Cluster algebras I: Foundations

By Sergey Fomin and Andrei Zelevinsky

To the memory of Sergei Kerov

Abstract

In an attempt to create an algebraic framework for dual canonical bases and total positivity in semisimple groups, we initiate the study of a new class of commutative algebras.

1. Introduction

In this paper, we initiate the study of a new class of algebras, which we call cluster algebras. Before giving precise definitions, we present some of the main features of these algebras. For any positive integer , a cluster algebra of rank is a commutative ring with unit and no zero divisors, equipped with a distinguished family of generators called cluster variables. The set of cluster variables is the (non-disjoint) union of a distinguished collection of -subsets called clusters. These clusters have the following exchange property: for any cluster and any element , there is another cluster obtained from by replacing with an element related to by a binomial exchange relation

where and are two monomials without common divisors in the variables . Furthermore, any two clusters can be obtained from each other by a sequence of exchanges of this kind.

The prototypical example of a cluster algebra of rank 1 is the coordinate ring of the group , viewed in the following way. Writing a generic element of as we consider the entries and as cluster variables, and the entries and as scalars. There are just two clusters and , and is the algebra over the polynomial ring generated by the cluster variables and subject to the binomial exchange relation

Another important incarnation of a cluster algebra of rank 1 is the coordinate ring of the base affine space of the special linear group ; here is the maximal unipotent subgroup of consisting of all unipotent upper triangular matrices. Using the standard notation for the Plücker coordinates on , we view and as cluster variables; then is the algebra over the polynomial ring generated by the two cluster variables and subject to the binomial exchange relation

This form of representing the algebra is closely related to the choice of a linear basis in it consisting of all monomials in the six Plücker coordinates which are not divisible by . This basis was introduced and studied in Reference 9 under the name “canonical basis”. As a representation of , the space is the multiplicity-free direct sum of all irreducible finite-dimensional representations, and each of the components is spanned by a part of the above basis. Thus, this construction provides a “canonical” basis in every irreducible finite-dimensional representation of . After Lusztig’s work Reference 13, this basis had been recognized as (the classical limit at of) the dual canonical basis, i.e., the basis in the -deformed algebra which is dual to Lusztig’s canonical basis in the appropriate -deformed universal enveloping algebra (a.k.a. quantum group). The dual canonical basis in the space was later constructed explicitly for a few other classical groups of small rank: for in Reference 16 and for in Reference 2. In both cases, can be seen to be a cluster algebra: there are 6 clusters of size 2 for , and 14 clusters of size 3 for .

We conjecture that the above examples can be extensively generalized: for any simply-connected connected semisimple group , the coordinate rings and , as well as coordinate rings of many other interesting varieties related to , have a natural structure of a cluster algebra. This structure should serve as an algebraic framework for the study of “dual canonical bases” in these coordinate rings and their -deformations. In particular, we conjecture that all monomials in the variables of any given cluster (the cluster monomials) belong to this dual canonical basis.

A particularly nice and well-understood example of a cluster algebra of an arbitrary rank is the homogeneous coordinate ring of the Grassmannian of -dimensional subspaces in . This ring is generated by the Plücker coordinates , for , subject to the relations

for all . It is convenient to identify the indices with the vertices of a convex -gon, and the Plücker coordinates with its sides and diagonals. We view the sides as scalars, and the diagonals as cluster variables. The clusters are the maximal families of pairwise noncrossing diagonals; thus, they are in a natural bijection with the triangulations of this polygon. It is known that the cluster monomials form a linear basis in . To be more specific, we note that this ring is naturally identified with the ring of polynomial -invariants of an -tuple of points in . Under this isomorphism, the basis of cluster monomials corresponds to the basis considered in Reference 11Reference 18. (We are grateful to Bernd Sturmfels for bringing these references to our attention.)

An essential feature of the exchange relations (Equation 1.1) is that the right-hand side does not involve subtraction. Recursively applying these relations, one can represent any cluster variable as a subtraction-free rational expression in the variables of any given cluster. This positivity property is consistent with a remarkable connection between canonical bases and the theory of total positivity, discovered by G. Lusztig Reference 14Reference 15. Generalizing the classical concept of totally positive matrices, he defined totally positive elements in any reductive group , and proved that all elements of the dual canonical basis in take positive values at them.

It was realized in Reference 15Reference 5 that the natural geometric framework for total positivity is given by double Bruhat cells, the intersections of cells of the Bruhat decompositions with respect to two opposite Borel subgroups. Different aspects of total positivity in double Bruhat cells were explored by the authors of the present paper and their collaborators in Reference 1Reference 3Reference 4Reference 5Reference 6Reference 7Reference 12Reference 17Reference 20. The binomial exchange relations of the form (Equation 1.1) played a crucial role in these studies. It was the desire to explain the ubiquity of these relations and to place them in a proper context that led us to the concept of cluster algebras. The crucial step in this direction was made in Reference 20, where a family of clusters and exchange relations was explicitly constructed in the coordinate ring of an arbitrary double Bruhat cell. However, this family was not complete: in general, some clusters were missing, and not any member of a cluster could be exchanged from it. Thus, we started looking for a natural way to “propagate” exchange relations from one cluster to another. The concept of cluster algebras is the result of this investigation. We conjecture that the coordinate ring of any double Bruhat cell is a cluster algebra.

This article, in which we develop the foundations of the theory, is conceived as the first in a forthcoming series. We attempt to make the exposition elementary and self-contained; in particular, no knowledge of semisimple groups, quantum groups or total positivity is assumed on the part of the reader.

One of the main structural features of cluster algebras established in the present paper is the following Laurent phenomenon: any cluster variable viewed as a rational function in the variables of any given cluster is in fact a Laurent polynomial. This property is quite surprising: in most cases, the numerators of these Laurent polynomials contain a huge number of monomials, and the numerator for moves into the denominator when we compute the cluster variable obtained from by an exchange (Equation 1.1). The magic of the Laurent phenomenon is that, at every stage of this recursive process, a cancellation will inevitably occur, leaving a single monomial in the denominator.

In view of the positivity property discussed above, it is natural to expect that all Laurent polynomials for cluster variables will have positive coefficients. This seems to be a rather deep property; our present methods do not provide a proof of it.

On the bright side, it is possible to establish the Laurent phenomenon in many different situations spreading beyond the cluster algebra framework. One such extension is given in Theorem 3.2. By a modification of the method developed here, a large number of additional interesting instances of the Laurent phenomenon are established in a separate paper Reference 8.

The paper is organized as follows. Section 2 contains an axiomatic definition, first examples and the first structural properties of cluster algebras. One of the technical difficulties in setting up the foundations involves the concept of an exchange graph whose vertices correspond to clusters, and the edges to exchanges among them. It is convenient to begin by taking the -regular tree as our underlying graph. This tree can be viewed as a universal cover for the actual exchange graph, whose appearance is postponed until Section 7.

The Laurent phenomenon is established in Section 3. In Sections 4 and 5, we scrutinize the main definition, obtain useful reformulations, and introduce some important classes of cluster algebras.

Section 6 contains a detailed analysis of cluster algebras of rank 2. This analysis exhibits deep and somewhat mysterious connections between cluster algebras and Kac-Moody algebras. This is just the tip of an iceberg: these connections will be further explored (for cluster algebras of an arbitrary rank) in the sequel to this paper. The main result of this sequel is a complete classification of cluster algebras of finite type, i.e., those with finitely many distinct clusters; cf. Example 7.6. This classification turns out to be yet another instance of the famous Cartan-Killing classification.

2. Main definitions

Let be a finite set of size ; the standard choice will be . Let denote the -regular tree, whose edges are labeled by the elements of , so that the edges emanating from each vertex receive different labels. By a common abuse of notation, we will sometimes denote by the set of the tree’s vertices. We will write if vertices are joined by an edge labeled by .

To each vertex , we will associate a cluster of generators (“variables”) . All these variables will commute with each other and satisfy the following exchange relations, for every edge in :

Here and are two monomials in the variables ; we think of these monomials as being associated with the two ends of the edge .

To be more precise, let be an abelian group without torsion, written multiplicatively. We call the coefficient group; a prototypical example is a free abelian group of finite rank. Every monomial in (Equation 2.2) will have the form

for some coefficient and some nonnegative integer exponents .

The monomials must satisfy certain conditions (axioms). To state them, we will need a little preparation. Let us write to denote that a polynomial divides a polynomial . Accordingly, means that the monomial contains the variable . For a rational function , the notation will denote the result of substituting for into . To illustrate, if , then .

Definition 2.1.

An exchange pattern on with coefficients in is a family of monomials of the form (Equation 2.3) satisfying the following four axioms:

We note that in the last axiom, the substitution is effectively monomial, since in the event that neither nor contain , condition (Equation 2.6) requires that both and do not depend on , thus making the whole substitution irrelevant.

One easily checks that axiom (Equation 2.7) is invariant under the “flip” , , so no restrictions are added if we apply it “backwards”. The axioms also imply at once that setting

for every edge , we obtain another exchange pattern ; this gives a natural involution on the set of all exchange patterns.

Remark 2.2.

Informally speaking, axiom (Equation 2.7) describes the propagation of an exchange pattern along the edges of . More precisely, let us fix the exchange monomials for all edges emanating from a given vertex . This choice uniquely determines the ratio for any vertex adjacent to and any edge (to see this, take and in (Equation 2.7), and allow to vary). In view of (Equation 2.5), this ratio in turn uniquely determines the exponents of all variables in both monomials and . There remains, however, one degree of freedom in determining the coefficients and because only their ratio is prescribed by (Equation 2.7). In Section 5 we shall introduce an important class of normalized exchange patterns for which this degree of freedom disappears, and so the whole pattern is uniquely determined by the monomials associated with edges emanating from a given vertex.

