associated with a symmetric Cartan matrix . We shall recall briefly the definition of the subalgebra of and its quantum cluster algebra structure by using the results of Reference 11 and Reference 23. Remark that we bring the results in Reference 11 through the isomorphism Equation 8.3.

For a given , fix a reduced expression .

For and , we set

We set

and

Note that , if . For , we set

The -subalgebra of generated by () is independent of the choice of a reduced expression of . We denote it by . Then every () is contained in Reference 11, Corollary 12.4. The set is a basis of as a -vector space Reference 23, Theorem 4.2.5. We call it the upper global basis of . We denote by the -module generated by . Then it is a -subalgebra of Reference 23, Section 4.7.2. We set .

Let , , and .

Definition 11.1.1.

We define the quiver with the set of vertices and the set of arrows as follows:

,

There are two types of arrows:

Let be the integer-valued -matrix associated to the quiver by Equation 5.2.

Lemma 11.1.2.

Assume that and

when ,

when .

Then and -commute; that is, there exists such that

Proof.

We may assume . Let be as in Equation 11.1. Take , , , and . Then we have

From Proposition 9.1.6, our assertion follows.

Hence we have an integer-valued skew-symmetric matrix such that

Proposition 11.1.3 (Reference 11, Proposition 10.1).

The pair is compatible with in Equation 5.3.

Theorem 11.1.4 (Reference 11, Theorem 12.3).

Let be the quantum cluster algebra associated to the initial quantum seed . Then we have an isomorphism of -algebras

where and .

11.2. Admissible seeds in the monoidal category

For , we set . It is a real simple module with .

Definition 11.2.1.

For , let be the smallest monoidal abelian full subcategory of satisfying the following properties:

(i)

is stable under the subquotients, extensions, and grading shifts,

(ii)

contains for all .

Then by Reference 11, belongs to if and only if belongs to . Hence we have a -algebra isomorphism

We set

Then, by Proposition 10.2.3, is a quantum monoidal seed in . We are now ready to state the main theorem in this section.

Theorem 11.2.2.

The pair is admissible.

As we already explained, combined with Theorem 7.1.3 and Corollary 7.1.4, this theorem implies the following theorem.

Theorem 11.2.3.

The category is a monoidal categorification of the quantum cluster algebra .

In the course of proving Theorem 11.2.2, we omit grading shifts if there is no danger of confusion.

We shall start the proof of Theorem 11.2.2 by proving that, for each , there exists a simple module such that

(11.2)

(a)

there exists a surjective homomorphism (up to a grading shift)

(b)

there exists a surjective homomorphism (up to a grading shift)

(c)

.

We set

Then is a real simple module.

Now we claim that the following simple module satisfies the conditions in 11.2:

Let us show 11.2 (a). The incoming arrows to are

for ,

.

Hence we have

Then the morphism in (a) is obtained as the composition,

Here the second epimorphism is given in Theorem 10.3.1, and Lemma 3.1.5 asserts that the composition Equation 11.3 is non-zero and hence an epimorphism.

Let us show 11.2 (b). The outgoing arrows from are

 for .

 if .

Hence we have

Lemma 11.2.4.

There exists an epimorphism (up to a grading)

Proof.

By the dual of Theorem 10.3.1 and the -system Equation 10.7 with , , and , we have morphisms

Here the last isomorphism follows from the fact that for any .

Thus we have a sequence of morphisms

By Lemma 3.1.5 (i), the composition is non-zero.

Since , Theorem 10.3.1 gives the morphisms

Here we have used Lemma 3.2.22 to obtain the morphism . Note that the module is simple. By applying Lemma 3.1.5 once again, is non-zero, and hence it is an epimorphism.

Lemma 11.2.5.

We have .

Proof.

Since and commute and by Proposition 10.3.3, we have

by Proposition 3.2.10 and Lemma 3.2.3. If and commute, then 11.2 (a) would imply that belongs to . It contradicts the result in Reference 10 that all the ’s are prime at .

Proposition 11.2.6.

The map factors through ; that is,

Here is the canonical surjection.

Proof.

We have by Lemma 11.2.5, and

by Proposition 10.3.3 with . Hence has a simple head by Proposition 3.2.16 (iii).

End of the proof of Theorem 11.2.2.

By the above arguments, we have proved the existence of which satisfies 11.2. By Proposition 3.2.17 and 11.2 (c), has composition length . Moreover, it has a simple socle and simple head. On the other hand, taking the dual of 11.2 (a), we obtain a monomorphism

in . Together with 11.2 (b), there exists a short exact sequence in :

for some . By Lemma 3.2.18 must be equal to .

It remains to prove that commutes with (). For any , we have

and

Hence we have

We conclude that commutes with if . Thus we complete the proof of Theorem 11.2.2.

As a corollary, we prove the following conjecture on the cluster monomials.

Theorem 11.2.7 (Reference 11, Conjecture 12.9, Reference 23, Conjecture 1.1(2)).

Every cluster variable in is a member of the upper global basis up to a power of .

Theorem 11.2.2 also implies Reference 11, Conjecture 12.7 in the refined form as follows.

Corollary 11.2.8.

has a quantum cluster algebra structure associated with the initial quantum seed

i.e.,

Acknowledgements

The authors would like to express their gratitude to Peter McNamara who informed us of his result. They would also like to express their gratitude to Bernard Leclerc and Yoshiyuki Kimura for many fruitful discussions. The third and fourth authors gratefully acknowledge the hospitality of Research Institute for Mathematical Sciences, Kyoto University, during their visits in 2014.

