Canonical bases for cluster algebras

By Mark Gross, Paul Hacking, Sean Keel, and Maxim Kontsevich

Abstract

In an earlier work (Publ. Inst. Hautes Études Sci., 122 (2015), 65–168) the first three authors conjectured that the ring of regular functions on a natural class of affine log Calabi–Yau varieties (those with maximal boundary) has a canonical vector space basis parameterized by the integral tropical points of the mirror. Further, the structure constants for the multiplication rule in this basis should be given by counting broken lines (certain combinatorial objects, morally the tropicalizations of holomorphic discs).

Here we prove the conjecture in the case of cluster varieties, where the statement is a more precise form of the Fock–Goncharov dual basis conjecture (Publ. Inst. Hautes Études Sci., 103 (2006), 1–211). In particular, under suitable hypotheses, for each the partial compactification of an affine cluster variety given by allowing some frozen variables to vanish, we obtain canonical bases for extending to a basis of . Each choice of seed canonically identifies the parameterizing sets of these bases with integral points in a polyhedral cone. These results specialize to basis results of combinatorial representation theory. For example, by considering the open double Bruhat cell in the basic affine space we obtain a canonical basis of each irreducible representation of , parameterized by a set which each choice of seed identifies with the integral points of a lattice polytope. These bases and polytopes are all constructed essentially without representation-theoretic considerations.

Along the way, our methods prove a number of conjectures in cluster theory, including positivity of the Laurent phenomenon for cluster algebras of geometric type.

Introduction

0.1. Statement of the main results

Fock and Goncharov conjectured that the algebra of functions on a cluster variety has a canonical vector space basis parameterized by the tropical points of the mirror cluster variety. Unfortunately, as shown in Reference GHK13 by the first three authors of this paper, this conjecture is usually false: in general the cluster variety may have far too few global functions. One can only expect a power series version of the conjecture, holding in the “large complex structure limit”, and honest global functions parameterized by a subset of the mirror tropical points. For the conjecture to hold as stated, one needs further affineness assumptions. Here we apply methods developed in the study of mirror symmetry, in particular scattering diagrams, introduced by Kontsevich and Soibelman in Reference KS06 for two dimensions and by Gross and Siebert in Reference GS11 for all dimensions, broken lines, introduced by Gross in Reference G09 and developed further by Carl, Pumperla, and Siebert in Reference CPS, and theta functions, introduced by Gross, Hacking, Keel, and Siebert, see Reference GHK11, Reference CPS, Reference GS12, and Reference GHKS, to prove the conjecture in this corrected form. We give in addition a formula for the structure constants in this basis, nonnegative integers given by counts of broken lines. Definitions of all these objects, essentially combinatorial in nature, in the context of cluster algebras will be given in later sections. Here are more precise statements of our results.

For basic cluster variety notions we follow the notation of Reference GHK13, §2, for convenience, as we have collected there a number of definitions across the literature; nothing there is original. We recall some of this notation in Appendices A and B. The various flavors of cluster varieties are all varieties of the form , where is a copy of the algebraic torus

over a field of characteristic , and is a lattice, indexed by running over a set of seeds (a seed being roughly an ordered basis for ). The birational transformations induced by the inclusions of two different copies of the torus are compositions of mutations. Fock and Goncharov introduced a simple way to dualize the mutations, and using this define the Fock–Goncharov dual,⁠Footnote1 . We write for the tropical semifield of integers under . There is a notion of the set of -valued points of , written as . This can also be viewed as being canonically in bijection with , the set of divisorial discrete valuations on the field of rational functions of where the canonical volume form has a pole; see §2. Each choice of seed determines an identification .

1

Roughly one can view the Fock–Goncharov dual as the mirror variety, but this is not always precisely the case. With some additional effort, one can make this precise “at the boundary”, but we shall not do so here.

Our main object of study is the cluster variety with principal coefficients, ; see Appendices A and B for notation. This comes with a canonical fibration over a torus and a canonical free action by a torus . We let . The fiber ( the identity) is the Fock–Goncharov variety (whose algebra of regular functions is the Fomin–Zelevinsky upper cluster algebra). The quotient is the Fock–Goncharov variety.

Definition 0.1.

A global monomial on a cluster variety is a regular function on which restricts to a character on some torus in the atlas. For an -type cluster variety, a global monomial is the same as a cluster monomial. One defines the upper cluster algebra associated to by and the ordinary cluster algebra to be the subalgebra of generated by global monomials.

For example, is the original cluster algebra defined by Fomin and Zelevinsky in Reference FZ02a, and is the corresponding upper cluster algebra as defined in Reference BFZ05.

Given a global monomial on , there is a seed such that is a character , . Because the seed gives an identification of with , we obtain an element , which we show is well-defined (independent of the open set ); see Lemma 7.10. This is the -vector of the global monomial . We show this notion of -vector coincides with the notion of -vector from Reference FZ07 in the case; see Corollary 5.9. Let be the set of -vectors of all global monomials on . Finally, we write for the -vector space with basis , i.e.,

(where for the moment indicates the abstract basis element corresponding to ).

Fock and Goncharov’s dual basis conjecture says that is canonically identified with the vector space , and so in particular should have a canonical -algebra structure. Note that such an algebra structure is determined by its structure constants, a function

such that for fixed , for all but finitely many and

With this in mind, we have:

Theorem 0.3.

Let be one of . The following hold:

(1)

There are canonically defined nonnegative structure constants

These are given by counts of broken lines, certain combinatorial objects which we will define. The value is not taken in the or case.

(2)

There is a canonically defined subset with such that the restriction of gives the vector subspace with basis indexed by the structure of an associative commutative -algebra.

(3)

, i.e., contains the -vector of each global monomial.

(4)

For the lattice structure on determined by any choice of seed, is closed under addition. Furthermore, is saturated: for and , if and only if .

(5)

There is a canonical -algebra map which sends for to the corresponding global monomial.

(6)

The image is a universal positive Laurent polynomial (i.e., a Laurent polynomial with nonnegative integral coefficients in the cluster variables for each seed

(7)

is injective for or . Furthermore, is injective for under the additional assumption that there is a seed for which all the covectors , , lie in a strictly convex cone. When is injective, we have canonical inclusions

There is an analogue to Theorem 0.3 for (the main difference is that the theta functions, i.e., the canonical basis for , are only defined up to scaling each individual element, and the structure constants will not in general be integers). Injectivity in (7) holds for very general ; see Theorem 7.16.

Note that (5)–(6) immediately imply:

Corollary 0.4 (Positivity of the Laurent phenomenon).

Each cluster variable of an -cluster algebra is a Laurent polynomial with nonnegative integer coefficients in the cluster variables of any given seed.

This was conjectured by Fomin and Zelevinsky in their original paper Reference FZ02a. Positivity was obtained independently in the skew-symmetric case by [LS13] by an entirely different argument. In our proof the positivity in (1) and (6) both come from positivity in the scattering diagram, a powerful tool fundamental to the entire paper; see Theorem 1.13.

We conjecture that injectivity in (7) holds for all (without the convexity assumption). Note (7) includes the linear independence of cluster monomials, which has already been established (without convexity assumptions) for skew-symmetric cluster algebras in Reference CKLP by a very different argument. The linear independence of cluster monomials in the principal case also follows easily from our scattering diagram technology, as pointed out to us by Greg Muller; see Theorem 7.20.

When there are frozen variables, one obtains a partial compactification (where the frozen variables are allowed to take the value ) for , or . The notions of , , , and extend naturally to ; see Construction B.9.

Of course if , and we have injectivity in (7), has a canonical basis with the given properties. Also, implies, under certain hypotheses, ; see Lemma 9.10. Such partial compactifications are essential for representation-theoretic applications:

Example 0.5.

Let . Choose a Borel subgroup of , a maximal torus, and let be the unipotent radical of . These choices determine a cluster variety structure (with frozen variables) on , with , the ring of regular functions on ; see Reference GLS, §10.4.2.

Theorem 0.3 implies that these choices canonically determine a vector space basis . Each basis element is an -eigenfunction for the natural (right) action of on . For each character , is a basis of the weight space . The are the collection of irreducible representations of , each of which thus inherits a basis, canonically determined by the choice of .