Let denote the group ring of with integer coefficients. For an edge , we refer to the binomial as the exchange polynomial. We will write or to indicate this fact. Note that, in view of the axiom (Equation 2.4), the right-hand side of the exchange relation (Equation 2.2) can be written as , which is the same as .

Let be an exchange pattern on with coefficients in . Note that since is torsion-free, the ring has no zero divisors. For every vertex , let denote the field of rational functions in the cluster variables , , with coefficients in . For every edge , we define a -linear field isomorphism by

Note that property (Equation 2.4) ensures that . The transition maps enable us to identify all the fields with each other. We can then view them as a single field that contains all the elements , for all and . Inside , these elements satisfy the exchange relations (Equation 2.1)–(Equation 2.2).

Definition 2.3.

Let be a subring with unit in containing all coefficients for and . The cluster algebra of rank over associated with an exchange pattern is the -subalgebra with unit in generated by the union of all clusters , for .

The smallest possible ground ring is the subring of generated by all the coefficients ; the largest one is itself. An intermediate choice of appears in Proposition 2.6 below.

Since is a subring of a field , it is a commutative ring with no zero divisors. We also note that if is obtained from by the involution (Equation 2.8), then the cluster algebra is naturally identified with .

Example 2.4.

Let . The tree has only one edge . The corresponding cluster algebra has two generators and satisfying the exchange relation

where and are arbitrary elements of the coefficient group . In the “universal” setting, we take to be the free abelian group generated by and . Then the two natural choices for the ground ring are the polynomial ring , and the Laurent polynomial ring . All other realizations of can be viewed as specializations of the universal one. Despite the seeming triviality of this example, it covers several important algebras: the coordinate ring of each of the varieties , and (cf. Section 1) is a cluster algebra of rank , for an appropriate choice of , , and .

Example 2.5.

Consider the case . The tree is shown below:

Let us denote the cluster variables as follows:

(the above equalities among the cluster variables follow from (Equation 2.1)). Then the clusters look like

We claim that the exchange relations (Equation 2.2) can be written in the following form:

where the integers and are either both positive or both equal to , and the coefficients and are elements of satisfying the relations

Furthermore, any such choice of parameters , , , results in a well-defined cluster algebra of rank 2.

To prove this, we notice that, in view of (Equation 2.4)–(Equation 2.5), both monomials and do not contain the variable , and at most one of them contains . If enters neither nor , then these two are simply elements of . But then (Equation 2.6) forces all monomials  to be elements of , while (Equation 2.7) implies that it is possible to give the names and to the two monomials corresponding to each edge so that (Equation 2.11)–(Equation 2.12) hold with .

Next, consider the case when precisely one of the monomials and contains . Applying if necessary the involution (Equation 2.8) to our exchange pattern, we may assume that and for some positive integer and some . Thus, the exchange relation associated to the edge takes the form . By (Equation 2.6), we have and for some positive integer and some . Then the exchange relation for the edge takes the form . At this point, we invoke (Equation 2.7):

By (Equation 2.5), we have and for some satisfying . Continuing in the same way, we obtain all relations (Equation 2.11)–(Equation 2.12).

For fixed and , the “universal” coefficient group is the multiplicative abelian group generated by the elements and for all subject to the defining relations (Equation 2.12). It is easy to see that this is a free abelian group of infinite rank. As a set of its free generators, one can choose any subset of that contains four generators and precisely one generator from each pair for .

A nice specialization of this setup is provided by the homogeneous coordinate ring of the Grassmannian . Recall (cf. Section 1) that this ring is generated by the Plücker coordinates , where and are distinct elements of the cyclic group . We shall write for , and adopt the convention ; see Figure 1.

The ideal of relations among the Plücker coordinates is generated by the relations

for . A direct check shows that these relations are a specialization of the relations (Equation 2.11), if we set , , , and for all . The coefficient group is the multiplicative free abelian group with generators . It is also immediate that the elements and defined in this way satisfy the relations (Equation 2.12).

We conclude this section by introducing two important operations on exchange patterns: restriction and direct product. Let us start with restriction. Let be an exchange pattern of rank with an index set and coefficient group . Let be a subset of size in . Let us remove from all edges labeled by indices in , and choose any connected component of the resulting graph. This component is naturally identified with . Let denote the restriction of to , i.e., the collection of monomials for all and . Then is an exchange pattern on whose coefficient group is the direct product of with the multiplicative free abelian group with generators , . We shall say that is obtained from by restriction from to . Note that depends on the choice of a connected component , so there can be several different patterns obtained from by restriction from to . (We thank the anonymous referee for pointing this out.)

Proposition 2.6.

Let be a cluster algebra of rank associated with an exchange pattern , and let be obtained from by restriction from to using a connected component . The -subalgebra of generated by is naturally identified with the cluster algebra , where is the polynomial ring .

Proof.

If , then (Equation 2.1) implies that stays constant as varies over . Therefore, we can identify this variable with the corresponding generator of the coefficient group , and the statement follows.

Let us now consider two exchange patterns and of ranks and , respectively, with index sets and , and coefficient groups and . We will construct the exchange pattern (the direct product of and ) of rank , with the index set , and coefficient group . Consider the tree whose edges are colored by , and, for , let be a map with the following property: if in and (resp., ), then in (resp., ). Clearly, such a map exists and is essentially unique: it is determined by specifying the image of any vertex of . We now introduce the exchange pattern on by setting, for every and , the monomial to be equal to , the latter monomial coming from the exchange pattern . The axioms (Equation 2.4)–(Equation 2.7) for are checked directly.

Proposition 2.7.

Let and be cluster algebras. Let and . Then the cluster algebra is canonically isomorphic to the tensor product of algebras all tensor products are taken over .

Proof.

Let us identify each cluster variable , for and (resp., ), with (resp., . Under this identification, the exchange relations for the exchange pattern become identical to the exchange relations for and .

3. The Laurent phenomenon

In this section we prove the following important property of cluster algebras.

Theorem 3.1.

In a cluster algebra, any cluster variable is expressed in terms of any given cluster as a Laurent polynomial with coefficients in the group ring .

We conjecture that each of the coefficients in these Laurent polynomials is actually a nonnegative integer linear combination of elements in .

We will obtain Theorem 3.1 as a corollary of a more general result, which applies to more general underlying graphs and more general (not necessarily binomial) exchange polynomials.

Since Theorem 3.1 is trivial for , we shall assume that . For every , let be a tree of the form shown in Figure 2. The tree has vertices of degree in its “spine” and vertices of degree 1. We label every edge of the tree by an element of an -element index set , so that the edges incident to each vertex on the spine receive different labels. (The reader may wish to think of the tree as being part of the -regular tree of the cluster-algebra setup.)

We fix two vertices and of that do not belong to the spine and are connected to its opposite ends. This gives rise to the orientation on the spine: away from and towards (see Figure 2).

As before, let be an abelian group without torsion, written multiplicatively. Let denote the additive semigroup generated by in the integer group ring . Assume that a nonzero polynomial in the variables , with coefficients in , is associated with every edge of . We call an exchange polynomial, and write to describe this situation. Suppose that the exchange polynomials associated with the edges of satisfy the following conditions:

(Note the orientation of the edge in (Equation 3.2).)

For every vertex on the spine, let denote the family of exchange polynomials associated with the edges emanating from . Also, let denote the collection of all Laurent polynomials that appear in condition (Equation 3.2), for all possible choices of , and let denote the subring with unit generated by all coefficients of the Laurent polynomials from , where is the vertex on the spine connected with .

As before, we associate a cluster to each vertex , and consider the field of rational functions in these variables with coefficients in . All these fields are identified with each other by the transition isomorphisms defined as in (Equation 2.9). We then view the fields as a single field that contains all the elements , for and . These elements satisfy the exchange relations (Equation 2.1) and the following version of (Equation 2.2):

for any edge in .

Theorem 3.2.

If conditions (Equation 3.1)–(Equation 3.2) are satisfied, then each element of the cluster is a Laurent polynomial in the cluster , with coefficients in the ring .

We note that Theorem 3.2 is indeed a generalization of Theorem 3.1, for the following reasons:

is naturally embedded into ;

conditions (Equation 3.1)–(Equation 3.2) are less restrictive than (Equation 2.4)–(Equation 2.7);

the claim being made in Theorem 3.2 about coefficients of the Laurent polynomials is stronger than that of Theorem 3.1, since .

Proof.

We start with some preparations. We shall write any Laurent polynomial in the variables in the form

where all coefficients are nonzero, is a finite subset of the lattice (i.e., the lattice of rank with coordinates labeled by ), and is the usual shorthand for . The set is called the support of and denoted by .

Notice that once we fix the collection , condition (Equation 3.2) can be used as a recursive rule for computing from , for any edge on the spine. It follows that the whole pattern of exchange polynomials is determined by the families of polynomials and . Moreover, since these polynomials have coefficients in , and the expression for in (Equation 3.2) does not involve subtraction, it follows that the support of any exchange polynomial is uniquely determined by the supports of the polynomials from and . Note that condition (Equation 3.1) can be formulated as a set of restrictions on these supports. In particular, it requires that in the situation of (Equation 3.2), the Laurent polynomial does not depend on and is a polynomial in ; in other words, every should have and .

We now fix a family of supports , for all , and assume that this family complies with (Equation 3.1). As is common in algebra, we shall view the coefficients , for all and , as indeterminates. Then all the coefficients in all exchange polynomials become “canonical” (i.e., independent of the choice of ) polynomials in these indeterminates, with positive integer coefficients.

The above discussion shows that it suffices to prove our theorem in the following “universal coefficients” setup: let be the free abelian group (written multiplicatively) with generators , for all and . Under this assumption, is simply the integer polynomial ring in the indeterminates .