Table of Contents

  1. Abstract
  2. Introduction
    1. Conjecture (11, Conjecture 12.9, 23, Conjecture 1.1(2)).
    2. Theorem (29, Conjecture 1).
    3. Main Theorem 1 (Theorem 7.1.3 and Corollary 7.1.4).
    4. Main Theorem 2 (Theorem 11.2.2).
  3. 1. Quantum groups and global bases
    1. 1.1. Quantum groups
    2. Definition 1.1.1.
    3. 1.2. Integrable representations
    4. 1.3. Crystal bases and global bases
    5. Lemma 1.3.1.
    6. Definition 1.3.2.
  4. 2. KLR algebras and R-matrices
    1. 2.1. KLR algebras
    2. Definition 2.1.1.
    3. Theorem 2.1.2 (2138).
    4. Definition 2.1.3.
    5. Theorem 2.1.4 (3940).
    6. 2.2. R-matrices for KLR algebras
    7. Definition 2.2.1.
    8. Definition 2.2.2.
    9. Lemma 2.2.3 (15, Lemma 3.1).
    10. Theorem 2.2.4 (15, Theorem 3.2).
    11. Definition 2.2.5.
  5. 3. Simplicity of heads and socles of convolution products
    1. 3.1. Homogeneous degrees of R-matrices
    2. Definition 3.1.1.
    3. Lemma 3.1.2.
    4. Definition 3.1.3.
    5. Lemma 3.1.4.
    6. Lemma 3.1.5.
    7. 3.2. Properties of and
    8. Lemma 3.2.1.
    9. Definition 3.2.2.
    10. Lemma 3.2.3 (15).
    11. Definition 3.2.4.
    12. Proposition 3.2.5.
    13. Definition 3.2.6.
    14. Lemma 3.2.7.
    15. Proposition 3.2.8.
    16. Proposition 3.2.9.
    17. Proposition 3.2.10.
    18. Lemma 3.2.11.
    19. Lemma 3.2.12.
    20. Proposition 3.2.13.
    21. Proposition 3.2.14.
    22. Proposition 3.2.15.
    23. Proposition 3.2.16.
    24. Proposition 3.2.17.
    25. Lemma 3.2.18.
    26. Lemma 3.2.19.
    27. Proposition 3.2.20.
    28. Corollary 3.2.21.
    29. Lemma 3.2.22.
  6. 4. Leclerc’s conjecture
    1. 4.1. Leclerc’s conjecture
    2. Theorem 4.1.1.
    3. Corollary 4.1.2.
    4. Theorem 4.1.3.
    5. Corollary 4.1.4.
    6. Corollary 4.1.5.
    7. 4.2. Geometric results
    8. Theorem 4.2.1 (34, Lemma 7.5).
    9. Corollary 4.2.2.
    10. Corollary 4.2.3.
    11. Corollary 4.2.4.
    12. 4.3. Proof of Theorem 4.2.1
  7. 5. Quantum cluster algebras
    1. 5.1. Quantum seeds
    2. Definition 5.1.1.
    3. 5.2. Mutation
    4. Definition 5.2.1.
  8. 6. Monoidal categorification of cluster algebras
    1. 6.1. Ungraded cases
    2. Definition 6.1.1.
    3. Definition 6.1.2.
    4. Definition 6.1.3.
    5. 6.2. Graded cases
    6. Definition 6.2.1.
    7. Lemma 6.2.2.
    8. Definition 6.2.3.
    9. Lemma 6.2.4.
    10. Definition 6.2.5.
  9. 7. Monoidal categorification via modules over KLR algebras
    1. 7.1. Admissible pair
    2. Definition 7.1.1.
    3. Proposition 7.1.2.
    4. Theorem 7.1.3.
    5. Corollary 7.1.4.
  10. 8. Quantum coordinate rings and modified quantized enveloping algebras
    1. 8.1. Quantum coordinate ring
    2. Definition 8.1.1.
    3. Proposition 8.1.2 (17, Proposition 7.2.2).
    4. Proposition 8.1.3.
    5. Proposition 8.1.4.
    6. Sublemma 8.1.5.
    7. 8.2. Unipotent quantum coordinate ring
    8. Lemma 8.2.1.
    9. Lemma 8.2.2.
    10. Remark 8.2.3.
    11. 8.3. Modified quantum enveloping algebra
    12. Proposition 8.3.1 (32).
    13. Theorem 8.3.2 (32).
    14. Lemma 8.3.3 (19).
    15. Theorem 8.3.4 (19).
    16. Corollary 8.3.5 (19).
    17. 8.4. Relationship of and
    18. Theorem 8.4.1 (19).
    19. 8.5. Relationship of and
    20. Definition 8.5.1.
    21. Proposition 8.5.2.
    22. Lemma 8.5.3.
    23. Proposition 8.5.4.
    24. 8.6. Global basis of and tensor products of -modules in
    25. Proposition 8.6.1.
    26. Proposition 8.6.2.
  11. 9. Quantum minors and -systems
    1. 9.1. Quantum minors
    2. Lemma 9.1.1.
    3. Lemma 9.1.2 (3, (9.13)).
    4. Corollary 9.1.3.
    5. Lemma 9.1.4.
    6. Lemma 9.1.5.
    7. Proposition 9.1.6 (3, (10.2)).
    8. Proposition 9.1.7.
    9. Lemma 9.1.8.
    10. Lemma 9.1.9.
    11. 9.2. -system
    12. Proposition 9.2.1 (11, Proposition 3.2).
    13. 9.3. Revisit of crystal bases and global bases
    14. Lemma 9.3.1.
    15. Lemma 9.3.2.
    16. Theorem 9.3.3.
    17. Lemma 9.3.4.
    18. 9.4. Generalized -system
    19. Proposition 9.4.1.
  12. 10. KLR algebras and their modules
    1. 10.1. Chevalley and Kashiwara operators
    2. Definition 10.1.1.
    3. Proposition 10.1.2.
    4. Proposition 10.1.3.
    5. Corollary 10.1.4.
    6. Proposition 10.1.5.
    7. Corollary 10.1.6.
    8. Corollary 10.1.7.
    9. Proposition 10.1.8.
    10. 10.2. Determinantial modules and -system
    11. Definition 10.2.1.
    12. Lemma 10.2.2.
    13. Proposition 10.2.3.
    14. Proposition 10.2.4.
    15. Proposition 10.2.5.
    16. 10.3. Generalized -system on determinantial module
    17. Theorem 10.3.1.
    18. Proposition 10.3.2.
    19. Proposition 10.3.3.
  13. 11. Monoidal categorification of
    1. 11.1. Quantum cluster algebra structure on
    2. Definition 11.1.1.
    3. Lemma 11.1.2.
    4. Proposition 11.1.3 (11, Proposition 10.1).
    5. Theorem 11.1.4 (11, Theorem 12.3).
    6. 11.2. Admissible seeds in the monoidal category
    7. Definition 11.2.1.
    8. Theorem 11.2.2.
    9. Theorem 11.2.3.
    10. Lemma 11.2.4.
    11. Lemma 11.2.5.
    12. Proposition 11.2.6.
    13. Theorem 11.2.7 (11, Conjecture 12.9, 23, Conjecture 1.1(2)).
    14. Corollary 11.2.8.
  14. Acknowledgements