We give, combining our results with results of T. Magee, much more precise results; see Corollary 0.20.

Canonical bases for have been constructed by Lusztig. Here we will obtain bases by a procedure very different from Lusztig’s, as a special case of the more general Reference GHK11, Conjecture 0.6, which applies in theory to any variety with the right sort of volume form. See Remark 0.16 for further commentary on this.

The tools necessary for the proof of Theorem 0.3 are developed in the first six sections of the paper, with the proof given in §7. This material is summarized in more detail in §0.2.

The second part of the paper turns to criteria for the full Fock–Goncharov conjecture to hold. Precisely:

Definition 0.6.

We say the full Fock–Goncharov conjecture holds for a cluster variety if the map of Theorem 0.3 is injective,

Note this implies .

We prove a number of criteria which guarantee the full Fock–Goncharov conjecture holds. One such condition, which seems to be very natural in our setup and is implied, say, by the existence of a maximal green sequence, is:

Proposition 0.7 (Proposition 8.25).

If the set of all -vectors of global monomials of in is not contained in a half-space under the identification of with induced by some choice of seed, then the full Fock–Goncharov conjecture holds for , , very general and, if the convexity condition of Theorem 0.3 holds, for .

Many of the results in the second part of the paper are proved using a generalized notion of convex function or convex polytope; see §§0.3 and 0.4 for more details.

In §8.5, we turn to results on partial compactifications. We first explain how convex polytopes in our sense give rise, under suitable hypotheses, to compactifications of -type cluster varieties and toric degenerations of such. This connects our constructions to the mirror symmetry picture described in Reference GHK11, and in particular describes a partial compactification of as giving a degeneration of a family of log Calabi–Yau varieties to a toric variety. Partial compactifications via frozen variables are also important in representation theoretic applications, as already indicated in Example 0.5. We prove results for such partial compactifications which, combined with recent results of T. Magee Reference Ma15, Reference Ma17, yield strong representation-theoretic results; see §0.4 for more details.

We now turn to a more detailed summary of the contents of the paper.

0.2. Toward the main theorem

Section 1 is devoted to the construction of the fundamental tool of the paper, scattering diagrams. While Reference GS11 defined these in much greater generality, here they are collections of walls living in a vector space with attached functions constructed canonically from a choice of seed data. A precise definition can be found in §1.1. Here we simply highlight the main new result, Theorem 1.13, whose proof, being fairly technical, is deferred to Appendix C. This says that the functions attached to walls of a scattering diagram associated to seed data have positive coefficients. All positivity results in this paper flow from this fundamental observation, and indeed many of our arguments use this in an essential way. For the reader’s convenience, we give in §1.2 an elementary construction of the relevant scattering diagrams, drawing on the method given in Reference KS13. Since a scattering diagram depends on a choice of seed, §1.3 shows how scattering diagrams associated to mutation equivalent seeds are related. This shows that a scattering diagram has a chamber structure indexed by seeds mutation equivalent to the initial choice of seed.

In §2 we review some notions of tropicalizations of cluster varieties, showing that scattering diagrams naturally live in such tropicalizations. Indeed, the scattering diagram which is associated to a cluster variety lives naturally in the tropical space of the Fock–Goncharov dual . These tropicalizations, crucially, can only be viewed as piecewise linear, rather than linear, spaces, with a choice of seed giving an identification of the tropicalization with a linear space. Already the mutation combinatorics becomes apparent:

Theorem 0.8 (Lemma 2.10 and Theorem 2.13).

For each seed of an -cluster variety, the (Fock–Goncharov) cluster chamber associated to is

where denotes the tropicalization of the monomial see §2. The collection of such subsets of over all mutation equivalent seeds form the maximal cones of a simplicial fan, the (Fock–Goncharov) cluster complex. The Fomin–Zelevinsky exchange graph is the dual graph of this fan.

The collection of cones was introduced by Fock and Goncharov, who conjectured they formed a fan. It is not at all obvious from the definition that the interiors of the cones cannot overlap. Our description of the chamber structure induced by a scattering diagram in fact shows that part of the chamber structure coincides with the collection of cones . This shows the fact that they form a fan directly. In addition, the set of Theorem 0.3 consists of the integral points of the union of cones in .

Section 3 gives the definition of broken line, the second principal combinatorial tool of the paper. These were originally introduced in Reference G09 and developed further in Reference CPS as tropical replacements for Maslov index two disks. In Reference GHK11, they were used to define theta functions, which are, in principle, formal sums over all broken lines with fixed boundary conditions. The relevance of theta functions for us comes in §4. Here we show the direct relationship between scattering diagrams and the cluster algebra. We show that if we associate a suitable torus to each chamber of the scattering diagram associated to a mutation of the initial seed, then the walls separating the chambers can be interpreted as giving birational maps between these tori. Gluing together these copies of gives the cluster variety; see Theorem 4.4. Further, a theta function depends on a point . If for a given choice of , is in fact a finite sum, then is a global function on . We show that this holds in particular when lies in the cluster complex , and in this case agrees with the cluster monomial with -vector given by . Because of the positivity result Theorem 1.13, is in any event always a power series with positive coefficients. Thus we get positivity of the Laurent phenomenon, Theorem 4.10, as an easy consequence of our formalism.

In §5 we begin with what is another essential observation for our approach. A choice of initial seed provides a partial compactification of by allowing the variables (the principal coefficients) to be zero. These variables induce a flat map , with being the fiber over . Our methods easily show:

Theorem 0.9 (Corollary 5.3(1)).

The central fiber is the algebraic torus .

Though immediate from our scattering diagram methods, the result is not obvious from the original definitions; indeed, it is equivalent to the sign-coherence of -vectors (see Corollary 5.5).

The last major ingredient in the proof of Theorem 0.3 is a formal version of the Fock–Goncharov conjecture. As mentioned above, this conjecture does not hold in general, but in §6, we show that the Fock–Goncharov conjecture holds in a formal neighborhood of the torus fiber of . We show the structure constants given in Theorem 0.3(1) have a tropical interpretation and determine an associative product on , except that will in general be an infinite sum of theta functions. Further, canonically associated to each universal Laurent polynomial is a formal power series which converges to in a formal neighborhood of the central fiber. For the precise statement see Theorem 6.8, which we interpret as saying that the Fock–Goncharov dual basis conjecture always holds in the large complex structure limit. This is all one should expect from log Calabi–Yau mirror symmetry in the absence of further affineness assumptions. A crucial point, shown in the proof of Theorem 6.8, is that the expansion of is independent of the choice of seed determining the compactification ; i.e., it is independent of which degeneration is used to perform the expansion.

In §7 we introduce the middle cluster algebra . The idea is that while we do not know that every regular function on can be written as a linear combination of theta functions, there is a set indexing those for which is a regular function on . These in fact yield a vector space basis for a subalgebra of which necessarily includes all cluster monomials, hence includes the ordinary cluster algebra. With this in hand, Theorem 0.3 becomes a summary of the results proved up to this point. We then deduce the result for and -type cluster varieties from the case.

0.3. Convexity conditions

We now turn to the use of convexity conditions to prove the Fock–Goncharov conjecture in a number of different situations, as covered in §8. To motivate the concepts, let us define a partial minimal model of a log Calabi–Yau variety . This is an inclusion as an open subset such that the canonical volume form on has a simple pole along each irreducible divisor of the boundary . For example, a partial minimal model for an algebraic torus is the same as a toric compactification. We wish to extend elementary constructions of toric geometry to the cluster case. For example, the partial compactification determined by frozen variables is a partial minimal model.

The generalization of the cocharacter lattice of the algebraic torus is the tropical set of . The main difference between the torus and the general case is that is not in general a vector space. Indeed, the identification of with the cocharacter lattices of various charts of induce piecewise linear (but not linear) identifications between the cocharacter lattices. As a result, a piecewise straight path in which is straight under one identification will be bent under another. Thus the usual notions of straight lines, convex functions, or convex sets do not make sense on .