Recall that we can view all cluster variables as elements of the field of rational functions in the cluster with coefficients in . For , let denote the ring of Laurent polynomials in the cluster , with coefficients in . We view each as a subring of the ambient field .

In this terminology, our goal is to show that the cluster is contained in . We proceed by induction on , the size of the spine. The claim is trivial for , so let us assume that , and furthermore assume that our statement is true for all “caterpillars” with smaller spine.

Let us abbreviate and , and suppose that the path from to starts with the following two edges: . Let be the vertex such that .

The following lemma plays a crucial role in our proof.

Lemma 3.3.

The clusters , , and are contained in . Furthermore, as elements of .

Note that is a unique factorization domain, so any two elements have a well-defined greatest common divisor , which is an element of defined up to a multiple from the group of units (that is, invertible elements) of . In our “universal” situation, consists of Laurent monomials in the cluster with coefficients .

Proof.

The only element from the clusters , , and whose inclusion in is not immediately obvious is . To simplify the notation, let us denote , , , , and , so that these variables appear in the clusters at , as shown below:

Note that the variables , for , do not change as we move among the four clusters under consideration. The lemma is then restated as saying that

Another notational convention will be based on the fact that each of the polynomials has a distinguished variable on which it depends, namely for and , and for . (In view of (Equation 3.1), and do not depend on , while does not depend on .) With this in mind, we will routinely write , , and as polynomials in one (distinguished) variable. In the same spirit, the notation , , etc., will refer to the partial derivative with respect to the distinguished variable.

We will prove the statements (Equation 3.3), (Equation 3.4), and (Equation 3.5) one by one, in this order.

By (Equation 3.2), the polynomial is given by

where is an “honest” polynomial in and a Laurent polynomial in the “mute” variables , . (Recall that does not depend on .) We then have:

Since

and

(Equation 3.3) follows.

We next prove (Equation 3.4). We have

Since and are invertible in , we conclude that . Now the trouble that we took in passing to universal coefficients finally pays off: since and are nonzero polynomials in the cluster whose coefficients are distinct generators of the polynomial ring , it follows that , proving (Equation 3.4).

It remains to prove (Equation 3.5). Let

Then

Our goal is to show that ; to this end, we are going to compute as “explicitly” as possible. We have, ,

Hence

Note that the right-hand side is a linear polynomial in , whose coefficients are Laurent polynomials in the rest of the variables of the cluster . Thus our claim will follow if we show that . This, again, is a consequence of our “universal coefficients” setup since the coefficients of , and are distinct generators of the polynomial ring .

We can now complete the proof of Theorem 3.2. We need to show that any variable belongs to . Since both and are closer to than , we can use the inductive assumption to conclude that belongs to both and . Since , it follows from (Equation 2.1) that can be written as for some and . On the other hand, since , it follows from (Equation 2.1) and from the inclusion guaranteed by Lemma 3.3 that has the form for some and some . The inclusion now follows from the fact that, by the last statement in Lemma 3.3, the denominators in the two obtained expressions for are coprime in .

Several examples that can be viewed as applications of Theorem 3.2 are given in Reference 8.

4. Exchange relations: The exponents

Let be an exchange pattern (see Definition 2.1). In this section we will ignore the coefficients in the monomials and take a closer look at the dynamics of their exponents. (An alternative point of view that the reader may find helpful is to assume that all exchange patterns considered in this section will have all their coefficients equal to .) For every edge in , let us write the ratio of the corresponding monomials as

where (cf. (Equation 2.3)); we note that ratios of this kind have already appeared in (Equation 2.7). Let us denote by the integer matrix whose entries are the exponents in (Equation 4.1). In view of (Equation 2.5), the exponents in and are recovered from :

Thus, the family of matrices encodes all the exponents in all monomials of an exchange pattern.

We shall describe the conditions on the family of matrices imposed by the axioms of an exchange pattern. To do this, we need some preparation.

Definition 4.1.

A square integer matrix is called sign-skew-symmetric if, for any and , either , or else and are of opposite sign; in particular, for all .

Definition 4.2.

Let and be square integer matrices of the same size. We say that is obtained from by the matrix mutation in direction  and write if

An immediate check shows that is involutive, i.e., its square is the identity transformation.

Proposition 4.3.

A family of integer matrices corresponds to an exchange pattern if and only if the following conditions hold:

(1) is sign-skew-symmetric for any .

(2) If , then .

Proof.

We start with the “only if” part, i.e., we assume that the matrices are determined by an exchange pattern via (Equation 4.1) and check the conditions (1)–(2). The condition follows from (Equation 2.4). The remaining part of (1) (dealing with ), follows at once from (Equation 2.6). Turning to part (2), the equality is immediate from the definition (Equation 4.1). Now suppose that . In this case, we apply the axiom (Equation 2.7) to the edge taken together with the two adjacent edges emanating from and and labeled by . Taking (Equation 4.2) into account, we obtain:

where . Comparing the exponents of on both sides of (Equation 4.4) yields . Finally, if , then comparing the exponents of on both sides of (Equation 4.4) gives

To complete the proof of (2), it remains to notice that, in view of the already proven part (1), the condition is equivalent to , which makes the last formula equivalent to (Equation 4.3).

To prove the “if” part, it suffices to show that if the matrices satisfy (1)–(2), then the monomials given by the first equality in (Equation 4.2) (with ) satisfy the axioms of an exchange pattern. This is done by a direct check.

Since all matrix mutations are involutive, any choice of an initial vertex and an arbitrary integer matrix gives rise to a unique family of integer matrices satisfying condition (2) in Proposition 4.3 and such that . Thus, the exponents in all monomials are uniquely determined by a single matrix . By Proposition 4.3, in order to determine an exchange pattern, must be such that all matrices obtained from it by a sequence of matrix mutations are sign-skew-symmetric. Verifying that a given matrix has this property seems to be quite nontrivial in general. Fortunately, there is another restriction on that is much easier to check, which implies the desired property, and still leaves us with a large class of matrices sufficient for most applications.

Definition 4.4.

A square integer matrix is called skew-symmetrizable if there exists a diagonal skew-symmetrizing matrix with positive integer diagonal entries such that is skew-symmetric, i.e., for all and .

Proposition 4.5.

For every choice of a vertex and a skew-symmetrizable matrix , there exists a unique family of matrices associated with an exchange pattern on and such that . Furthermore, all the matrices are skew-symmetrizable, sharing the same skew-symmetrizing matrix.

Proof.

The proof follows at once from the following two observations:

1. Every skew-symmetrizable matrix is sign-skew-symmetric.

2. If is skew-symmetrizable and , then is also skew-symmetrizable, with the same skew-symmetrizing matrix.

We call an exchange pattern—and the corresponding cluster algebra—skew-symmetrizable if all the matrices given by (Equation 4.1) (equivalently, one of them) are skew-symmetrizable. In particular, all cluster algebras of rank are skew-symmetrizable: for we have , while for , the calculations in Example 2.5 show that one can take

for all , in the notation of (Equation 2.10)–(Equation 2.11).

Remark 4.6.

Skew-symmetrizable matrices are closely related to symmetrizable (generalized) Cartan matrices appearing in the theory of Kac-Moody algebras. More generally, to every sign-skew-symmetric matrix we can associate a generalized Cartan matrix of the same size by setting

There seem to be deep connections between the cluster algebra corresponding to and the Kac-Moody algebra associated with . We exhibit such a connection for the rank case in Section 6 below. This is however just the tip of an iceberg: a much more detailed analysis will be presented in the sequel to this paper.

In order to show that non-skew-symmetrizable exchange patterns do exist, we conclude this section by exhibiting a 3-parameter family of such patterns of rank .

Proposition 4.7.

Let , , and be three positive integers such that . There exists a unique family of matrices associated with a non-skew-symmetrizable exchange pattern on and such that the matrix at a given vertex is equal to

Proof.

First of all, the matrix is sign-skew-symmetric but not skew-symmetrizable. Indeed, any skew-symmetrizable matrix satisfies the equation . However, this equation for holds only when is equal to or .

For the purpose of this proof only, we refer to a matrix as cyclical if its entries follow one of the two sign patterns

In particular, is cyclical; to prove the proposition, it suffices to show that any matrix obtained from it by a sequence of matrix mutations is also cyclical.

The set of cyclical matrices is not stable under matrix mutations. Let us define some subsets of cyclical matrices that behave nicely with respect to matrix mutations. For a matrix , we denote

For , we say that is -biased if we have

for any . For , we have

so it is -biased for every .

Our proposition becomes an immediate consequence of the following lemma.

Lemma 4.8.

Suppose is cyclical and -biased, and let . Then is cyclical and -biased.

Proof.

Without loss of generality we can assume that and . Denote , and let us write , , etc. By (Equation 4.3), we have therefore, and . We also have

Since is cyclical, the two summands on the right-hand side of each of the equalities in (Equation 4.8) have opposite signs. Note that

and

It follows that (resp., ) has the opposite sign to (resp., ). Thus, is cyclical, and it only remains to show that is -biased. To this effect, we note that , and so

therefore, both and do not exceed . As for , we have , and so

since , we conclude that . This completes the proof that is -biased. Lemma 4.8 and Proposition 4.7 are proved.

5. Exchange relations: The coefficients

In this section we fix a family of matrices satisfying the conditions in Proposition 4.3, and discuss possible choices of coefficients that can appear in the corresponding exchange pattern. We start with the following simple characterization.

Proposition 5.1.

Assume that matrices satisfy conditions (1)–(2) in Proposition 4.3. A family of elements of a coefficient group gives rise, via (Equation 4.2), to an exchange pattern if and only if they satisfy the following relations, whenever :

Note that by (Equation 4.3), so at most one of the and actually enters (Equation 5.1).

Proof.