Mathematical Fragments

Equation (0.2)
Equations (1.1), (1.2)
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 .

Equation (1.3)
Equation (2.1)
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

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.

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 .

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 .

Lemma 3.1.2.

For non-zero -modules and , we have

Lemma 3.1.4.

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

is a self-dual simple module.

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

Lemma 3.2.1.

Let and be simple -modules. Then we have

(i)

(ii)

If for some , then

up to constant multiples.

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.

Proposition 3.2.5.

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

is a real simple module.

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 .

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

Proposition 3.2.9.

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

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 .

Lemma 3.2.11.

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

commutes.

Lemma 3.2.12.

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

commutes.

Proposition 3.2.13.

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

(i)

If and commute, then

(ii)

If and commute, then

Proposition 3.2.14.

Let be a real simple module, and let be a module with a simple socle. If the following diagram

commutes up to a non-zero constant multiple, then is equal to the socle of . In particular, has a simple socle.

Proposition 3.2.15.

Let be a real simple module. Let be a module with a simple head. If the following diagram

commutes up to a non-zero constant multiple, then is equal to the simple head of .

Proposition 3.2.16.

Let , , and be simple modules. We assume that is real and one of and is real.

(i)

If , then has a simple head and has a simple socle.

(ii)

If , then has a simple head and has a simple socle.

(iii)

If , then and have simple heads, and and have simple socles.

Proposition 3.2.17.

Let and be simple modules. Assume that one of them is real and . Then we have an exact sequence

In particular, has length .

Lemma 3.2.18.

Let and be simple modules. Assume that one of them is real. If there is an exact sequence

for self-dual simple modules , and integers , , then we have

Lemma 3.2.19.

Let and be simple modules. Assume that one of them is real. If the equation

holds in for self-dual simple modules , and integers , such that , then we have

(i)

,

(ii)

there exists an exact sequence

(iii)

is the socle of and is the head of .

Proposition 3.2.20.

Let , and be simple -modules. Assume that there is an exact sequence

and are simple, and are ungraded modules. Then is a real simple module.

Equation (3.2)
Corollary 3.2.21.

Let be simple -modules, and let be a real simple -module. If we have an exact sequence

and if and are simple, then is a real simple module.

Lemma 3.2.22.

Let and be a pair of commuting families of real simple modules. We assume that

(a)

is a commuting family of real simple modules,

(b)

commutes with for any .

Then we have

Theorem 4.1.1.

Let and be simple modules. We assume that is real. Then we have the equalities in the Grothendieck group as follows:

(i)

with simple modules such that ,

(ii)

with simple modules such that ,

(iii)

with simple modules such that ,

(iv)

with simple modules such that .

In particular, as well as appears only once in the Jordan–Hölder series of in .

Theorem 4.1.3.

Let be an element of the Grothendieck group given by

where is the self-dual simple module corresponding to and . Let be a real simple module in . Assume that we have an equality

in for some . Then commutes with and

for every such that .

Corollary 4.1.4.

Let and be simple modules. Assume that one of them is real. If and q-commute (i.e., for some ), then and commute. In particular, is simple.

Corollary 4.1.5.

Assume that the generalized Cartan matrix is symmetric and that satisfy the following conditions:

(i)

one of and is a member of the upper global basis up to a power of ,

(ii)

and q-commute.

Then their product is a member of the upper global basis of up to a power of .

Theorem 4.2.1 (Reference 34, Lemma 7.5).

Assume that the base field is of characteristic . Assume that has a head with a self-dual simple module and . Then we have the equality in the Grothendieck group

with self-dual simple modules and .

Equation (4.1)
Equation (5.1)
Equation (5.2)
Equation (5.3)
Equation (5.4)
Equation (5.5)
Equation (5.6)
Equation (6.1)
Definition 6.2.1.

We call a quadruple a quantum monoidal seed in if it satisfies the following conditions:

(i)

is an integer-valued -matrix whose principal part is skew-symmetric,

(ii)

is an integer-valued skew-symmetric -matrix,

(iii)

is a family of elements in ,

(iv)

is a family of simple objects such that for any ,

(v)

for all ,

(vi)

is simple for any sequence in ,

(vii)

The pair is compatible in the sense of Equation 5.3 with ,

(viii)

for all ,

(ix)

for all .

Lemma 6.2.2.

Set , the mutation of as in Equation 5.6. Set . Then we have

where

Definition 6.2.3.

We say that a quantum monoidal seed admits a mutation in direction if there exists a simple object such that

(i)

there exist exact sequences in ,

where and are as in Equation 6.3.

(ii)

is a quantum monoidal seed in .

We call the mutation of in direction .

Definition 7.1.1.