The idea for generalizing the notion of convexity is to instead make use of broken lines, which are piecewise linear paths in . Using broken lines in place of straight lines, we can say which piecewise linear functions, and thus which polytopes, are convex; see Definition 8.2. Each regular function has a canonical piecewise linear tropicalization , which we conjecture is convex in the sense of Definition 8.2; see Conjecture 8.11. The conjecture is easy for ; see Proposition 8.13. Each convex piecewise linear gives a convex polytope and a convex cone , where italics indicates convexity in our broken line sense. We believe the existence of a bounded polytope is equivalent to the full Fock–Goncharov conjecture:

Conjecture 0.10.

The full Fock–Goncharov conjecture holds for if and only if the tropical space contains a full-dimensional bounded polytope, convex in our sense.

The examples of Reference GHK13, §7, show that for the full Fock–Goncharov conjecture to hold, we need to assume has enough global functions. In that case tropicalizing a general function gives (conjecturally) a bounded convex polytope. As we are unable to prove Conjecture 8.11 except in the monomial case, we use a restricted version (which happily still has wide application):

Definition 0.11.

A cluster variety has Enough Global Monomials (EGM) if for each valuation there is a global monomial with .

The condition that has EGM is equivalent to the existence of whose associated convex polytope is bounded; see Lemma 8.15.

The following theorem demonstrates the value of the EGM condition:

Theorem 0.12.

Let be a cluster variety. Then:

(1)

(Corollaries 8.18 and 8.21) If satisfies the EGM condition, then the multiplication rule on is polynomial, i.e., for given , for all but finitely many . This gives the structure of a finitely generated commutative associative -algebra.

(2)

(Proposition 8.22) If and satisfies the EGM condition, then there are canonical inclusions

Remark 0.13.

We believe, based on calculations in Reference M13, §7.1, that the conditions of the theorem ( has EGM, and ) hold for the cluster variety associated with the once-punctured torus; see some details in Examples 2.14 and 7.18. However, the equality is expected to fail, and in particular in this case we expect the full Fock–Goncharov conjecture holds for and very general but not for .

We note that has EGM in many cases:

Proposition 0.14.

Consider the following conditions on a cluster algebra :

(1)

The exchange matrix has full rank, is generated by finitely many cluster variables, and is a smooth affine variety.

(2)

has an acyclic seed.

(3)

has a seed with a maximal green sequence.

(4)

For some seed, the cluster complex is not contained in a half-space.

(5)

has EGM.

Then implies Proposition 8.27). Furthermore, implies implies implies Propositions 8.24 and 8.25 Finally, implies the full Fock–Goncharov conjecture, for , , or very general , or, under the convexity assumption of Theorem 0.3, for (Proposition 8.25

Example 0.15.

A recent paper Reference GY13 of Goodearl and Yakimov announces the equality for all double Bruhat cells in semisimple groups. In this case, Yakimov has furthermore announced the existence of a maximal green sequence. Many cluster varieties associated to a marked bordered surface with at least two punctures also have a maximal green sequence; see Reference CLS, §1.3 for a summary of known results on this. The recent Reference GS16, Theorems 1.12 and 1.17, shows that (4) holds for the Fock–Goncharov cluster varieties of local systems on most decorated surfaces. Together with Proposition 0.14, these results imply the full Fock–Goncharov theorem in any of these cases.

We note that for the cluster algebra associated to a marked bordered surface, a canonical basis of parameterized by has been previously obtained by Fock–Goncharov Reference FG06, Theorem 12.3. They show that the and varieties have natural modular meaning as moduli spaces of local systems. They identify with a space of integer laminations (isotopy classes of disjoint loops with integer weights) and their associated basis element is a natural function given by trace of monodromy around a loop. We checked, together with A. Neitzke, that our basis agrees with the Fock–Goncharov basis of trace functions in the case of a sphere with four punctures, for primitive elements of the tropical set. Our theta function basis comes canonically from the cluster structure (it does not depend on any modular interpretation).

Remark 0.16.

In general, we conjecture the bases we construct for rings of global functions on cluster varieties or partial compactifications are intrinsic to the underlying log Calabi–Yau variety and do not depend on the particular cluster structure on . This is a nontrivial statement: there exist varieties with multiple cluster structures (in particular different atlases of tori for the same variety). Yan Zhou will show in her PhD thesis that the (principal coefficient version of) the cluster variety associated to the once-punctured torus is an example.

This conjecture is suggested by Reference GHK11, Conjecture 0.6, and the results of Reference GHK11, Reference GHK12, and Reference GHKII prove this in the case of the cluster varieties where the skew-symmetric form has rank 2, which includes the case of the sphere with four punctures. Thus we have the (at least to us) remarkable conclusion that in many cases where bases occur because of some extrinsic interpretation of the spaces, in fact this extrinsic interpretation is irrelevant. For example, the theta functions given by trace functions above, which would appear to depend on the realization of the cluster variety as a moduli space of local systems, are actually intrinsic to the underlying variety. In the case of Example 0.5, where bases may arise from representation theory, our basis does not use the group-theoretic aspects of the spaces. The suggestion that the canonical basis is independent of the cluster structure may surprise some, as understanding the canonical basis was the initial motivation for the Fomin–Zelevinsky definition of cluster algebras.

Returning to the role of convexity notions, we note that our formula for the structure constants of Theorem 0.3(1) is given by counting broken lines. As a result, our notion of convexity interacts nicely with the multiplication rule. This allows us to generalize basic polyhedral constructions from toric geometry in a straightforward way.

A polytope convex in our sense determines (by familiar Rees-type constructions for graded rings) a compactification of . Furthermore, for any choice of seed, is identified with a linear space and with an ordinary convex polytope. Our construction also gives a flat degeneration of this compactification of to the ordinary polarized toric variety for ; see §8.5. We expect this specializes to a uniform construction of many degenerations of representation theoretic objects to toric varieties; see, e.g., Reference C02, Reference AB, and Reference KM05. Applied to the Fock–Goncharov moduli spaces of -local systems, this will give for the first time compactifications of character varieties with nice (e.g., toroidal anticanonical) boundary; see Remark 8.34. The polytope can be chosen so that the boundary of the compactification is very simple, a union of toric varieties. For example, let be the open subset where the frozen variables for the standard cluster structure are nonvanishing. Then the boundary consists of a union of certain Schubert cells. Using a polytope, we obtain an alternative compactification where the Schubert cells (which are highly nontoric) are replaced by toric varieties; see Theorem 8.35.

The Fock–Goncharov conjecture is the cluster special case of Reference GHK11, Conjecture 0.6, which says (roughly) that affine log Calabi–Yau varieties with maximal boundary come in canonical dual pairs with the tropical set of one parameterizing a canonical basis of functions on the other. We can view the conjecture as having two parts: First, the vector space, , with this basis is naturally an algebra in a such a way that is an affine log CY. And then furthermore, this log CY is the mirror—in the cluster case the Fock–Goncharov dual (it is natural to further ask if this is the mirror in the sense of homological mirror symmetry but we do not consider this question here). Our deepest mirror theoretic result is the following weakening of the first part:

Theorem 0.17.

Assume has EGM. Let , , or for very general , or let and assume the convexity condition of Theorem 0.3 holds. Then the structure constants of Theorem 0.3 define an algebra structure on such that it is a finitely generated -algebra and is a log canonical Gorenstein -trivial affine variety of dimension .

For the proof see Theorem 8.32.

0.4. Representation-theoretic applications

We turn to §9. Here we study features of partial compactifications coming from frozen variables. As explained in Example 0.5, these partial compactifications are often the relevant ones in representation-theoretic examples. In particular, for a partial minimal model , often the vector subspace is more important than itself. For example there is a cluster structure with frozen variables for the open double Bruhat cell in a semisimple group . Then is the ring of functions on the open double Bruhat cell and . Of course is the most important representation of . However, one cannot expect a canonical basis of , i.e., one determined by the intrinsic geometry of . For example, has no nonconstant global functions which are eigenfunctions for the action of on itself. But we expect, and in the myriad cases above can prove, that the affine log Calabi–Yau open subset has a canonical basis , and we believe that , the set of theta functions on that extends regularly to all of , is a basis for , canonically associated to the choice of log Calabi–Yau open subset ; see Reference GHK13, Remark 1.10. This is not a basis of -eigenfunctions, but they are eigenfunctions for the associated maximal torus, which is the subgroup of that preserves . This is exactly what one should expect: the basis is not intrinsic to , instead it is (we conjecture) intrinsic to the pair ; see Remark 0.16.