The only axiom of an exchange pattern that involves the coefficients is (Equation 2.7), and the relation (Equation 5.1) is precisely what it prescribes.

First of all, let us mention the trivial solution of (Equation 5.1) when all the coefficients are equal to .

Moving in the opposite direction, we introduce the universal coefficient group (with respect to a fixed family ) as the abelian group generated by the elements , for all and , which has (Equation 5.1) as the system of defining relations. The torsion-freeness of this group is guaranteed by the following proposition whose straightforward proof is omitted.

Proposition 5.2.

The universal coefficient group is a free abelian group. More specifically, let , and let be a collection of pairs that contains both and for any edge , and precisely one of the pairs and for each edge with and different from . Then is a set of free generators for .

We see that, in contrast to (Equation 4.3), relations (Equation 5.1) leave infinitely many degrees of freedom in determining the coefficients (cf. Remark 2.2).

The rest of this section is devoted to important classes of exchange patterns within which all the coefficients are completely determined by specifying of them corresponding to the edges emanating from a given vertex.

Suppose that, in addition to the multiplicative group structure, the coefficient group is supplied with a binary operation that we call auxiliary addition. Furthermore, suppose that this operation is commutative, associative, and distributive with respect to multiplication; thus is a semifield. (By the way, under these assumptions is automatically torsion-free as a multiplicative group: indeed, if for some and , then

Definition 5.3.

An exchange pattern and the corresponding cluster algebra are called normalized if is a semifield and for any edge .

Proposition 5.4.

Fix a vertex and elements and of a semifield such that for all . Then every family of matrices satisfying the conditions in Proposition 4.3 gives rise to a unique normalized exchange pattern such that, for every edge , we have and . Thus a normalized exchange pattern is completely determined by the monomials and , for all edges .

Proof.

In a normalized exchange pattern, the coefficients and corresponding to an edge are determined by their ratio

via

Clearly, we have

for any edge . We can also rewrite the relation (Equation 5.1) as follows:

for any edge and any . This form of the relations for the normalized coefficients makes our proposition obvious.

Remark 5.5.

It is natural to ask whether the normalization condition imposes additional multiplicative relations among the coefficients that are not consequences of (Equation 5.1). In other words: can a normalized system of coefficients generate the universal coefficient group? In the next section, we present a complete answer to this question in the rank case (see Remark 6.5 below).

One example of a semifield is the multiplicative group of positive real numbers, being ordinary addition. However the following example is more important for our purposes.

Example 5.6.

Let be a free abelian group, written multiplicatively, with a finite set of generators , and with auxiliary addition defined by

Then is a semifield; specifically, it is a product of copies of the tropical semifield (see, e.g., Reference 1). We denote this semifield by . Note that if all exponents and in (Equation 5.6) are nonnegative, then the monomial on the right-hand side is the gcd of the two monomials on the left.

Definition 5.7.

We say that an exchange pattern is of geometric type if it is normalized, has the coefficient semifield , and each coefficient is a monomial in the generators with all exponents nonnegative.

For an exchange pattern of geometric type, the normalization condition simply means that, for every edge , the two monomials and in the generators are coprime, i.e., have no variable in common.

In all our examples of cluster algebras of geometric origin, including those discussed in the introduction, the exchange patterns turn out to be of geometric type. These patterns have the following useful equivalent description.

Proposition 5.8.

Let . A family of coefficients gives rise to an exchange pattern of geometric type if and only if they are given by

for some family of integers ) satisfying the following property: for every edge in , the matrices and are related by

Proof.

As in the proof of Proposition 5.4, for every edge we consider the ratio . We introduce the matrices by setting

The expression (Equation 5.7) then becomes a specialization of the first equality in (Equation 5.3), with auxiliary addition given by (Equation 5.6). To derive (Equation 5.8) from (Equation 5.5), first replace by and by , respectively, then specialize, then pick up the exponent of .

Comparing (Equation 5.8) with (Equation 4.3), we see that it is natural to combine a pair of matrices into one rectangular integer matrix by setting for and . Then the matrices for are related to each other by the same matrix mutation rule (Equation 4.3), now applied to any and . We refer to as the principal part of . Combining Propositions 5.8 and 4.5, we obtain the following corollary.

Corollary 5.9.

Let be an integer matrix with skew-symmetrizable principal part. There is a unique exchange pattern of geometric type such that at a given vertex .

In the geometric type case, there is a distinguished choice of ground ring for the corresponding cluster algebra: take to be the polynomial ring . In the situation of Corollary 5.9, we will denote the corresponding cluster algebra simply by . In the notation of Section 2, is the subring of the ambient field generated by cluster variables for all and together with the generators of .

The set is allowed to be empty: this simply means that all the coefficients in the corresponding exchange pattern are equal to . In this case, we have for all .

We note that the class of exchange patterns of geometric type (and the corresponding cluster algebras ) is stable under the operations of restriction and direct product introduced in Section 2. The restriction from to a subset amounts to removing from the columns labeled by ; the direct product operation replaces two matrices and by the matrix .

6. The rank  case

In this section, we illustrate the above results and constructions by treating in detail the special case . We label vertices and edges of the tree as in (Equation 2.10). For , it will be convenient to denote by the element of congruent to modulo . Thus, consists of vertices and edges for all . We use the notation of Example 2.5, so the clusters are of the form , and the exchange relations are given by (Equation 2.11), with coefficients and satisfying (Equation 2.12). More specifically, we have

for . Choose as the initial cluster. According to Theorem 3.1, each cluster variable can be expressed as a Laurent polynomial in and with coefficients in . Let us write this Laurent polynomial as

where is a polynomial not divisible by or . We will investigate in detail the denominators of these Laurent polynomials.

Recall that for the matrices are given by (Equation 4.5). Thus, all these matrices have the same associated Cartan matrix

(see (Equation 4.6)). We will show that the denominators in (Equation 6.2) have a nice interpretation in terms of the root system associated to . Let us recall some basic facts about this root system (cf., e.g., Reference 10).

Let be a lattice of rank with a fixed basis of simple roots. The Weyl group is a group of linear transformations of generated by two simple reflections and whose action in the basis of simple roots is given by

Since both and are involutions, each element of is one of the following:

here both products are of length . It is well known that is finite if and only if ; we shall refer to this as the finite case. The Coxeter number of is the order of in ; it is given by Table 1. In the finite case, is the dihedral group of order , and its elements can be listed as follows: (the identity element), (the longest element), and distinct elements for . In the infinite case, all elements and for are distinct.

A vector is a real root for if it is -conjugate to a simple root. Let denote the set of real roots for . It is known that , where

is the set of positive real roots. In the finite case, has cardinality , and we have

In the infinite case, we have

with all the roots and distinct.

We will represent the denominators in (Equation 6.2) as vectors in the root lattice by setting

for all . In particular, we have and .

Theorem 6.1.

In the rank case (finite or infinite alike), cluster variables are uniquely up to a multiple from determined by their denominators in the Laurent expansions with respect to a given cluster. The set of these denominators is naturally identified with . More precisely:

(i) In the infinite case, we have

In particular, all for have different denominators .

(ii) In the finite case, we have

and for all , so the denominators are periodic with the period . Moreover, for .

Proof.

Before giving a unified proof of Theorem 6.1, we notice that part (ii) can be proved by a direct calculation. We will express all coefficients in terms of , , and for (cf. Proposition 5.2). With the help of Maple we find that in each finite case, the elements for are the following Laurent polynomials in and .

Type : .

Type : .

Type : .

Type : .

Type : .

Type : .

By inspection, the exponents of the variables in the denominators above agree with (Equation 6.6). Furthermore, we see that for , and therefore for any .

Now let us give a unified proof of both parts of Theorem 6.1. For , let us denote

It then follows from the relations (Equation 2.11) that

Starting with and , we use (Equation 6.7) to see that

and also

This proves (Equation 6.5) and (Equation 6.6) for and .

Proceeding by induction on , let us now assume that , and that both vectors and are positive roots given by (Equation 6.6); in the finite case, we also assume that . If is odd, it follows from (Equation 6.7) that

Thus, is also given by (Equation 6.6). The same argument works with even, and, in the infinite case, with negative. This completes the proof of (Equation 6.5) and (Equation 6.6).

We conclude by a unified proof of the periodicity in the finite case. Without loss of generality, we assume that . By the above inductive argument, we know that is given by (Equation 6.6). Thus, we have . It is known—and easy to check—that if (the only case when is odd), and in the other three finite cases. Hence in all cases, and we are done.

In the rest of the section, we discuss the normalized case (see Definition 5.3). In this context, the periodicity property in Theorem 6.1 (ii) can be sharpened considerably.

Proposition 6.2.

In the normalized finite case of a cluster algebra of rank , we have , , and for all .

Proof.

We set for . Then and are recovered from by

(cf. (Equation 5.3)). To establish the periodicity of and , it suffices to show that . Furthermore, it is enough to check that and in each of the following four cases: (type ), (type ), (type ), and (type ). This is done by direct calculation using the relation

which, in view of (Equation 6.1), is a consequence of (Equation 5.4) and (Equation 5.5). Below are the intermediate steps of this calculation, which in each case expresses for as a rational function in and , and then confirms and .

Type : .

Type : .

Type : .

Type : .

To complete the proof of Proposition 6.2, it remains to show that and in each of the four cases. By Theorem 6.1 (ii), both ratios and belong to , and we only need to show that these two elements of are equal to in the normalized case. The ratios in question were explicitly computed in the course of the proof of Theorem 6.1. We see that it all boils down to the following identities:

All these identities can be proved by a direct calculation (preferably, with a computer): first express both sides in terms of the using (Equation 6.8), and then replace each by the above expression in terms of and .

Remark 6.3.