A pair is called admissible if

(i)

is a family of real simple self-dual objects of which commute with each other,

(ii)

is an integer-valued -matrix with a skew-symmetric principal part,

(iii)

for each , there exists a self-dual simple object of such that there is an exact sequence in

and commutes with for any .

Proposition 7.1.2.

Let be an admissible pair in , and let be as in Definition 7.1.1. Then we have the following properties:

(a)

The quadruple is a quantum monoidal seed in .

(b)

The self-dual simple object is real for every .

(c)

The quantum monoidal seed admits a mutation in each direction .

(d)

and are simply linked for any (i.e., ).

(e)

For any and , we have

Theorem 7.1.3.

Let be an admissible pair in and set

as in Proposition 7.1.2. We set . We assume further that

Then, for each , the pair is admissible in .

Equation (7.5)
Equation (7.6)
Equation (7.7)
Equations (7.8), (7.9)
Equation (7.10)
Equation (7.11)
Equation (7.12)
Equation (7.13)
Equation (7.14)
Corollary 7.1.4.

Let be an admissible pair in . Under the assumption Equation 7.4, is a monoidal categorification of the quantum cluster algebra . Furthermore, the following statements hold:

(i)

The quantum monoidal seed admits successive mutations in all directions.

(ii)

Any cluster monomial in is the isomorphism class of a real simple object in up to a power of .

(iii)

Any cluster monomial in is a Laurent polynomial of the initial cluster variables with a coefficient in .

Proposition 8.1.2 (Reference 17, Proposition 7.2.2).

We have an isomorphism of -bimodules

given by . Namely,

Proposition 8.1.3.

Let , , and . Then, is a member of the upper global basis of .

Proposition 8.1.4.
Equation (8.2)
Remark 8.2.3.

Note that the multiplication on given in Reference 11 is different from ours. Indeed, by denoting the product of and in Reference 11, Section 4.2 by , for , we have

where for . By Lemma 8.5.3 below, we have

for , where . In particular, we have a -algebra isomorphism from to given by

Note also that the bar-involution is a ring anti-isomorphism between and .

Equation (8.4)
Proposition 8.5.2.

For , set

. Then we have

Lemma 8.5.3.

For , if , then

Proposition 8.5.4.

For , we have

Proposition 8.6.2.

Let and . Then for any , vanishes or is a member of the lower global basis of .

Equation (8.6)
Lemma 9.1.1.

is a member of the upper global basis of . Moreover, is either a member of the upper global basis of or zero.

Corollary 9.1.3.

For and , we have

Lemma 9.1.5.

Let , such that and .

(i)

If , then

(ii)

If and , then .

(iii)

If , then

(iv)

If and , then .

Proposition 9.1.6 (Reference 3, (10.2)).

Let and such that and . Then we have

(i)

.

(ii)

If we assume further that and , then we have

or equivalently

Proposition 9.1.7.

For and , set and with . Then we have

Lemma 9.1.8.

For and such that , we have

Lemma 9.1.9.

For and , we have

Proposition 9.2.1 (Reference 11, Proposition 3.2).

Assume that the Kac–Moody algebra is of symmetric type. Assume that and satisfy and . Then

and

where .

Lemma 9.3.1.

is the largest upper crystal lattice of contained in the lower crystal lattice .

Equation (9.4)
Lemma 9.3.2.

Let and such that . Then we have

Theorem 9.3.3.

Let and such that . Then we have

Lemma 9.3.4.

Let and such that . Then we have

Equation (9.6)
Proposition 9.4.1.

Let , and set . Then we have

Equation (9.8)
Proposition 10.1.2.

Let with . Assume that an -module satisfies . Then the left -module homomorphism given by

extends uniquely to an -bilinear homomorphism

Equation (10.4)
Proposition 10.1.5.

Let , be modules and .

(i)

If and , then we have

(ii)

If and , then we have

Corollary 10.1.6.

Let , and let be a real simple module. Then is also real simple.

Proposition 10.1.8.

Let and be simple modules. We assume that one of them is real. If , then we have an isomorphism in

Similarly, if , then we have

Proposition 10.2.3.

Let , and such that , , , and . Then

(i)

and commute,

(ii)

,

(iii)

,.

Proposition 10.2.4.

Let , such that and .

(i)

If , then

(ii)

If and , then .

(iii)

If , then

(iv)

If and , then .

Proposition 10.2.5.

Assume that and satisfy and .

(i)

We have exact sequences

and

where .

(ii)

.

Theorem 10.3.1.

Let and such that . Then there exists a canonical epimorphism

which is equivalent to saying that .

In particular, we have

Proposition 10.3.2.

Let and .

(i)

If , , , and , then we have

(ii)

If , , , and , then we have

Equation (10.8)
Proposition 10.3.3.

Let such that and . Then we have

Equation (11.1)
Definition 11.2.1.

For , let be the smallest monoidal abelian full subcategory of satisfying the following properties:

(i)

is stable under the subquotients, extensions, and grading shifts,

(ii)

contains for all .

Theorem 11.2.2.

The pair is admissible.

Equation (11.3)
Lemma 11.2.5.

We have .