We shall now describe in more detail what can be proved for partial compactifications of cluster varieties coming from frozen variables. A key point is a technical but combinatorial hypothesis that each variable has an optimized seed; see Definition 9.1 and Lemmas 9.2 and 9.3. The main need for this hypothesis is Proposition 9.7, which states that if a linear combination of theta functions extends across a boundary divisor, then each theta function in the sum extends across the divisor. Thus the middle cluster algebra, in this case, behaves well with respect to boundary divisors. Happily, this condition holds for the cluster structures on the Grassmannian, and, for , for the cluster structure on a maximal unipotent subgroup , the basic affine space , and the Fock–Goncharov cluster structure on ; see Remark 9.5.

Let us now work with the principal cluster variety . Consider the partial compactification by allowing the frozen variables to be zero. Each boundary divisor gives a point and thus (in general conjecturally) a canonical theta function on . We then define the potential as the sum of these theta functions. We have its piecewise linear tropicalization . This defines a cone

Theorem 0.19 (Corollaries 9.17 and 9.18).

Assume that each frozen index has an optimized seed. Then:

(1)

and are convex in our sense.

(2)

The set parameterizes a canonical basis of an algebra , and

(3)

Now assume further that we have EGM on . If for some seed , is contained in the convex hull of (which itself contains the convex hull of then , is finitely generated, and the integer points parameterize a canonical basis.

Each choice of seed identifies with a lattice and the cone with a rational polyhedral cone, described by canonical linear inequalities given by the tropicalization of the potential. Note that is convex in our generalized sense.

We show, making use of recent results of Magee Reference Ma15, Reference Ma17 and Goncharov and Shen Reference GS16 that in the representation-theoretic examples, which were the original motivation for the definition of cluster algebras, our polyhedral cones specialize to the piecewise linear parameterizations of canonical bases of Berenstein and Zelevinsky Reference BZ01, Knutson and Tao Reference KT99, and Goncharov and Shen Reference GS13:

Corollary 0.20.

Let and let be the Fomin–Zelevinsky cluster variety for the basic affine space .

(1)

All the hypotheses, and thus the conclusions, of Theorem 0.19 hold. In particular parameterizes a canonical theta function basis of .

(2)

Our potential agrees with the (representation theoretically defined) potential function of Berenstein and Kazhdan Reference BK07.

(3)

The maximal torus acts canonically on , preserving the open set .

(4)

Each theta function is an -eigenfunction, and there is a canonical map

(the target is the character lattice of linear for the linear structure given by any seed, which sends an integer point to the -weight of the corresponding theta function. The slice

parameterizes a canonical theta function basis of the eigenspace , the corresponding irreducible representation of .

(5)

For a natural choice of seed, the cone is canonically identified with the Gelfand–Tsetlin cone.

Corollary 0.21.

Let be the Fock–Goncharov cluster variety for

(1)

All the hypotheses, and thus the conclusions, of Theorem 0.19 hold. In particular the cone parameterizes a canonical theta function basis of .

(2)

Our potential function agrees with the (representation theoretically defined) potential function of Goncharov and Shen Reference GS13.

(3)

acts canonically on , preserving the open subset .

(4)

Each theta function is an -eigenfunction, and there is a canonical map

linear for the linear structure given by any seed, which sends an integer point to the -weight of the corresponding theta function. The slice

parameterizes a canonical theta function basis of the eigenspace

In particular, the number of integral points in is the corresponding Littlewood–Richardson coefficient.

(5)

For a natural choice of seed, the cone is canonically identified with the Knutson–Tao hive cone.

These corollaries are proven at the end of §9.2.

We stress here that the above representation-theoretic results come for free from general properties of our mirror symmetry construction: any partial minimal model of an affine log Calabi–Yau variety with maximal boundary determines (in general conjecturally) a cone with the analogous meaning. We are getting these basic representation-theoretic results without representation theory!

We recover the remarkable Gelfand–Tsetlin and hive polytopes for a particular choice of seed. Different (among the infinitely many possible) choices of seed give in general combinatorially different cones, whose integer points parameterize the same theta function basis. The canonical object is the convex cone cut out by , different (by piecewise linear mutation) identifications of with a vector space give different incarnations of as convex cones in the usual sense.

Potentials were considered in the work of Goncharov and Shen Reference GS13, which in turn built on work of Berenstein and Zelevinsky Reference BZ01 and Berenstein and Kazhdan Reference BK00, Reference BK07. The potential constructed by Goncharov and Shen has a beautiful representation-theoretic definition and was found in many situations to coincide with known constructions of Landau–Ginzburg potentials. On the other hand, the construction of the potential in terms of theta functions coincides precisely with the construction of the mirror Landau–Ginzburg potential as carried out in Reference G09, Reference CPS. The latter work can be viewed as a tropicalization of the descriptions of the potential in terms of holomorphic disks in Reference CO06, Reference A07. Thus our construction explains the emergence of the Landau–Ginzburg potentials in Reference GS13. Our potentials are determined by the cluster structure (and conjecturally, just the underlying log Calabi–Yau variety), and in particular are independent of any modular or representation-theoretic interpretation of the cluster variety. This gives, as in Remark 0.16, the remarkable suggestion that, e.g., the representation theoretically defined Goncharov–Shen potential, which would seem to depend heavily on the modular interpretation of , is actually intrinsic to the partial minimal model .

1. Scattering diagrams and chamber structures

1.1. Definition and constructions

Here we recall the basic properties of scattering diagrams, the main technical tool in this paper. Scattering diagrams appeared first in Reference KS06 in two dimensions, and then in all dimensions in Reference GS11, with another approach in a more specific case in Reference KS13. Here we give a self-contained treatment restricted to the specific case needed in this paper.

We start with a choice of fixed data as defined in Reference GHK13, which for the reader’s convenience is described at the beginning of Appendix A. In brief, this entails a lattice with dual lattice , a skew-symmetric form

sublattices with a saturated sublattice and a sublattice of finite index with dual lattice , an index set with and a subset with , as well as positive integers , . Finally, we also choose an initial seed , i.e., a basis of . See Appendix A for the precise properties that all this data must satisfy.

For the construction of the scattering diagram associated to this data, we will require

The injectivity assumption

The map given by is injective.

While this does not hold for a general choice of fixed data, it does hold in the principal coefficient case (see Appendix B) and results in this paper about arbitrary cluster varieties and algebras will be proved via the principal case.

Set

Choose a linear function such for .

Under the injectivity assumption, one can choose a strictly convex top-dimensional cone , with associated monoid , such that for all . Here is the group of units of the monoid . This gives the monomial ideal in the monoid ring over a field of characteristic , and we write for the completion with respect to .

We define the module of log derivations of as

with the action of on being given by

so we write as . Let denote the completion of with respect to the ideal .

Using this action, if , then

makes sense using the Taylor series for the exponential. We have the Lie bracket

Then can be viewed as a subgroup of the group of continuous automorphisms of which are the identity modulo , with the group law of composition coinciding with the group law coming from the Baker–Campbell–Hausdorff formula.

Define the sub-Lie algebra of

where is the one-dimensional subspace of spanned by . We calculate that is in fact closed under Lie bracket:

We have

a Lie subalgebra, and a nilpotent Lie algebra. We let be the corresponding nilpotent group. This group, as a set, is just , but multiplication is given by the Baker–Campbell–Hausdorff formula. We set

the corresponding pronilpotent group. We have the canonical set bijections

For we define

Note that by the commutator formula Equation 1.1, , hence , is abelian.

In what follows, noting that is a subgroup of , we will often describe elements of as follows.

Definition 1.2.

Let , and . Define to be the automorphism of given by

where is the generator of the monoid .

Lemma 1.3.

For , is the subgroup of automorphisms of the form for as in Definition 1.2 with the given . More specifically, acts as the automorphism with , where is the smallest positive rational number such that .

Proof.