Periodicity of the recurrence (Equation 6.9) is a very special case of the periodicity phenomenon for -system recurrences in the theory of the thermodynamic Bethe ansatz Reference 19. We plan to address the case of arbitrary rank in a forthcoming paper.

We have shown that in a finite normalized case, any coefficient can be written as a monomial in . There is a nice uniform way to write down these monomials using the dual root system of the system considered above. Recall that the dual root system is associated with the transposed Cartan matrix. The simple roots of are called simple coroots and denoted and . They generate the coroot lattice . The Weyl groups of and are naturally isomorphic to each other, and it is common to identify them. The same Weyl group acts on so that the action of simple reflections and in the basis of simple coroots is given by the matrices

(cf. (Equation 6.4)). By inspection (cf. (Equation 6.10)–Equation 6.13)), one obtains the following unified expression for .

Proposition 6.4.

For and , let denote the coefficient of in the root . Then, in every finite case, we have

for all .

Remark 6.5.

The relations (Equation 6.14) together with the periodicity relation imply that in each finite normalized case the subgroup of the coefficient group generated by all and is in fact generated by . Thus, this group is dramatically different from the universal coefficient group. One might ask if the normalization condition imposes some relations among . The answer to this question is negative: it is possible to take (cf. Example 5.6), so that multiplicatively, is a free abelian group generated by . With the help of (Equation 6.14) and the periodicity relation , we can view all and for as elements of . By the definition of auxiliary addition (see (Equation 5.6)), we have for all . Hence these coefficients give rise to a normalized exchange pattern (which is in fact of geometric type according to Definition 5.7).

In the infinite case, the situation is very different: there exists a normalized exchange pattern such that the coefficients and generate the universal coefficient group (see Proposition 5.2). To see this, take to be the group of formal infinite products , where the exponents can be arbitrary integers. Make into a semifield by setting

Taking (Equation 6.14) as an inspiration, let us define the elements for all by

with the same exponents as in (Equation 6.14). Also, view each as an element of via . A direct check then shows that these coefficients give rise to a normalized exchange pattern (that is, satisfy (Equation 2.12)), and also that in there are no relations among the elements , , and , for .

7. The exchange graph

In this section, we restrict our attention to normalized exchange patterns . Let be the family of cluster variables associated to . As in Section 2, we view them as elements of the ambient field .

A normalized exchange pattern gives rise to a natural equivalence relation on the set of vertices of . Informally speaking, two vertices are equivalent with respect to if they have the same clusters and the same exchange relations, up to relabeling of cluster variables. Here is a precise definition.

Definition 7.1.

We say that two vertices and in are -equivalent if there exists a permutation of the index set such that:

(1) for all ;

(2) if and for some , then and .

Remark 7.2.

We believe that condition (2) in Definition 7.1 is in fact a consequence of (1). Thus, once a cluster has been fixed as a set (forgetting the labels and the cluster’s location in the tree ), the exchange relations involving its elements are uniquely determined.

In view of Proposition 5.4, a normalized exchange pattern can be defined in terms of a family of integer matrices together with a family of elements satisfying relations (Equation 5.4)–(Equation 5.5). In these terms, property (2) above can be rephrased as saying that and for all .

The following property is an immediate consequence of Proposition 5.4.

Proposition 7.3.

Suppose and are -equivalent. For every vertex adjacent to , there is a unique vertex adjacent to and equivalent to . In the notation of Definition 7.1, if for some , then .

Definition 7.4.

The exchange graph associated with a normalized exchange pattern has the -equivalence classes as vertices, joined by an edge if they have adjacent representatives in .

As an immediate corollary of Proposition 7.3, we obtain the following.

Corollary 7.5.

The exchange graph is -regular, i.e., every vertex has precisely edges emanating from it.

Note that passing from to may result in losing the coloring of edges by elements of the index set . This will happen if a permutation in Definition 7.1 is nontrivial. (This permutation can be thought of as “discrete monodromy” of our graph.)

Example 7.6.

The results in Section 6 (more specifically, Theorem 6.1 and Proposition 6.2) show that, in the rank case, the -equivalence and the exchange graph can be described as follows. In the infinite case, no two distinct vertices of are -equivalent to each other, so the exchange graph is itself. In the finite case, and are -equivalent if and only if , so the exchange graph is a cycle of length . For example, in the type case, we have , and the graph is a pentagon. Figure 3 shows the corresponding exchange graph, together with its clusters and exchange relations, written with the help of (Equation 2.11) and (Equation 6.11). The discrete monodromy is present in this special case, as the variables and get switched after a full 5-cycle of exchanges:

The corresponding cluster algebra can be realized as the homogeneous coordinate ring of the Grassmannian ; see Example 2.5.

We see that, in the rank case, the following conditions on a normalized exchange pattern are equivalent:

(1) The exchange graph is finite.

(2) A cluster algebra associated to has finitely many distinct cluster variables.

(3) There exists a vertex such that the Cartan counterpart of the matrix is a Cartan matrix of finite type, i.e., we have .

In the sequel to this paper, we shall extend this result to arbitrary rank.

Returning to the general case, we shall now provide some important instances of -equivalence.

Theorem 7.7.

Suppose for some vertex and two distinct indices and . Let be the Coxeter number of the corresponding rank root system, and let be a vertex joined with by a path of length whose edge labels alternate between and . Then is -equivalent to , so that the path becomes a cycle of length in the exchange graph.

Proof.

For the sake of convenience, assume that , ; thus, the path joining and has the form

(cf. (Equation 2.10)), where as before stands for the element of congruent to modulo . We abbreviate and , and assume (without loss of generality) that and are either both positive or both equal to . This agrees with the convention (Equation 4.5) (also used in Section 6), where was denoted by , and by . Recall that the value of is given by Table 1 in Section 6; in particular, is even unless (the type ). We claim that the conditions (1) and (2) in Definition 7.1 hold with the permutation equal to identity for even, and equal to the transposition of indices and in the only case when is odd. Using the restriction operation on exchange patterns (cf. Proposition 2.6), it suffices to treat the case , so we can take .

In view of the exchange property (Equation 2.1), the cluster variable does not depend on . Therefore, verifying condition (1) amounts to checking that for . Again using restriction, we can assume for the purpose of this checking only that and , in which case the required property was established in Proposition 6.2.

The same argument shows that the only part of condition (2) that does not follow from Proposition 6.2 is the equalities and for . Let us abbreviate and ; the latter notation is chosen to be consistent with the notation in the proof of Proposition 6.2. Iterating (Equation 5.5) (with = 3), we obtain

Thus, we need to show that the product on the right-hand side of (Equation 7.1) is equal to . To do this, we need some preparation.

Recall that the coefficients satisfy the relation (Equation 6.9), which in our present notation can be rewritten as

On the other hand, the matrix mutation rules (Equation 4.3) (with , applied along the edges ) readily imply the following recurrence for the exponents :

The sequence is periodic with period , by virtue of Proposition 6.2. Furthermore, the relation (Equation 7.3) can be seen, somewhat surprisingly, as a special case of (Equation 7.2), for the following version of the tropical semifield: take with the multiplication given by ordinary addition, and the auxiliary addition given by . This implies in particular that the sequence is also periodic with period . We also note that the same periodicity holds for the sequence : when is even, this is clear from the definition of ; and in the only case when is odd, we have .

To show that the right-hand side of (Equation 7.1) is equal to , we first treat the case . Then , and we have and for all . It follows that

as desired.

In the remaining case when and are both positive, we use (Equation 7.2) and (Equation 7.3) as well as the above-mentioned periodicity, to obtain:

To conclude that , just recall that the coefficient group is assumed to be torsion-free.

To complete the proof of the theorem, it remains to show that for . But this is equivalent to saying that the sequence is periodic with period , which is already proven.

We conjecture that the -equivalence relation is generated by its instances described in Theorem 7.7.

Example 7.8.

Let be an exchange pattern of geometric type associated with the skew-symmetric matrix

as prescribed by Corollary 5.9. Thus, the corresponding cluster algebra is of rank , and all the coefficients are equal to . The exchange graph for this pattern is a “two-layer brick wall” shown in Figure 4. In this figure, distinct cluster variables are associated with regions: there are variables associated with bounded regions (“bricks”), and two more variables and associated with the two unbounded regions. The cluster at any vertex consists of the three cluster variables associated to the three regions adjacent to . The binomial exchange relations are as follows: for , we have

To see all this, we first show that the graph in Figure 4 is a cover of the exchange graph . Pick an initial vertex with the matrix given by (Equation 7.4) (abusing notation, we will use the same symbol for a vertex in and its image in ). Denote the cluster variables at by , , and , so that their order agrees with that of rows and columns of . Since every principal submatrix of is of type , Theorem 7.7 provides that is a common vertex of three -cycles in . These cycles are depicted in Figure 4 as perimeters of the three bricks surrounding . The variable inside each brick indicates that this brick corresponds to the rank 2 exchange pattern obtained from via restriction from to ; equivalently, this means that appears in every cluster on the perimeter of the brick. Now we move from to an adjacent vertex with the cluster . According to Proposition 4.3,

This matrix differs from by a simultaneous cyclic permutation of rows and columns. Therefore, we can apply the same construction to , obtaining the three bricks surrounding this vertex. Continuing in the same way, we produce the entire graph in Figure 4. Since this graph is -regular, we have covered all the vertices and edges of the exchange graph.

For every vertex on the median of the wall, the matrix is obtained from by a simultaneous permutation of rows and columns. Now take a vertex on the outer boundary, say the vertex in Figure 5. Then

Again, for every other vertex on the outer boundary, the corresponding matrix is obtained from by a simultaneous permutation of rows and columns. Substituting the matrices into (Equation 4.2), we generate all the exchange relations (Equation 7.5) (cf. Figure 5).