References

Reference [1]
A. A. Beĭlinson, J. Bernstein, and P. Deligne, Faisceaux pervers (French), Analysis and topology on singular spaces, I (Luminy, 1981), Astérisque, vol. 100, Soc. Math. France, Paris, 1982, pp. 5–171. MR751966,
Show rawAMSref \bib{BBDG}{article}{ author={Be\u \i linson, A. A.}, author={Bernstein, J.}, author={Deligne, P.}, title={Faisceaux pervers}, language={French}, conference={ title={Analysis and topology on singular spaces, I}, address={Luminy}, date={1981}, }, book={ series={Ast\'erisque}, volume={100}, publisher={Soc. Math. France, Paris}, }, date={1982}, pages={5\textendash 171}, review={\MR {751966}}, }
Reference [2]
Arkady Berenstein and Andrei Zelevinsky, String bases for quantum groups of type , I. M. Gel′fand Seminar, Adv. Soviet Math., vol. 16, Amer. Math. Soc., Providence, RI, 1993, pp. 51–89. MR1237826,
Show rawAMSref \bib{BZ93}{article}{ author={Berenstein, Arkady}, author={Zelevinsky, Andrei}, title={String bases for quantum groups of type $A_r$}, conference={ title={I. M. Gel\cprime fand Seminar}, }, book={ series={Adv. Soviet Math.}, volume={16}, publisher={Amer. Math. Soc., Providence, RI}, }, date={1993}, pages={51\textendash 89}, review={\MR {1237826}}, }
Reference [3]
Arkady Berenstein and Andrei Zelevinsky, Quantum cluster algebras, Adv. Math. 195 (2005), no. 2, 405–455. MR2146350,
Show rawAMSref \bib{BZ05}{article}{ author={Berenstein, Arkady}, author={Zelevinsky, Andrei}, title={Quantum cluster algebras}, journal={Adv. Math.}, volume={195}, date={2005}, number={2}, pages={405\textendash 455}, issn={0001-8708}, review={\MR {2146350}}, }
Reference [4]
Giovanni Cerulli Irelli, Bernhard Keller, Daniel Labardini-Fragoso, and Pierre-Guy Plamondon, Linear independence of cluster monomials for skew-symmetric cluster algebras, Compos. Math. 149 (2013), no. 10, 1753–1764. MR3123308,
Show rawAMSref \bib{CKLP12}{article}{ author={Cerulli Irelli, Giovanni}, author={Keller, Bernhard}, author={Labardini-Fragoso, Daniel}, author={Plamondon, Pierre-Guy}, title={Linear independence of cluster monomials for skew-symmetric cluster algebras}, journal={Compos. Math.}, volume={149}, date={2013}, number={10}, pages={1753\textendash 1764}, issn={0010-437X}, review={\MR {3123308}}, }
Reference [5]
Ben Davison, Davesh Maulik, Jörg Schürmann, and Balázs Szendrői, Purity for graded potentials and quantum cluster positivity, Compos. Math. 151 (2015), no. 10, 1913–1944. MR3414389,
Show rawAMSref \bib{DMSS}{article}{ author={Davison, Ben}, author={Maulik, Davesh}, author={Sch\"urmann, J\"org}, author={Szendr\H oi, Bal\'azs}, title={Purity for graded potentials and quantum cluster positivity}, journal={Compos. Math.}, volume={151}, date={2015}, number={10}, pages={1913\textendash 1944}, issn={0010-437X}, review={\MR {3414389}}, }
Reference [6]
Sergey Fomin and Andrei Zelevinsky, Cluster algebras. I. Foundations, J. Amer. Math. Soc. 15 (2002), no. 2, 497–529. MR1887642,
Show rawAMSref \bib{FZ02}{article}{ author={Fomin, Sergey}, author={Zelevinsky, Andrei}, title={Cluster algebras. I. Foundations}, journal={J. Amer. Math. Soc.}, volume={15}, date={2002}, number={2}, pages={497\textendash 529}, issn={0894-0347}, review={\MR {1887642}}, }
Reference [7]
Christof Geiss, Bernard Leclerc, and Jan Schröer, Semicanonical bases and preprojective algebras (English, with English and French summaries), Ann. Sci. Éc. Norm. Supér. (4) 38 (2005), no. 2, 193–253. MR2144987,
Show rawAMSref \bib{GLS05}{article}{ author={Geiss, Christof}, author={Leclerc, Bernard}, author={Schr\"oer, Jan}, title={Semicanonical bases and preprojective algebras}, language={English, with English and French summaries}, journal={Ann. Sci. \'Ec. Norm. Sup\'er. (4)}, volume={38}, date={2005}, number={2}, pages={193\textendash 253}, issn={0012-9593}, review={\MR {2144987}}, }
Reference [8]
Christof Geiß, Bernard Leclerc, and Jan Schröer, Kac–Moody groups and cluster algebras, Adv. Math. 228 (2011), no. 1, 329–433. MR2822235,
Show rawAMSref \bib{GLS11}{article}{ author={Gei\ss , Christof}, author={Leclerc, Bernard}, author={Schr\"oer, Jan}, title={Kac--Moody groups and cluster algebras}, journal={Adv. Math.}, volume={228}, date={2011}, number={1}, pages={329\textendash 433}, issn={0001-8708}, review={\MR {2822235}}, }
Reference [9]
Christof Geiß, Bernard Leclerc, and Jan Schröer, Cluster algebra structures and semicanonical bases for unipotent groups, arXiv:0703039v4 [math.RT].
Reference [10]
Christof Geiss, Bernard Leclerc, and Jan Schröer, Factorial cluster algebras, Doc. Math. 18 (2013), 249–274. MR3064982,
Show rawAMSref \bib{GLS13}{article}{ author={Geiss, Christof}, author={Leclerc, Bernard}, author={Schr\"oer, Jan}, title={Factorial cluster algebras}, journal={Doc. Math.}, volume={18}, date={2013}, pages={249\textendash 274}, issn={1431-0635}, review={\MR {3064982}}, }
Reference [11]
C. Geiß, B. Leclerc, and J. Schröer, Cluster structures on quantum coordinate rings, Selecta Math. (N.S.) 19 (2013), no. 2, 337–397. MR3090232,
Show rawAMSref \bib{GLS}{article}{ author={Gei\ss , C.}, author={Leclerc, B.}, author={Schr\"oer, J.}, title={Cluster structures on quantum coordinate rings}, journal={Selecta Math. (N.S.)}, volume={19}, date={2013}, number={2}, pages={337\textendash 397}, issn={1022-1824}, review={\MR {3090232}}, }
Reference [12]
David Hernandez and Bernard Leclerc, Cluster algebras and quantum affine algebras, Duke Math. J. 154 (2010), no. 2, 265–341. MR2682185,
Show rawAMSref \bib{HL10}{article}{ author={Hernandez, David}, author={Leclerc, Bernard}, title={Cluster algebras and quantum affine algebras}, journal={Duke Math. J.}, volume={154}, date={2010}, number={2}, pages={265\textendash 341}, issn={0012-7094}, review={\MR {2682185}}, }
Reference [13]
David Hernandez and Bernard Leclerc, Monoidal categorifications of cluster algebras of type and , Symmetries, integrable systems and representations, Springer Proc. Math. Stat., vol. 40, Springer, Heidelberg, 2013, pp. 175–193. MR3077685,