Let be the set of of the given form. Then is a subgroup as . Note that , where is as described in the statement. The exponential of this vector field is easily seen to act as with . Hence . From this, we see also that if , then , and the latter lies in .

Definition 1.4.

A wall in (for and ) is a pair such that

(1)

for some primitive .;

(2)

is a -dimensional convex (but not necessarily strictly convex) rational polyhedral cone.

The set is called the support of the wall .

Remark 1.5.

Using Lemma 1.3, we often write a wall as for , necessarily of the form . We shall use this notation interchangeably without comment.

Definition 1.6.

A scattering diagram for and is a set of walls such that for every degree , there are only a finite number of with the image of in not the identity.

If is a scattering diagram, we write

for the support and singular locus of the scattering diagram. If is a finite scattering diagram, then its support is a finite polyhedral cone complex. A joint is an -dimensional cell of this complex, so that is the union of all joints of .

Remark 1.7.

We will often (especially in Appendix C) want to use a slightly more general notion of scattering diagram, where the elements attached to walls lie in some other choice of group arising from an -graded Lie algebra . In this case we talk about a scattering diagram for . For example, any scattering diagram for induces a finite scattering diagram for by taking the image of the attached group elements under the projection .

Given a scattering diagram , we obtain the path-ordered product. Assume given a smooth immersion

with endpoints not contained in the support of . Assume is transversal to each wall of that it crosses. For each degree , we can find numbers

and elements with the image of in nontrivial such that

if , and taken as large as possible. (The are the times at which the path hits a wall. We allow because we may have two different walls which span the same hyperplane.)

For each , define

where with . We then define

If , then span the same hyperplane , hence . Thus, since this latter group is abelian, and commute, so this product is well-defined. We then take

We note that depends only on its homotopy class (with fixed endpoints) in . We also note that the definition can easily be extended to piecewise smooth paths , provided that the path always crosses a wall if it intersects it.

Definition 1.8.

Two scattering diagrams , are equivalent if for all paths for which both are defined.

Call general if there is at most one rational hyperplane with . For general and a scattering diagram, let . One checks easily:

Lemma 1.9.

Two scattering diagrams are equivalent if and only if for all general .

Definition 1.10.

A scattering diagram is consistent if only depends on the endpoints of for any path for which is defined.

Definition 1.11.

We say a wall is incoming if

Otherwise, we say the wall is outgoing (note in any case lies in the span of the wall ).

We call the direction of the wall. (This terminology comes from the case , where an outgoing wall is then a ray containing its direction vector, thus one that points outward.)

We need one particular scattering diagram, determined by the fixed data and seed data. Setting , , we start with the scattering diagram

The main result on scattering diagrams, which follows easily from Theorem 1.21, is the following. A more general version of this was proved in two dimensions in Reference KS06 and in a much more general context in all dimensions in Reference GS11. A simpler argument which applies to the case at hand was given in Reference KS13, which shall be reviewed in §1.2 .

Theorem 1.12.

There is a scattering diagram satisfying:

(1)

is consistent,

(2)

,

(3)

consists only of outgoing walls.

Moreover, satisfying these three properties is unique up to equivalence.

The crucial positivity result satisfied by is now easily stated:

Theorem 1.13.

The scattering diagram is equivalent to a scattering diagram all of whose walls satisfy for some and a positive integer. In particular, all nonzero coefficients of are positive integers.

The proof is given in Appendix C. The basic idea is that the construction of the scattering diagram can be reduced to repeated applications of the following example:

Example 1.14.

Take , and the skew-symmetric form given by the matrix , where . Let be the dual basis of , and write , . We get

Then one checks easily that

See Figure 1.1. (See for example Reference GPS, Example 1.6.)

Example 1.15.

Take , with basis , and take to be the sublattice generated by . Further, take , , where are two positive integers, and take the skew-symmetric form to be the same as in the previous example. Then , . Taking as before , , we get

For most choices of and , this is a very complicated scattering diagram. A very similar scattering diagram, with functions and , has been analyzed in Reference GP10, but it is easy to translate this latter diagram to the one considered here by replacing by and using the change of lattice trick, which is given in Step IV of the proof of Proposition C.13. All rays of are contained strictly in the fourth quadrant (i.e., in particular are not contained in an axis). Without giving the details, we summarize the results. There are two linear operators given by the matrices in the basis as

Then is invariant under and , in the sense that if , we have provided is contained strictly in the fourth quadrant. It is also the case that applying to or to gives an element of . Further, contains a discrete series of rays consisting of those rays in the fourth quadrant obtained by applying and alternately to the above rays supported on and . These rays necessarily have functions of the form or for various choices of and . If , we obtain a finite diagram. (Moreover, the corresponding cluster variety is the cluster variety of finite type Reference FZ03a associated to the root system , , or for , , or respectively.) See Figure 1.2 for the case . If , these rays converge to the rays contained in the two eigenspaces of and . These are rays of slope . This gives a complete description of the rays outside of the cone spanned by these two rays. The expectation is that every ray of rational slope appears in the interior of this cone, and the attached functions are in general unknown (see Figure 1.3). However, in the case, it is known Reference R12 that the function attached to the ray of slope is

The chamber structure one sees outside the quadratic irrational cone is very well-behaved and familiar in cluster algebra theory. In particular, the interiors of the first, second, and third quadrants are all connected components of , and there are for an infinite number of connected components in the fourth quadrant. We will see in §2 that this chamber structure is precisely the Fock–Goncharov cluster complex.

On the other hand, it is precisely the rich structure inside the quadratic irrational cone which scattering diagram technology brings into the cluster algebra picture.

1.2. Construction of consistent scattering diagrams

In this subsection we give more details about the construction of scattering diagrams, and in particular give results leading to the proof of Theorem 1.12. This material can be skipped on first reading but is recommended before reading the more difficult material on scattering diagrams in Appendix C.

Let be a scattering diagram. If we set

then since any wall spans a hyperplane , for some ,

In particular, if is a consistent scattering diagram, then for a path with initial point in and final point in is independent of the particular choice of path (or endpoints in ). Thus we obtain a well-defined element which only depends on the scattering diagram .

Theorem 1.17 (Kontsevich and Soibelman).

The assignment of to gives a one-to-one correspondence between equivalence classes of consistent scattering diagrams and elements .

This is a special case of Reference KS13, 2.1.6. For the reader’s convenience we include the short proof:

Proof.

We need to show how to construct given . To do so, choose any primitive and a point general. Then we can determine as follows, noting by Lemma 1.9 that this information for all such and general determines up to equivalence. We can write

with

Each of these subspaces of are closed under Lie bracket, thus defining subgroups of . Note by the generality assumption on , we in fact have . This splitting induces a unique factorization for any element . Applying this to gives a well-defined element . We need to show that the set of data determines a scattering diagram such that for all general . To do this, one needs to know that to any finite order , the hyperplane is subdivided into a finite number of polyhedral cones such that the image of in is constant for . This is clear because the number of with is finite, as then the decomposition Equation 1.18 varies discretely with to order .

We need to show that satisfies the condition that for any path from the positive to the negative chamber and that only depends on endpoints of . To do so, we work modulo for any , so we can assume has a finite number of walls. Choose a general point . Take a general two-dimensional subspace of containing , and after choosing a metric, let be a semicircle in the two-dimensional subspace with endpoints and and center . Then for points contained in walls crossed by , and is the element of determined by the factorization of above. Note that if lies in the hyperplane , all the wall-crossing automorphisms of walls traversed by before crossing lie in and all those from walls traversed by after crossing lie in . It then follows inductively that the factorization of given by takes the form for some . Indeed, for , this just follows from the definition of , while if true for , then we have that is a decomposition of induced by the splitting , and the claim then follows by the definition of . In particular, for , taking and noting that , one sees that .

Next we show the independence of path for the we have constructed, again modulo . It is sufficient to check as an element of for any small loop around any joint of . Take a general point in , such that , and choose to be points in near on either side of the joint . Let be two semicircular paths with endpoints and and passing through respectively. Then up to orientation is freely homotopic to in . Thus .

Thus we have established the one-to-one correspondence between consistent scattering diagrams and elements of .