To prove that our brick wall is indeed the exchange graph, it remains to show that all the cluster variables , , and are distinct; recall that we view them as elements of the ambient field . Following the methodology of Theorem 6.1, it suffices to show that all of them have different denominators in the Laurent expansion in terms of the initial cluster . We write the denominator of a cluster variable as (cf. (Equation 6.2)), and encode it by a vector . In particular, we have

A moment’s reflection shows that the vectors , and satisfy the “tropical version” of each of the exchange relations (Equation 7.5) (with multiplication replaced by vector addition, and addition replaced by component-wise maximum), as long as the two vectors on the right-hand side are different. For example, the first equation in (Equation 7.5) will take the form

provided ; here stands for component-wise maximum. All ’s are uniquely determined from (Equation 7.6) by recursive application of these conditions. As a result, we obtain, for every :

(cf. Figure 6, where each vector is shown within the corresponding region).

Acknowledgments

This work began in May 2000 when the authors got together as participants in the program “Representation Theory-2000,” organized by Victor Kac and Alexander A. Kirillov at the Erwin Schrödinger International Institute for Mathematical Physics in Vienna, Austria. We thank the organizers for inviting us, and we are grateful to Peter Michor and the staff of the Institute for creating ideal working conditions. The paper was finished at the Isaac Newton Institute for Mathematical Sciences in Cambridge, UK, whose support is gratefully acknowledged.

It was during our stay in Vienna that we learned of the terminal illness of Sergei Kerov, an outstanding mathematician and a good friend of ours. He passed away on July 30, 2000. We dedicate this paper to his memory.

Figures

Figure 1.

The Grassmannian

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{4.0pt} \begin{picture}(40,34)(-4,-2) \put(6,0){\line(1,0){20}} \put(0,19){\line(1,0){32}} \qbezier(6,0)(11,15.5)(16,31) \qbezier(26,0)(21,15.5)(16,31) \qbezier(6,0)(3,9.5)(0,19) \qbezier(26,0)(29,9.5)(32,19) \qbezier(0,19)(13,9.5)(26,0) \qbezier(32,19)(19,9.5)(6,0) \qbezier(0,19)(8,25)(16,31) \qbezier(32,19)(24,25)(16,31) \par\put(6,0){\circle*{1}} \put(26,0){\circle*{1}} \put(0,19){\circle*{1}} \put(32,19){\circle*{1}} \put(16,31){\circle*{1}} \par\put(5,-2){\makebox(0,0){$\bar1$}} \put(27,-2){\makebox(0,0){$\bar5$}} \put(-1,21){\makebox(0,0){$\bar2$}} \put(33,21){\makebox(0,0){$\bar4$}} \put(16,33){\makebox(0,0){$\bar3$}} \put(16,-2){\makebox(0,0){$q_4$}} \put(1.5,9){\makebox(0,0){$q_2$}} \put(30.5,9){\makebox(0,0){$q_1$}} \put(7.5,26.5){\makebox(0,0){$q_5$}} \put(24.5,26.5){\makebox(0,0){$q_3$}} \put(16,20.5){\makebox(0,0){$y_4$}} \put(23,16){\makebox(0,0){$y_2$}} \put(9.5,16){\makebox(0,0){$y_1$}} \put(12,8.5){\makebox(0,0){$y_3$}} \put(20.5,8.5){\makebox(0,0){$y_5$}} \par\end{picture}
Figure 2.

The “caterpillar” tree , for ,

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{picture}(280,60)(0,0) \thicklines\put(0,30){\line(1,0){280}} \multiput(20,30)(40,0){7}{\vector(1,0){7}} \multiput(0,30)(40,0){8}{\line(1,-2){10}} \multiput(0,30)(40,0){8}{\line(-1,-2){10}} \put(0,30){\line(-2,1){20}} \put(280,30){\line(2,1){20}} \multiput(0,30)(40,0){8}{\circle*{4}} \multiput(-10,10)(20,0){16}{\circle*{4}} \put(-20,40){\circle*{4}} \put(300,40){\circle*{4}} \put(-30,47){$t_{\textup{tail}}$} \put(-8,37){$t_{\textup{base}}$} \put(290,47){$t_{\textup{head}}$} \end{picture}
Table 1.

The Coxeter number

0123
2346
Figure 3.

The exchange graph for a cluster algebra of type

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{4.2pt} \begin{picture}(35,35)(-2,-2) \put(6,0){\line(1,0){20}} \qbezier(6,0)(3,9.5)(0,19) \qbezier(26,0)(29,9.5)(32,19) \qbezier(0,19)(8,25)(16,31) \qbezier(32,19)(24,25)(16,31) \par\put(6,0){\circle*{1}} \put(26,0){\circle*{1}} \put(0,19){\circle*{1}} \put(32,19){\circle*{1}} \put(16,31){\circle*{1}} \par\put(2,0){\makebox(0,0){$y_5,y_1$}} \put(30,0){\makebox(0,0){$y_4,y_5$}} \put(-4,19){\makebox(0,0){$y_1,y_2$}} \put(36,19){\makebox(0,0){$y_3,y_4$}} \put(16,33){\makebox(0,0){$y_2,y_3$}} \par\put(-2,26.5){\makebox(0,0){$y_1 y_3=q_2 y_2 + q_4 q_5$}} \put(34,26.5){\makebox(0,0){$y_2 y_4=q_3 y_3 + q_5 q_1$}} \put(39.5,9){\makebox(0,0){$y_3 y_5=q_4 y_4 + q_1 q_2$}} \put(-7.5,9){\makebox(0,0){$y_5 y_2=q_1 y_1 + q_2 q_4$}} \put(16,-2.5){\makebox(0,0){$y_4 y_1=q_5 y_5 + q_2 q_3$}} \par\end{picture}
Figure 4.

The two-layer brick wall

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{2.0pt} \begin{picture}(140,55)(0,0) \multiput(0,10)(0,20){3}{\line(1,0){140}} \multiput(10,30)(40,0){4}{\line(0,1){20}} \multiput(30,10)(40,0){3}{\line(0,1){20}} \par\multiput(10,30)(40,0){4}{\circle*{2}} \multiput(10,50)(40,0){4}{\circle*{2}} \multiput(30,30)(40,0){3}{\circle*{2}} \multiput(30,10)(40,0){3}{\circle*{2}} \par\put(10,20){\makebox(0,0){$y_0$}} \put(50,20){\makebox(0,0){$y_2$}} \put(90,20){\makebox(0,0){$y_4$}} \put(130,20){\makebox(0,0){$y_6$}} \par\put(30,40){\makebox(0,0){$y_1$}} \put(70,40){\makebox(0,0){$y_3$}} \put(110,40){\makebox(0,0){$y_5$}} \par\put(53,33){\makebox(0,0){$t_0$}} \put(73,33){\makebox(0,0){$t_1$}} \put(70,2){\makebox(0,0){$z$}} \put(70,58){\makebox(0,0){$w$}} \par\end{picture}
Figure 5.

Close-up of a brick

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{6.0pt} \begin{picture}(50,30)(0,-4) \multiput(0,0)(0,20){2}{\line(1,0){50}} \multiput(25,20)(40,0){1}{\line(0,1){5}} \multiput(5,0)(40,0){2}{\line(0,1){20}} \par\multiput(5,0)(40,0){2}{\circle*{0.7}} \multiput(5,20)(20,0){3}{\circle*{0.7}} \par\put(15,19){\makebox(0,0){$y_0 y_3\!=\!y_1 y_2 \!+\! 1$}} \put(35,19){\makebox(0,0){$y_1 y_4\!=\!y_2 y_3 \!+\! 1$}} \put(10,10){\makebox(0,0){$y_1 z\!=\!y_0\!+\! y_2$}} \put(40,10){\makebox(0,0){$y_3 z\!=\!y_2\!+\! y_4$}} \put(25,1.5){\makebox(0,0){$y_0 y_4\!=\!y_2^2\!+\! z$}} \put(26.5,21){\makebox(0,0){$t_0$}} \put(4,18.5){\makebox(0,0){$t_4$}} \put(46.5,18.5){\makebox(0,0){$t_1$}} \put(4,1.5){\makebox(0,0){$t_3$}} \put(46.5,1.5){\makebox(0,0){$t_2$}} \par\put(5,24){\makebox(0,0){$y_1$}} \put(45,24){\makebox(0,0){$y_3$}} \put(25,10){\makebox(0,0){$y_2$}} \put(1,10){\makebox(0,0){$y_0$}} \put(49,10){\makebox(0,0){$y_4$}} \put(25,-3){\makebox(0,0){$z$}} \par\end{picture}
Figure 6.