Show rawAMSref \bib{HL13}{article}{ author={Hernandez, David}, author={Leclerc, Bernard}, title={Monoidal categorifications of cluster algebras of type $A$ and $D$}, conference={ title={Symmetries, integrable systems and representations}, }, book={ series={Springer Proc. Math. Stat.}, volume={40}, publisher={Springer, Heidelberg}, }, date={2013}, pages={175\textendash 193}, review={\MR {3077685}}, }
Reference [14]
Seok-Jin Kang, Masaki Kashiwara, Myungho Kim, and Se-Jin Oh, Symmetric quiver Hecke algebras and -matrices of quantum affine algebras IV, Selecta Math. (N.S.) 22 (2016), no. 4, 1987–2015. MR3573951,
Show rawAMSref \bib{K^3}{article}{ author={Kang, Seok-Jin}, author={Kashiwara, Masaki}, author={Kim, Myungho}, author={Oh, Se-Jin}, title={Symmetric quiver Hecke algebras and $R$-matrices of quantum affine algebras IV}, journal={Selecta Math. (N.S.)}, volume={22}, date={2016}, number={4}, pages={1987\textendash 2015}, issn={1022-1824}, review={\MR {3573951}}, }
Reference [15]
Seok-Jin Kang, Masaki Kashiwara, Myungho Kim, and Se-jin Oh, Simplicity of heads and socles of tensor products, Compos. Math. 151 (2015), no. 2, 377–396. MR3314831,
Show rawAMSref \bib{KKKO14}{article}{ author={Kang, Seok-Jin}, author={Kashiwara, Masaki}, author={Kim, Myungho}, author={Oh, Se-jin}, title={Simplicity of heads and socles of tensor products}, journal={Compos. Math.}, volume={151}, date={2015}, number={2}, pages={377\textendash 396}, issn={0010-437X}, review={\MR {3314831}}, }
Reference [16]
M. Kashiwara, On crystal bases of the -analogue of universal enveloping algebras, Duke Math. J. 63 (1991), no. 2, 465–516. MR1115118,
Show rawAMSref \bib{Kash91}{article}{ author={Kashiwara, M.}, title={On crystal bases of the $Q$-analogue of universal enveloping algebras}, journal={Duke Math. J.}, volume={63}, date={1991}, number={2}, pages={465\textendash 516}, issn={0012-7094}, review={\MR {1115118}}, }
Reference [17]
Masaki Kashiwara, Global crystal bases of quantum groups, Duke Math. J. 69 (1993), no. 2, 455–485. MR1203234,
Show rawAMSref \bib{Kas93}{article}{ author={Kashiwara, Masaki}, title={Global crystal bases of quantum groups}, journal={Duke Math. J.}, volume={69}, date={1993}, number={2}, pages={455\textendash 485}, issn={0012-7094}, review={\MR {1203234}}, }
Reference [18]
Masaki Kashiwara, The crystal base and Littelmann’s refined Demazure character formula, Duke Math. J. 71 (1993), no. 3, 839–858. MR1240605,
Show rawAMSref \bib{Kas93a}{article}{ author={Kashiwara, Masaki}, title={The crystal base and Littelmann's refined Demazure character formula}, journal={Duke Math. J.}, volume={71}, date={1993}, number={3}, pages={839\textendash 858}, issn={0012-7094}, review={\MR {1240605}}, }
Reference [19]
Masaki Kashiwara, Crystal bases of modified quantized enveloping algebra, Duke Math. J. 73 (1994), no. 2, 383–413. MR1262212,
Show rawAMSref \bib{Kash94}{article}{ author={Kashiwara, Masaki}, title={Crystal bases of modified quantized enveloping algebra}, journal={Duke Math. J.}, volume={73}, date={1994}, number={2}, pages={383\textendash 413}, issn={0012-7094}, review={\MR {1262212}}, }
Reference [20]
Masaki Kashiwara, On crystal bases, Representations of groups (Banff, AB, 1994), CMS Conf. Proc., vol. 16, Amer. Math. Soc., Providence, RI, 1995, pp. 155–197. MR1357199,
Show rawAMSref \bib{Kas95}{article}{ author={Kashiwara, Masaki}, title={On crystal bases}, conference={ title={Representations of groups}, address={Banff, AB}, date={1994}, }, book={ series={CMS Conf. Proc.}, volume={16}, publisher={Amer. Math. Soc., Providence, RI}, }, date={1995}, pages={155\textendash 197}, review={\MR {1357199}}, }
Reference [21]
Mikhail Khovanov and Aaron D. Lauda, A diagrammatic approach to categorification of quantum groups. I, Represent. Theory 13 (2009), 309–347. MR2525917,
Show rawAMSref \bib{KL09}{article}{ author={Khovanov, Mikhail}, author={Lauda, Aaron D.}, title={A diagrammatic approach to categorification of quantum groups. I}, journal={Represent. Theory}, volume={13}, date={2009}, pages={309\textendash 347}, issn={1088-4165}, review={\MR {2525917}}, }
Reference [22]
Mikhail Khovanov and Aaron D. Lauda, A diagrammatic approach to categorification of quantum groups II, Trans. Amer. Math. Soc. 363 (2011), no. 5, 2685–2700. MR2763732,
Show rawAMSref \bib{KL11}{article}{ author={Khovanov, Mikhail}, author={Lauda, Aaron D.}, title={A diagrammatic approach to categorification of quantum groups II}, journal={Trans. Amer. Math. Soc.}, volume={363}, date={2011}, number={5}, pages={2685\textendash 2700}, issn={0002-9947}, review={\MR {2763732}}, }
Reference [23]
Yoshiyuki Kimura, Quantum unipotent subgroup and dual canonical basis, Kyoto J. Math. 52 (2012), no. 2, 277–331. MR2914878,
Show rawAMSref \bib{Kimu12}{article}{ author={Kimura, Yoshiyuki}, title={Quantum unipotent subgroup and dual canonical basis}, journal={Kyoto J. Math.}, volume={52}, date={2012}, number={2}, pages={277\textendash 331}, issn={2156-2261}, review={\MR {2914878}}, }
Reference [24]
Yoshiyuki Kimura and Fan Qin, Graded quiver varieties, quantum cluster algebras and dual canonical basis, Adv. Math. 262 (2014), 261–312. MR3228430,
Show rawAMSref \bib{KQ14}{article}{ author={Kimura, Yoshiyuki}, author={Qin, Fan}, title={Graded quiver varieties, quantum cluster algebras and dual canonical basis}, journal={Adv. Math.}, volume={262}, date={2014}, pages={261\textendash 312}, issn={0001-8708}, review={\MR {3228430}}, }
Reference [25]
Atsuo Kuniba, Tomoki Nakanishi, and Junji Suzuki, -systems and -systems in integrable systems, J. Phys. A 44 (2011), no. 10, 103001, 146. MR2773889,
Show rawAMSref \bib{KNS11}{article}{ author={Kuniba, Atsuo}, author={Nakanishi, Tomoki}, author={Suzuki, Junji}, title={$T$-systems and $Y$-systems in integrable systems}, journal={J. Phys. A}, volume={44}, date={2011}, number={10}, pages={103001, 146}, issn={1751-8113}, review={\MR {2773889}}, }
Reference [26]