Following Reference KS13, we give an alternative parameterization of , as follows. For any primitive, we get the splitting

where

These give rise to subgroups of . We drop the when it is clear from context. Again, this allows us to factor any as with , . We can further decompose , where , while involves those summands of coming from not proportional to . Note that . Indeed, if with for , we then have so that by (Equation 1.1). Thus we have a projection homomorphism with kernel . In particular, the factorization yields an element via this projection. We then have a map (of sets)

Proposition 1.20.

is a set bijection.

Proof.

is induced by an analogous map to order ,

One checks easily that this is a bijection order by order.

Theorem 1.21.

Let be a consistent scattering diagram corresponding to . The following hold:

(1)

For each , to any fixed finite order, there is an open neighborhood of such that for all general . Here denotes the component of indexed by .

(2)

is equivalent to a diagram with only one wall in containing for each , and the group element attached to this wall is .

(3)

Set

Then is equivalent to a consistent scattering diagram such that and consists only of outgoing walls. Furthermore, up to equivalence, is the unique consistent scattering diagram with this property.

(4)

The equivalence class of a consistent scattering diagram is determined by its set of incoming walls.

We note first that (3) of Theorem 1.21 implies Theorem 1.12. Indeed, let for be the group element corresponding to , so that the initial scattering can be written as . By Proposition 1.20 there is a unique element with

Now apply Theorem 1.21 with .

Proof of Theorem 1.21.

First note that statement (1) implies (2). Further, (1), along with Theorem 1.17 and Proposition 1.20, implies (4), which in turn gives the uniqueness in (3). Note (1) implies that, to the given finite order, is equivalent to a diagram having only one incoming wall contained in , and the attached group element is . Now we can replace this single wall by an equivalent collection of walls consisting of and a number of outgoing walls contained in with attached group element . This gives the existence in (3).

Thus it suffices to prove (1). We work modulo , so we may assume is finite, and compare the splittings Equation 1.18 coming from a choice of near and Equation 1.19, after replacing with . For each , there exists an open neighborhood of such that (resp. ) implies (resp. ) for all . Since is now a finite sum of ’s, we can find a single so that for all . If is general inside , we also have .

Now write

as in Equation 1.19. Then we can further factor

as in Equation 1.18. Note , . Since the projection is a group homomorphism with kernel , the image of in is , which thus coincides with by definition of the latter. We have

which is then the (unique) factorisation from Equation 1.18. Thus

for any general .

1.3. Mutation invariance of the scattering diagram

We now study how the scattering diagram constructed from seed data defined in the previous subsection changes under mutation. This is crucial for uncovering the chamber structure of these diagrams and giving the connection with the exchange graph and cluster complex.

Thus let and be the mutated seed (see, e.g., Reference GHK13, (2.3)). To distinguish the two Lie algebras involved, we write and for the Lie algebras arising from these two different seeds. We recall that the injectivity assumption is independent of the choice of seed.

Definition 1.22.

We set

For , define the piecewise linear transformation by, for ,

As we will explain in §2, is the tropicalization of viewed as a birational map between tori. We will write and to be the linear transformations used to define in the regions and , respectively.

Define the scattering diagram to be the scattering diagram obtained by the following:

(1)

For each wall , where , we have one or two walls in given as

throwing out the first or second of these if or , respectively. Here for linear, we write for the formal power series obtained by applying to each exponent in .

(2)

also contains the wall .

The main result of this subsection is:

Theorem 1.24.

Suppose the injectivity assumption is satisfied. Then is a consistent scattering diagram for and . Furthermore, and are equivalent.

The main point in the proof, which is not at all obvious from the definition, is that is a scattering diagram for , , where . Formally, consistency will be easy to check using consistency of . It will follow easily that by construction and have the same incoming walls, so the theorem will then follow from the uniqueness in Theorem 1.12.

The main problem to overcome is that the functions attached to walls of and live in two different completed monoid rings, and , for a monoid chosen to contain , and a monoid chosen to contain . We need first a common monoid containing both and .

Definition 1.25.

Let be a top-dimensional cone containing , , and , and such that . Set , and .

Given such a choice of , we can find , contained in . However, we have an additional problem that is not trivial modulo . Indeed, , while one of the initial walls of is . In particular, the wall-crossing automorphism associated to

is not an automorphism of the ring , but rather of the localized ring . (Here the hats denote completion with respect to .) This kind of situation is dealt with in Reference GS11, see especially §4.3. However the current situation is quite a bit simpler, so we will give the complete necessary arguments here and in Appendix C.

We will use the notation for the automorphism of associated to crossing the wall from to . Explicitly,

In this situation, define

We note that by the definition of the mutated seed , , so we indicate it by .

We now extend the definition of scattering diagram.

Definition 1.27.

A wall for and ideal is a pair with as in Definition 1.4, but with , and congruent to mod . The slab for the seed means the pair . Note since this does not qualify as a wall. Now a scattering diagram is a collection of walls and possibly this single slab, with the condition that for each , for all but finitely many walls in .

Note that crossing a wall or slab now induces an automorphism of of the form (with the localization only needed when a slab is crossed).

The following is proved in Appendix C:

Theorem 1.28.

There exists a scattering diagram in the sense of Definition 1.27 such that

(1)

,

(2)

consisting only of outgoing walls, and

(3)

as an automorphism of only depends on the endpoints of .

Furthermore, with these properties is unique up to equivalence.

Finally, is also a scattering diagram for the data and, as such, is equivalent to .

Remark 1.29.

Note in particular that the theorem implies does not contain any walls contained in besides . Indeed, no wall of is contained in : only the slab is contained in .

Proof of Theorem 1.24.

We write , .

We first note that we can choose representatives for , , which are scattering diagrams in the sense of Definition 1.27, by Theorem 1.28. Furthermore, is also a scattering diagram in the sense of Definition 1.27 for the seed : this follows since if for some , we also have . Thus by the uniqueness statement of Theorem 1.28, and are equivalent if (1) these diagrams are equivalent to diagrams which have the same set of slabs and incoming walls; (2) is consistent. We carry out these two steps.

Step I. Up to equivalence, and have the same set of slabs and incoming walls.

If is outgoing, the wall contributes to and is also outgoing, so let us consider the incoming walls of . Setting , already contains the slab for

which lies in by construction. Next consider the wall , for . We have three cases to consider, based on whether is zero, positive, or negative.

First if , then takes the plane to itself (in a piecewise linear way), and . Thus the wall contributes two walls whose union is the wall , as and in this case. Up to equivalence, we can replace these two walls with the single wall .

If , then consider the wall

This wall contains the ray , so this is an incoming wall. Note that if , we have, with as given in Equation A.1,

Thus is a half-space contained in , and furthermore since

Thus we see that the wall of is half of the wall of .

If , then the wall coincides with , and also contains , so is an incoming wall. But also , in this case. Thus is again half of the wall .

In summary, we find that after splitting some of the walls of in two, and have the same set of incoming walls, and thus, making a similar change to , we see that and have the same set of incoming walls.

Step II. for any loop for which this automorphism is defined.

Indeed, the only place a problem can occur is for a loop around a joint of contained in the slab , as this is where fails to be linear. To test this, consider a loop around a joint contained in . Assume that it has basepoint in the half-space and is split up as , where immediately crosses , is contained entirely in , crossing all walls of which contain and intersect the interior of , crosses again, and then crosses all relevant walls in the half-space .

Let , be the wall-crossing automorphisms for crossing or passing from to , as in Equation 1.26. Then by Remark 1.29, and .

Let be the automorphism induced by , i.e.,

Then note that

Thus to show , it is enough to show that

But

as desired.

Construction 1.30 (The chamber structure).

Suppose given fixed data satisfying the injectivity assumption and seed data . We then obtain for every seed obtained from via mutation a scattering diagram . In each case we will choose a representative for the scattering diagram with minimal support.

Note by construction and Remark 1.29, irrespective of the representative of used, contains walls whose union of supports is . Furthermore, we have given by Equation 1.16, which can be written more explicitly as

Then are the closures of connected components of . Similarly, we see that taking to be the chambers where all are positive (or negative), we have that is the closure of a connected component of , so that is the closure of a connected component of . Note that the closures of and have a common codimension 1 face given by the intersection with . This gives rise to the following chamber structure for a subset of .