The brick wall denominators

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.38pt} \begin{picture}(260,63)(0,0) \multiput(0,10)(0,20){3}{\line(1,0){260}} \multiput(10,30)(40,0){7}{\line(0,1){20}} \multiput(30,10)(40,0){6}{\line(0,1){20}} \par\multiput(10,30)(40,0){7}{\circle*{2}} \multiput(10,50)(40,0){7}{\circle*{2}} \multiput(30,30)(40,0){6}{\circle*{2}} \multiput(30,10)(40,0){6}{\circle*{2}} \par\par\par\put(10,20){\makebox(0,0){$\begin{array}{c}y_{-4}\\(2,2,3)\end{array}$}} \put(50,20){\makebox(0,0){$\begin{array}{c}y_{-2}\\(1,1,2)\end{array}$}} \put(90,20){\makebox(0,0){$\begin{array}{c}y_0\\(0,0,1)\end{array}$}} \put(130,20){\makebox(0,0){$\begin{array}{c}y_2\\(0,-1,0)\end{array}$}} \put(170,20){\makebox(0,0){$\begin{array}{c}y_4\\(1,0,0)\end{array}$}} \put(210,20){\makebox(0,0){$\begin{array}{c}y_6\\(2,1,1)\end{array}$}} \put(250,20){\makebox(0,0){$\begin{array}{c}y_8\\(3,2,2)\end{array}$}} \par\put(30,40){\makebox(0,0){$\begin{array}{c}y_{-3}\\(1,2,2)\end{array}$}} \put(70,40){\makebox(0,0){$\begin{array}{c}y_{-1}\\(0,1,1)\end{array}$}} \put(110,40){\makebox(0,0){$\begin{array}{c}y_1\\(-1,0,0)\end{array}$}} \put(150,40){\makebox(0,0){$\begin{array}{c}y_3\\(0,0,-1)\end{array}$}} \put(190,40){\makebox(0,0){$\begin{array}{c}y_5\\(1,1,0)\end{array}$}} \put(230,40){\makebox(0,0){$\begin{array}{c}y_7\\(2,2,1)\end{array}$}} \par\put(130,0){\makebox(0,0){$\begin{array}{c}z\\(1,0,1)\end{array}$}} \put(130,61){\makebox(0,0){$\begin{array}{c}w\\(0,1,0)\end{array}$}} \par\end{picture}

Mathematical Fragments

Equation (1.1)
Equations (2.1), (2.2)
Equation (2.3)
Definition 2.1.

An exchange pattern on with coefficients in is a family of monomials of the form (Equation 2.3) satisfying the following four axioms:

Equation (2.8)
Remark 2.2.

Informally speaking, axiom (Equation 2.7) describes the propagation of an exchange pattern along the edges of . More precisely, let us fix the exchange monomials for all edges emanating from a given vertex . This choice uniquely determines the ratio for any vertex adjacent to and any edge (to see this, take and in (Equation 2.7), and allow to vary). In view of (Equation 2.5), this ratio in turn uniquely determines the exponents of all variables in both monomials and . There remains, however, one degree of freedom in determining the coefficients and because only their ratio is prescribed by (Equation 2.7). In Section 5 we shall introduce an important class of normalized exchange patterns for which this degree of freedom disappears, and so the whole pattern is uniquely determined by the monomials associated with edges emanating from a given vertex.

Equation (2.9)
Example 2.5.

Consider the case . The tree is shown below:

Let us denote the cluster variables as follows:

(the above equalities among the cluster variables follow from (Equation 2.1)). Then the clusters look like

We claim that the exchange relations (Equation 2.2) can be written in the following form:

where the integers and are either both positive or both equal to , and the coefficients and are elements of satisfying the relations

Furthermore, any such choice of parameters , , , results in a well-defined cluster algebra of rank 2.

To prove this, we notice that, in view of (Equation 2.4)–(Equation 2.5), both monomials and do not contain the variable , and at most one of them contains . If enters neither nor , then these two are simply elements of . But then (Equation 2.6) forces all monomials  to be elements of , while (Equation 2.7) implies that it is possible to give the names and to the two monomials corresponding to each edge so that (2.11)–(2.12) hold with .

Next, consider the case when precisely one of the monomials and contains . Applying if necessary the involution (Equation 2.8) to our exchange pattern, we may assume that and for some positive integer and some . Thus, the exchange relation associated to the edge takes the form . By (Equation 2.6), we have and for some positive integer and some . Then the exchange relation for the edge takes the form . At this point, we invoke (Equation 2.7):

By (Equation 2.5), we have and for some satisfying . Continuing in the same way, we obtain all relations (2.11)–(2.12).

For fixed and , the “universal” coefficient group is the multiplicative abelian group generated by the elements and for all subject to the defining relations (2.12). It is easy to see that this is a free abelian group of infinite rank. As a set of its free generators, one can choose any subset of that contains four generators and precisely one generator from each pair for .

A nice specialization of this setup is provided by the homogeneous coordinate ring of the Grassmannian . Recall (cf. Section 1) that this ring is generated by the Plücker coordinates , where and are distinct elements of the cyclic group . We shall write for , and adopt the convention ; see Figure 1.

The ideal of relations among the Plücker coordinates is generated by the relations

for . A direct check shows that these relations are a specialization of the relations (2.11), if we set , , , and for all . The coefficient group is the multiplicative free abelian group with generators . It is also immediate that the elements and defined in this way satisfy the relations (2.12).

Proposition 2.6.

Let be a cluster algebra of rank associated with an exchange pattern , and let be obtained from by restriction from to using a connected component . The -subalgebra of generated by is naturally identified with the cluster algebra , where is the polynomial ring .

Theorem 3.1.

In a cluster algebra, any cluster variable is expressed in terms of any given cluster as a Laurent polynomial with coefficients in the group ring .

Equations (3.1), (3.2)
Theorem 3.2.

If conditions (Equation 3.1)–(Equation 3.2) are satisfied, then each element of the cluster is a Laurent polynomial in the cluster , with coefficients in the ring .

Lemma 3.3.

The clusters , , and are contained in . Furthermore, as elements of .

Equations (3.3), (3.4), (3.5)
Equation (4.1)
Equation (4.2)
Definition 4.2.

Let and be square integer matrices of the same size. We say that is obtained from by the matrix mutation in direction  and write if

Proposition 4.3.

A family of integer matrices corresponds to an exchange pattern if and only if the following conditions hold:

(1) is sign-skew-symmetric for any .

(2) If , then .

Equation (4.4)
Proposition 4.5.

For every choice of a vertex and a skew-symmetrizable matrix , there exists a unique family of matrices associated with an exchange pattern on and such that . Furthermore, all the matrices are skew-symmetrizable, sharing the same skew-symmetrizing matrix.

Equation (4.5)
Remark 4.6.

Skew-symmetrizable matrices are closely related to symmetrizable (generalized) Cartan matrices appearing in the theory of Kac-Moody algebras. More generally, to every sign-skew-symmetric matrix we can associate a generalized Cartan matrix of the same size by setting

There seem to be deep connections between the cluster algebra corresponding to and the Kac-Moody algebra associated with . We exhibit such a connection for the rank case in Section 6 below. This is however just the tip of an iceberg: a much more detailed analysis will be presented in the sequel to this paper.

Proposition 4.7.

Let , , and be three positive integers such that . There exists a unique family of matrices associated with a non-skew-symmetrizable exchange pattern on and such that the matrix at a given vertex is equal to

Lemma 4.8.

Suppose is cyclical and -biased, and let . Then is cyclical and -biased.

Equation (4.8)
Proposition 5.1.

Assume that matrices satisfy conditions (1)–(2) in Proposition 4.3. A family of elements of a coefficient group gives rise, via (Equation 4.2), to an exchange pattern if and only if they satisfy the following relations, whenever :

Proposition 5.2.

The universal coefficient group is a free abelian group. More specifically, let , and let be a collection of pairs that contains both and for any edge , and precisely one of the pairs and for each edge with and different from . Then is a set of free generators for .

Definition 5.3.

An exchange pattern and the corresponding cluster algebra are called normalized if is a semifield and for any edge .

Proposition 5.4.

Fix a vertex and elements and of a semifield such that for all . Then every family of matrices satisfying the conditions in Proposition 4.3 gives rise to a unique normalized exchange pattern such that, for every edge , we have and . Thus a normalized exchange pattern is completely determined by the monomials and , for all edges .

Equation (5.3)
Equation (5.4)
Equation (5.5)
Example 5.6.

Let be a free abelian group, written multiplicatively, with a finite set of generators , and with auxiliary addition defined by

Then is a semifield; specifically, it is a product of copies of the tropical semifield (see, e.g., Reference 1). We denote this semifield by . Note that if all exponents and in (5.6) are nonnegative, then the monomial on the right-hand side is the gcd of the two monomials on the left.

Definition 5.7.

We say that an exchange pattern is of geometric type if it is normalized, has the coefficient semifield , and each coefficient is a monomial in the generators with all exponents nonnegative.

Proposition 5.8.

Let . A family of coefficients gives rise to an exchange pattern of geometric type if and only if they are given by

for some family of integers ) satisfying the following property: for every edge in , the matrices and are related by

Corollary 5.9.

Let be an integer matrix with skew-symmetrizable principal part. There is a unique exchange pattern of geometric type such that at a given vertex .

Equation (6.1)
Equation (6.2)
Equation (6.4)
Theorem 6.1.

In the rank case (finite or infinite alike), cluster variables are uniquely up to a multiple from determined by their denominators in the Laurent expansions with respect to a given cluster. The set of these denominators is naturally identified with . More precisely:

(i) In the infinite case, we have

In particular, all for have different denominators .

(ii) In the finite case, we have

and for all , so the denominators are periodic with the period . Moreover, for .

Equation (6.7)
Proposition 6.2.

In the normalized finite case of a cluster algebra of rank , we have , , and for all .

Equation (6.8)
Equation (6.9)
Equations (6.10), (6.11), (6.12), (6.13)
Proposition 6.4.

For and , let denote the coefficient of in the root . Then, in every finite case, we have

for all .

Remark 6.5.

The relations (Equation 6.14) together with the periodicity relation imply that in each finite normalized case the subgroup of the coefficient group generated by all and is in fact generated by . Thus, this group is dramatically different from the universal coefficient group. One might ask if the normalization condition imposes some relations among . The answer to this question is negative: it is possible to take (cf. Example 5.6), so that multiplicatively, is a free abelian group generated by . With the help of (Equation 6.14) and the periodicity relation , we can view all and for as elements of . By the definition of auxiliary addition (see (Equation 5.6)), we have for all . Hence these coefficients give rise to a normalized exchange pattern (which is in fact of geometric type according to Definition 5.7).