Philipp Lampe, A quantum cluster algebra of Kronecker type and the dual canonical basis, Int. Math. Res. Not. IMRN 13 (2011), 2970–3005. MR2817684,
Show rawAMSref \bib{Lampe11}{article}{ author={Lampe, Philipp}, title={A quantum cluster algebra of Kronecker type and the dual canonical basis}, journal={Int. Math. Res. Not. IMRN}, date={2011}, number={13}, pages={2970\textendash 3005}, issn={1073-7928}, review={\MR {2817684}}, }
Reference [27]
P. Lampe, Quantum cluster algebras of type and the dual canonical basis, Proc. Lond. Math. Soc. (3) 108 (2014), no. 1, 1–43. MR3162819,
Show rawAMSref \bib{Lampe14}{article}{ author={Lampe, P.}, title={Quantum cluster algebras of type $A$ and the dual canonical basis}, journal={Proc. Lond. Math. Soc. (3)}, volume={108}, date={2014}, number={1}, pages={1\textendash 43}, issn={0024-6115}, review={\MR {3162819}}, }
Reference [28]
Aaron D. Lauda and Monica Vazirani, Crystals from categorified quantum groups, Adv. Math. 228 (2011), no. 2, 803–861. MR2822211,
Show rawAMSref \bib{LV11}{article}{ author={Lauda, Aaron D.}, author={Vazirani, Monica}, title={Crystals from categorified quantum groups}, journal={Adv. Math.}, volume={228}, date={2011}, number={2}, pages={803\textendash 861}, issn={0001-8708}, review={\MR {2822211}}, }
Reference [29]
B. Leclerc, Imaginary vectors in the dual canonical basis of , Transform. Groups 8 (2003), no. 1, 95–104. MR1959765,
Show rawAMSref \bib{L03}{article}{ author={Leclerc, B.}, title={Imaginary vectors in the dual canonical basis of $U_q(\mathfrak {n})$}, journal={Transform. Groups}, volume={8}, date={2003}, number={1}, pages={95\textendash 104}, issn={1083-4362}, review={\MR {1959765}}, }
Reference [30]
Kyungyong Lee and Ralf Schiffler, Positivity for cluster algebras, Ann. of Math. (2) 182 (2015), no. 1, 73–125. MR3374957,
Show rawAMSref \bib{LS13}{article}{ author={Lee, Kyungyong}, author={Schiffler, Ralf}, title={Positivity for cluster algebras}, journal={Ann. of Math. (2)}, volume={182}, date={2015}, number={1}, pages={73\textendash 125}, issn={0003-486X}, review={\MR {3374957}}, }
Reference [31]
G. Lusztig, Canonical bases arising from quantized enveloping algebras, J. Amer. Math. Soc. 3 (1990), no. 2, 447–498. MR1035415,
Show rawAMSref \bib{Lusz90}{article}{ author={Lusztig, G.}, title={Canonical bases arising from quantized enveloping algebras}, journal={J. Amer. Math. Soc.}, volume={3}, date={1990}, number={2}, pages={447\textendash 498}, issn={0894-0347}, review={\MR {1035415}}, }
Reference [32]
G. Lusztig, Canonical bases in tensor products, Proc. Natl. Acad. Sci. USA 89 (1992), no. 17, 8177–8179. MR1180036,
Show rawAMSref \bib{Lusz92}{article}{ author={Lusztig, G.}, title={Canonical bases in tensor products}, journal={Proc. Natl. Acad. Sci. USA}, volume={89}, date={1992}, number={17}, pages={8177\textendash 8179}, issn={0027-8424}, review={\MR {1180036}}, }
Reference [33]
George Lusztig, Introduction to quantum groups, Progress in Mathematics, vol. 110, Birkhäuser Boston, Inc., Boston, MA, 1993. MR1227098,
Show rawAMSref \bib{Lus93}{book}{ author={Lusztig, George}, title={Introduction to quantum groups}, series={Progress in Mathematics}, volume={110}, publisher={Birkh\"auser Boston, Inc., Boston, MA}, date={1993}, pages={xii+341}, isbn={0-8176-3712-5}, review={\MR {1227098}}, }
Reference [34]
Peter J. McNamara, Representations of Khovanov–Lauda–Rouquier algebras III: symmetric affine type, Math. Z. 287 (2017), no. 1-2, 243–286. MR3694676,
Show rawAMSref \bib{Mc14}{article}{ author={McNamara, Peter J.}, title={Representations of Khovanov--Lauda--Rouquier algebras III: symmetric affine type}, journal={Math. Z.}, volume={287}, date={2017}, number={1-2}, pages={243\textendash 286}, issn={0025-5874}, review={\MR {3694676}}, }
Reference [35]
Hiraku Nakajima, Cluster algebras and singular supports of perverse sheaves, Advances in representation theory of algebras, EMS Ser. Congr. Rep., Eur. Math. Soc., Zürich, 2013, pp. 211–230. MR3220538,
Show rawAMSref \bib{Nak13}{article}{ author={Nakajima, Hiraku}, title={Cluster algebras and singular supports of perverse sheaves}, conference={ title={Advances in representation theory of algebras}, }, book={ series={EMS Ser. Congr. Rep.}, publisher={Eur. Math. Soc., Z\"urich}, }, date={2013}, pages={211\textendash 230}, review={\MR {3220538}}, }
Reference [36]
Hiraku Nakajima, Quiver varieties and cluster algebras, Kyoto J. Math. 51 (2011), no. 1, 71–126. MR2784748,
Show rawAMSref \bib{Nak11}{article}{ author={Nakajima, Hiraku}, title={Quiver varieties and cluster algebras}, journal={Kyoto J. Math.}, volume={51}, date={2011}, number={1}, pages={71\textendash 126}, issn={2156-2261}, review={\MR {2784748}}, }
Reference [37]
Fan Qin, Triangular bases in quantum cluster algebras and monoidal categorification conjectures, Duke Math. J. 166 (2017), no. 12, 2337–2442. MR3694569,
Show rawAMSref \bib{Qin15}{article}{ author={Qin, Fan}, title={Triangular bases in quantum cluster algebras and monoidal categorification conjectures}, journal={Duke Math. J.}, volume={166}, date={2017}, number={12}, pages={2337\textendash 2442}, issn={0012-7094}, review={\MR {3694569}}, }
Reference [38]
R. Rouquier, 2-Kac–Moody algebras, arXiv:0812.5023v1.
Reference [39]
Raphaël Rouquier, Quiver Hecke algebras and 2-Lie algebras, Algebra Colloq. 19 (2012), no. 2, 359–410. MR2908731,
Show rawAMSref \bib{R11}{article}{ author={Rouquier, Rapha\"el}, title={Quiver Hecke algebras and 2-Lie algebras}, journal={Algebra Colloq.}, volume={19}, date={2012}, number={2}, pages={359\textendash 410}, issn={1005-3867}, review={\MR {2908731}}, }
Reference [40]
M. Varagnolo and E. Vasserot, Canonical bases and KLR-algebras, J. Reine Angew. Math. 659 (2011), 67–100. MR2837011,
Show rawAMSref \bib{VV09}{article}{ author={Varagnolo, M.}, author={Vasserot, E.}, title={Canonical bases and KLR-algebras}, journal={J. Reine Angew. Math.}, volume={659}, date={2011}, pages={67\textendash 100}, issn={0075-4102}, review={\MR {2837011}}, }