We refer the reader to Appendix A for the definition of the infinite oriented tree (or ) used for parameterizing seeds obtained via mutation of . In particular, for any vertex of , there is a simple path from the root vertex to , indicating a sequence of mutations and hence a piecewise linear transformation

Note that is defined using the basis vector of the seed , not the basis vector of the original seed . By applying Theorem 1.24 repeatedly, we see that

(where applied to the scattering diagram is interpreted as the composition of the actions of each ) and

is the closure of a connected component of .

Note that the map from vertices of to chambers of is never one-to-one. Indeed, if is the vertex obtained by following the edge labeled twice starting at the root vertex, one checks that , even though (see Reference GHK13, Remark 2.5).

Thus we have a chamber structure on a subset of ; in general, the union of the cones do not form a dense subset of .

Since we will often want to compare various aspects of this geometry for different seeds, we will write the short-hand for an object parameterized by a vertex where the root of the tree is labeled with the seed . In particular:

Definition 1.32.

We write for the chamber of corresponding to the vertex . We write for the set of chambers for running over all vertices of . We call elements of cluster chambers.

2. Basics on tropicalization and the Fock–Goncharov cluster complex

We now explain that the chamber structure of Construction 1.30 coincides with the Fock–Goncharov cluster complex. To do so, we first recall the basics of tropicalization.

For a lattice with , let be the subset of elements of the field of fractions of which can be expressed as a ratio of Laurent polynomials with nonnegative integer coefficients. Then is a semifield under ordinary multiplication and addition. For any semifield , restriction to the monomials gives a canonical bijection

where the first is maps of semifields, means the multiplicative group of , and in the last tensor product we mean viewed as -module. Following Reference FG09 we define the -valued points of to be . A positive birational map means a birational map for which the pullback induces an isomorphism on . Obviously, it gives an isomorphism on -valued points. Thus it makes sense to talk about for any variety with a positive atlas of tori, for example many of the various flavors of cluster variety.

There are two equally good semifield structures on , the max-plus and the min-plus structures. Here addition is either maximum or minimum, and multiplication is addition. We notate these as and , respectively, thinking of capital for the max-plus tropicalization and little for the min-plus tropicalization. We similarly define and . Thus taking or , we obtain the sets of tropical points or . The former is the convention used by Fock and Goncharov in Reference FG09, so we refer to this as the Fock–Goncharov tropicalization. The latter choice in fact coincides with as defined in Reference GHK13, Def. 1.7, defined as a subset of the set of discrete valuations. We refer to this as the geometric tropicalization. It will turn out both are useful. There is the obvious isomorphism of semifields from . This induces a canonical sign-change identification .

Given a positive birational map , we use and to indicate the induced maps and respectively. For the geometric tropicalization, this coincides with the map on discrete valuations induced by pullback of functions; see Reference GHK13, §1. For cluster varieties the two types of tropicalization are obviously equivalent. The geometric tropicalization has the advantage that it makes sense for any log Calabi–Yau variety, while the Fock–Goncharov tropicalization is restricted to (Fock–Goncharov) positive spaces, i.e., spaces obtained by gluing together algebraic tori via positive birational maps. We will use both notions: because in many cases it is more natural to think in terms of valuations/boundary divisors, and because, as we indicate below, the scattering diagram for building lives naturally in (because of already established cluster sign conventions).

One computes easily that for the basic mutation

the Fock–Goncharov tropicalization is

while the geometric tropicalization (see Reference GHK13, (1.4)) is

Thus:

Proposition 2.4.

defined in Equation 1.23 is the Fock–Goncharov tropicalization of

A rational function on a cluster variety is called positive if its restriction to each seed torus is positive, i.e., can be expressed as a ratio of sums of characters with positive integer coefficients. We can then define its Fock–Goncharov tropicalization by . Similarly, for positive, we have its geometric tropicalization which for each has value . Using the identification of with , is interpreted as a valuation and coincides with , the value of on . In particular, this value is defined regardless of whether is positive. We have a commutative diagram

where is the canonical isomorphism determined by the sign-change isomorphism. The definition of in terms of valuations extends the definition of , and hence via this diagram, to any nonzero rational function. We note that

where

is the canonical restriction isomorphism. We will almost always leave out of the notation.

Lemma 2.8.
(1)

For a positive Laurent polynomial (i.e., , and

where is the canonical isomorphism Equation 2.7.

(2)

If is a divisorial discrete valuation and is any Laurent polynomial (so now then

Proof.

By definition

This gives the first statement. For the second, we can assume is primitive, so part of a basis. Then the statement reduces to an obvious statement about the degree of a linear combination of monomials in .

Note the mutations are precisely the mutations between the tori in the atlas for (see Appendix A for the definition of the Fock–Goncharov dual , and Reference GHK13, (2.5) for the mutations between tori in our notation). Thus by Theorem 1.24 and Proposition 2.4, the support of viewed as a subset of under the identification (induced canonically from the open set ) is independent of seed. In particular it makes sense to talk about as being completely canonically defined without choosing any seed. For any seed, the chambers are connected components of .

We recall from Reference FG11:

Definition 2.9.

Suppose we are given fixed data and an initial seed. For a seed obtained by mutation from the initial seed, the Fock–Goncharov cluster chamber associated to is the subset

identified with

via . The (Fock–Goncharov) cluster complex is the set of all such chambers.

Lemma 2.10.

Suppose we are given fixed data satisfying the injectivity assumption, and suppose we are given an initial seed. For a seed obtained by mutation from the initial seed, the chamber (also identified with via is the Fock–Goncharov cluster chamber associated to . Hence the Fock–Goncharov cluster chambers are the maximal cones of a simplicial fan (of not necessarily strictly convex cones). In particular is identified with for any choice of seed giving an identification of with .

Proof.

The identification of the chamber is immediate from the definition. The result then follows from the chamber structure of Construction 1.30 and the fact that the are the Fock–Goncharov tropicalizations of the mutations for . It is obvious each maximal cone is simplicial, and each adjacent pair of maximal cones meets along a codimension 1 face of each. Hence we obtain a simplicial fan.

Construction 2.11.

See Appendix B for a review of the cluster variety with principal coefficients, . Any seed gives rise to a scattering diagram living in

the second equality by Proposition B.2(3). Indeed in this situation, the injectivity assumption is satisfied since the form on is nondegenerate (which is the reason we use instead of or ). Indeed, the vectors are linearly independent. Note by Theorem 1.21, contains the scattering diagram

Recall from Proposition B.2 that we have a canonical map which is defined on cocharacter lattices by the canonical projection ; see Equation B.4. Thus the tropicalization

coincides with this projection, which can be viewed as the quotient of an action of translation by . By Definition 1.4, walls of are of the form for . Thus all walls are invariant under translation by , and thus are inverse images of walls under . So even though may not satisfy the injectivity assumption necessary to build a scattering diagram, we see that is the inverse image of a subset of canonically defined independently of the seed. In particular, note that the Fock–Goncharov cluster chamber in associated to the seed (where for all ) pulls back to the corresponding Fock–Goncharov cluster chamber in .

The following was conjectured by Fock and Goncharov, Reference FG11, §1.5:

Theorem 2.13.

For any initial data the Fock–Goncharov cluster chambers in are the maximal cones of a simplicial fan.

Proof.

When the injectivity assumption holds, this follows from Lemma 2.10. In particular it holds for . Now the general case follows by the above invariance of under the translation by .

Example 2.14.

Consider the rank 3 skew-symmetric cluster algebra given by the matrix

Then projecting the walls of to via , one obtains a collection of walls in a three-dimensional vector space. One can visualize this by intersecting the walls with the affine hyperplane . The collection of resulting rays and lines appears on the first page of Reference FG11. While Fock and Goncharov were not aware of scattering diagrams in this context, in fact there the picture represents the same slice of the cluster complex, and hence coincides with the scattering diagram.

The cluster complex in fact fills up the half-space . There is no path through chambers connecting and .

This example is particularly well known in cluster theory, and gives the cluster algebra associated with triangulations of the once-punctured torus.