In the infinite case, the situation is very different: there exists a normalized exchange pattern such that the coefficients and generate the universal coefficient group (see Proposition 5.2). To see this, take to be the group of formal infinite products , where the exponents can be arbitrary integers. Make into a semifield by setting

Taking (Equation 6.14) as an inspiration, let us define the elements for all by

with the same exponents as in (Equation 6.14). Also, view each as an element of via . A direct check then shows that these coefficients give rise to a normalized exchange pattern (that is, satisfy (Equation 2.12)), and also that in there are no relations among the elements , , and , for .

Definition 7.1.

We say that two vertices and in are -equivalent if there exists a permutation of the index set such that:

(1) for all ;

(2) if and for some , then and .

Proposition 7.3.

Suppose and are -equivalent. For every vertex adjacent to , there is a unique vertex adjacent to and equivalent to . In the notation of Definition 7.1, if for some , then .

Example 7.6.

The results in Section 6 (more specifically, Theorem 6.1 and Proposition 6.2) show that, in the rank case, the -equivalence and the exchange graph can be described as follows. In the infinite case, no two distinct vertices of are -equivalent to each other, so the exchange graph is itself. In the finite case, and are -equivalent if and only if , so the exchange graph is a cycle of length . For example, in the type case, we have , and the graph is a pentagon. Figure 3 shows the corresponding exchange graph, together with its clusters and exchange relations, written with the help of (Equation 2.11) and (Equation 6.11). The discrete monodromy is present in this special case, as the variables and get switched after a full 5-cycle of exchanges:

The corresponding cluster algebra can be realized as the homogeneous coordinate ring of the Grassmannian ; see Example 2.5.

We see that, in the rank case, the following conditions on a normalized exchange pattern are equivalent:

(1) The exchange graph is finite.

(2) A cluster algebra associated to has finitely many distinct cluster variables.

(3) There exists a vertex such that the Cartan counterpart of the matrix is a Cartan matrix of finite type, i.e., we have .

In the sequel to this paper, we shall extend this result to arbitrary rank.

Theorem 7.7.

Suppose for some vertex and two distinct indices and . Let be the Coxeter number of the corresponding rank root system, and let be a vertex joined with by a path of length whose edge labels alternate between and . Then is -equivalent to , so that the path becomes a cycle of length in the exchange graph.

Equation (7.1)
Equation (7.2)
Equation (7.3)
Example 7.8.

Let be an exchange pattern of geometric type associated with the skew-symmetric matrix

as prescribed by Corollary 5.9. Thus, the corresponding cluster algebra is of rank , and all the coefficients are equal to . The exchange graph for this pattern is a “two-layer brick wall” shown in Figure 4. In this figure, distinct cluster variables are associated with regions: there are variables associated with bounded regions (“bricks”), and two more variables and associated with the two unbounded regions. The cluster at any vertex consists of the three cluster variables associated to the three regions adjacent to . The binomial exchange relations are as follows: for , we have

To see all this, we first show that the graph in Figure 4 is a cover of the exchange graph . Pick an initial vertex with the matrix given by (7.4) (abusing notation, we will use the same symbol for a vertex in and its image in ). Denote the cluster variables at by , , and , so that their order agrees with that of rows and columns of . Since every principal submatrix of is of type , Theorem 7.7 provides that is a common vertex of three -cycles in . These cycles are depicted in Figure 4 as perimeters of the three bricks surrounding . The variable inside each brick indicates that this brick corresponds to the rank 2 exchange pattern obtained from via restriction from to ; equivalently, this means that appears in every cluster on the perimeter of the brick. Now we move from to an adjacent vertex with the cluster . According to Proposition 4.3,

This matrix differs from by a simultaneous cyclic permutation of rows and columns. Therefore, we can apply the same construction to , obtaining the three bricks surrounding this vertex. Continuing in the same way, we produce the entire graph in Figure 4. Since this graph is -regular, we have covered all the vertices and edges of the exchange graph.

For every vertex on the median of the wall, the matrix is obtained from by a simultaneous permutation of rows and columns. Now take a vertex on the outer boundary, say the vertex in Figure 5. Then

Again, for every other vertex on the outer boundary, the corresponding matrix is obtained from by a simultaneous permutation of rows and columns. Substituting the matrices into (Equation 4.2), we generate all the exchange relations (7.5) (cf. Figure 5).

To prove that our brick wall is indeed the exchange graph, it remains to show that all the cluster variables , , and are distinct; recall that we view them as elements of the ambient field . Following the methodology of Theorem 6.1, it suffices to show that all of them have different denominators in the Laurent expansion in terms of the initial cluster . We write the denominator of a cluster variable as (cf. (Equation 6.2)), and encode it by a vector . In particular, we have

A moment’s reflection shows that the vectors , and satisfy the “tropical version” of each of the exchange relations (7.5) (with multiplication replaced by vector addition, and addition replaced by component-wise maximum), as long as the two vectors on the right-hand side are different. For example, the first equation in (7.5) will take the form

provided ; here stands for component-wise maximum. All ’s are uniquely determined from (7.6) by recursive application of these conditions. As a result, we obtain, for every :

(cf. Figure 6, where each vector is shown within the corresponding region).

References

Reference [1]
A. Berenstein, S. Fomin and A. Zelevinsky, Parametrizations of canonical bases and totally positive matrices, Adv. Math. 122 (1996), 49–149. MR 98j:17008
Reference [2]
A. Berenstein and A. Zelevinsky, String bases for quantum groups of type , in I. M. Gelfand Seminar, 51–89, Adv. Soviet Math. 16, Part 1, Amer. Math. Soc., Providence, RI, 1993. MR 94g:17019
Reference [3]
A. Berenstein and A. Zelevinsky, Total positivity in Schubert varieties, Comment. Math. Helv. 72 (1997), 128–166. MR 99g:14064
Reference [4]
A. Berenstein and A. Zelevinsky, Tensor product multiplicities, canonical bases and totally positive varieties, Invent. Math. 143 (2001), 77–128. CMP 2001:06
Reference [5]
S. Fomin and A. Zelevinsky, Double Bruhat cells and total positivity, J. Amer. Math. Soc. 12 (1999), 335–380. MR 2001f:20097
Reference [6]
S. Fomin and A. Zelevinsky, Total positivity: tests and parametrizations, Math. Intelligencer 22 (2000), no. 1, 23–33. MR 2001b:15030
Reference [7]
S. Fomin and A. Zelevinsky, Totally nonnegative and oscillatory elements in semisimple groups, Proc. Amer. Math. Soc. 128 (2000), 3749–3759. MR 2001b:22014
Reference [8]
S. Fomin and A. Zelevinsky, The Laurent phenomenon, to appear in Adv. in Applied Math.
Reference [9]
I. M. Gelfand and A. Zelevinsky, Canonical basis in irreducible representations of and its applications, in: Group theoretical methods in physics, Vol. II (Jurmala, 1985), 127–146, VNU Sci. Press, Utrecht, 1986. MR 89a:17010
Reference [10]
V. Kac, Infinite dimensional Lie algebras, 3rd edition, Cambridge University Press, 1990. MR 92k:17038
Reference [11]
J. Kung and G.-C. Rota, The invariant theory of binary forms, Bull. Amer. Math. Soc. (N.S.) 10 (1984), 27–85. MR 85g:05002
Reference [12]
B. Leclerc and A. Zelevinsky, Quasicommuting families of quantum Plucker coordinates, Amer. Math. Soc. Transl. (2), Vol. 181, Kirillov’s Seminar on Representation Theory, 85–108, Amer. Math. Soc., Providence, RI, 1998. MR 99g:14066
Reference [13]
G. Lusztig, Canonical bases arising from quantized enveloping algebras, J. Amer. Math. Soc. 3 (1990), 447–498. MR 90m:17023
Reference [14]
G. Lusztig, Introduction to quantum groups, Progress in Mathematics 110, Birkhäuser, 1993. MR 94m:17016
Reference [15]
G. Lusztig, Total positivity in reductive groups, in: Lie theory and geometry: in honor of Bertram Kostant, Progress in Mathematics 123, Birkhäuser, 1994. MR 96m:20071
Reference [16]
V. Retakh and A. Zelevinsky, The base affine space and canonical bases in irreducible representations of the group , Soviet Math. Dokl. 37 (1988), no. 3, 618–622. MR 89g:22025
Reference [17]
B. Shapiro, M. Shapiro, A. Vainshtein and A. Zelevinsky, Simply-laced Coxeter groups and groups generated by symplectic transvections, Michigan Math. J. 48 (2000), 531–551. MR 2001g:20050
Reference [18]
B. Sturmfels, Algorithms in invariant theory. Texts and Monographs in Symbolic Computation. Springer-Verlag, Vienna, 1993. MR 94m:13004
Reference [19]
Al. B. Zamolodchikov, On the thermodynamic Bethe ansatz equations for reflectionless scattering theories, Phys. Lett. B 253 (1991), 391–394. MR 92a:81196
Reference [20]
A. Zelevinsky, Connected components of real double Bruhat cells, Intern. Math. Res. Notices 2000, No. 21, 1131–1153. MR 2001k:14094

Article Information

MSC 1991
Primary: 14M99 (None of the above but in this section)
Secondary: 17B99 (None of the above but in this section)
Keywords
  • Cluster algebra
  • exchange pattern
  • Laurent phenomenon
Author Information
Sergey Fomin
Department of Mathematics, University of Michigan, Ann Arbor, Michigan 48109
fomin@umich.edu
ORCID
MathSciNet
Andrei Zelevinsky
Department of Mathematics, Northeastern University, Boston, Massachusetts 02115
andrei@neu.edu
Additional Notes

The authors were supported in part by NSF grants #DMS-0049063, #DMS-0070685 (S.F.), and #DMS-9971362 (A.Z.).

Journal Information
Journal of the American Mathematical Society, Volume 15, 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 2001 by Sergey Fomin and Andrei Zelevinsky
Article References

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.