Article Information

MSC 2010
Primary: 13F60 (Cluster algebras), 81R50 (Quantum groups and related algebraic methods), 16Gxx (Representation theory of rings and algebras), 17B37 (Quantum groups and related deformations)
Keywords
  • Cluster algebra
  • quantum cluster algebra
  • monoidal categorification
  • Khovanov–Lauda–Rouquier algebra
  • unipotent quantum coordinate ring
  • quantum affine algebra
Author Information
Seok-Jin Kang
Research Institute of Computers, Information and Communication, Pusan National University, 2, Busandaehak-ro Pusan 46241, Korea
soccerkang@hotmail.com
MathSciNet
Masaki Kashiwara
Research Institute for Mathematical Sciences, Kyoto University, Kyoto 606-8502, Japan
masaki@kurims.kyoto-u.ac.jp
MathSciNet
Myungho Kim
Department of Mathematics, Kyung Hee University, Seoul 02447, Korea
mkim@khu.ac.kr
MathSciNet
Se-jin Oh
Department of Mathematics Ewha Womans University, Seoul 03760, Korea
sejin092@gmail.com
MathSciNet
Additional Notes

This work was supported by Grant-in-Aid for Scientific Research (B) 22340005, Japan Society for the Promotion of Science.

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIP) (No. NRF-2017R1C1B2007824).

This work was supported by NRF Grant # 2016R1C1B2013135.

This research was supported by Ministry of Culture, Sports and Tourism (MCST) and Korea Creative Content Agency (KOCCA) in the Culture Technology (CT) Research & Development Program 2017.

Journal Information
Journal of the American Mathematical Society, Volume 31, Issue 2, ISSN 1088-6834, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , , and published on .
Copyright Information
Copyright 2017 American Mathematical Society
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/jams/895
  • MathSciNet Review: 3758148
  • Show rawAMSref \bib{3758148}{article}{ author={Kang, Seok-Jin}, author={Kashiwara, Masaki}, author={Kim, Myungho}, author={Oh, Se-jin}, title={Monoidal categorification of cluster algebras}, journal={J. Amer. Math. Soc.}, volume={31}, number={2}, date={2018-04}, pages={349-426}, issn={0894-0347}, review={3758148}, doi={10.1090/jams/895}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.