3. Broken lines

We will explain how a scattering diagram determines a class of piecewise straight paths which will allow for the construction of theta functions. The notion of broken line was introduced in Reference G09 and was developed from the point of view of defining canonical functions in Reference CPS and Reference GHK11.

We choose fixed data and a seed as described in Appendix A, and assume it satisfies the injectivity assumption. This gives rise to the group described in §1.1 which acts by automorphisms of for a choice of monoid containing and with . The group also acts on the rank 1 free -module for any , with a log derivation acting on as usual to give .

We then have:

Definition 3.1.

Let be a scattering diagram in the sense of Definition 1.6, and let and . A broken line for with endpoint is a piecewise linear continuous proper path with a finite number of domains of linearity. This path comes along with a monomial for each domain of linearity of . This data satisfies the following properties:

(1)

.

(2)

If is the first (and therefore unbounded) domain of linearity of , then .

(3)

For in a domain of linearity , .

(4)

Let be a point at which is not linear, passing from domain of linearity to . Let

Then is a term in the formal power series .

Remark 3.2.

Note that since a broken line does not pass through a singular point of , we can write

where is primitive, vanishes on each , and is positive by item (3) of the definition of broken line. It is an important feature of broken lines that we never need to invert.

Definition 3.3.

Let be a scattering diagram, and let and . For a broken line for with endpoint , define

(where is for initial),

and

to be the monomial attached to the final (where is for final) domain of linearity of . Define

where the sum is over all broken lines for with endpoint .

For , we define for any endpoint

In general, is an infinite sum, but makes sense formally:

Proposition 3.4.

.

Proof.

It is clear by construction that for any broken line with , we have . So it is enough to show that for any , there are only a finite number of broken lines such that , , and .

First note by the assumption that , there are only a finite number of choices for such that . Fix a choice for . Second, to test that there are finitely many broken lines with , and , we can throw out any wall with , so we can assume is finite. Third, no broken line with can bend more than times. Thus there are only a finite number of possible ordered sequences of walls at which can bend. Fix one such sequence. One then sees there are at most a finite number of broken lines with , bending at . Indeed, one can start at and trace a broken line backward, using that the final direction is . Crossing a wall and passing from domain of linearity (for smaller ) to domain of linearity (for larger ), one sees that knowing the monomial attached to restricts the choices of monomial on to a finite number of possibilities. This shows the desired finiteness.

The most important general feature of broken lines is the following:

Theorem 3.5.

Let be a consistent scattering diagram, and let and be two points. Suppose further the coordinates of are linearly independent over , and the same is true for . Then for any path with endpoints and for which is defined, we have

Proof.

This is a special case of results of Reference CPS, §4. The generality condition on and guarantees that we do not have to worry about broken lines which pass through joints (which we are not allowing). Indeed, the dimension of the -span of the coordinates can drop by at most one along a line with rational slope, and a point in a joint has two independent -linear relations in its coordinates.

Let us next consider how broken lines change under mutation. Let be a seed, and let be as in Definition 1.25.

Proposition 3.6.

defines a one-to-one correspondence between broken lines for with endpoint for and broken lines for with endpoint for . This correspondence satisfies, depending on whether or ,

where acts linearly on the exponents. In particular, we have

where the superscript indicates which scattering diagram is used to define the theta function.

Remark 3.7.

By Propositions 3.6 and 2.4, when the injectivity assumption holds, broken lines make sense in independent of a choice of seed.

Proof.

Given a broken line for , we define to have underlying map . Subdivide domains of linearity of so that we can assume any domain of linearity satisfies or . In the two cases, the attached monomial becomes or respectively. We show that is a broken line for with endpoint , with respect to the scattering diagram , which is equal to , by Theorem 1.24. Indeed, the only thing to do is to analyze what happens when crosses . So suppose in passing from a domain of linearity to a domain of linearity , crosses , so that is a term in

Assume first that passes from to . Then is a term in

showing that satisfies the correct rules for bending as it crosses the slab of .

If instead crosses from to , then is a term in

so again satisfies the bending rule at the slab .

The map on broken lines is then shown to be a bijection by observing , similarly defined, is the inverse to on the set of broken lines.

The following, which shows that cluster variables are theta functions, is the key observation for proving positivity of the Laurent phenomenon.

Proposition 3.8.

Let be a basepoint, and let . Then .

Proof.

This says the only broken line with asymptotic direction and basepoint has image , with attached monomial . To see this, suppose we are given a broken line with asymptotic direction which bends successively at walls . For each , there is an such that . Multiplying by a positive integer if necessary, we can assume that the monomial attached to upon crossing the wall changes by a factor . Now if is the image of the th linear segment of , we show inductively that

Indeed, for some , so initially is contained on the positive side of , i.e., is positive on , and hence after bending at , we see . Next, assume true for . Then , and if is the time when bends at the wall , we have and by the induction hypothesis. Thus . In addition, the derivative of along is , and

by skew symmetry of and . Thus

Since for all , any broken line with asymptotic direction which bends cannot terminate at the basepoint . This shows that there is only one broken line for terminating at .

Corollary 3.9.

Let be a cluster chamber, and let , . Then .

Proof.

Note for some vertex of , with associated seed . There is then a piecewise linear map with ; see Equation 1.31. Then the result follows by applying Proposition 3.8 to , and Proposition 3.6.

In the next section, we will identify theta functions which are polynomials with universal Laurent polynomials, i.e., elements of the cluster algebra associated to the fixed and seed data. It will follow from the above corollary that cluster monomials are in fact theta functions.

Example 3.10.

Figures 3.1 and 3.2 show some examples of broken lines in the case of Example 1.15 with . In the first figure, we take ; in the second, . Neither of these lie in the cluster complex: the union of all cones in the cluster complex is . In this case the only bends occur on the original lines of , as any bending along the additional rays of the scattering diagram will result in the broken line shooting back out, unable to reach the first quadrant containing the basepoint . In the figures, the final line segment is labeled with its attached monomial, so that the theta function is a sum of these labels. One finds

In Reference CGMMRSW it was shown that for any , with lying in the first quadrant, the with ranging over all elements of coincides with the greedy basis Reference LLZ13.

4. Building from the scattering diagram and positivity of the Laurent phenomenon

Throughout this section we work with initial data satisfying the injectivity assumption, so we obtain the cluster chamber structure from described in Construction 1.30. In particular, this condition holds for initial data ; see Appendix B.

In what follows, we will often want to deal with multiple copies of etc., indexed either by vertices of or chambers . To distinguish these (identical) copies, we will use subscripts or ; e.g., the scattering diagram lives in , and chambers in give, under the identification , the Fock–Goncharov cluster complex by Lemma 2.10. In particular the cluster chambers of and are in canonical bijection.

Construction 4.1.

Fix a seed . We use the cluster chambers to build a positive space. We attach a copy of the torus to each cluster chamber .

Given any two cluster chambers of , we can choose a path from to . We then get an automorphism which is independent of choice of path. If we choose the path to lie in the support of the cluster complex, then by Remark 1.29 (which shows in particular that the scattering functions on walls of the cluster complex are polynomials, as opposed to formal power series), the wall crossings give birational maps of the torus, and hence we can view as giving a well-defined map of fields of fractions

This induces a birational map

which is in fact positive.

We can then construct a space by gluing together all the tori , via these birational maps; see Reference GHK13, Proposition 2.4. We call this space (with its atlas of tori) .

We write if we need to make clear which seed is being used.

We check first that mutation equivalent seeds give canonically isomorphic spaces.

We recall first something of the construction of . Fix a seed . Then we have positive spaces

where each atlas is parameterized by vertices of the infinite tree . We write, e.g., for the open subset parameterized by . If we obtain a seed by mutation from , then we can think of the tree as a subtree of rooted at , and thus we obtain natural open immersions

These are easily seen to be isomorphisms. Under this immersion, the open cover of is identified canonically with the subcover of indexed by vertices of (but in either atlas there are many tori identified with the same open set of the union). Because of this we view as independent of the choice of seed in a given mutation equivalence class.

Given vertices of , we have birational maps

induced by the inclusions and respectively.

In what follows, we use the same notation for the restriction of a piecewise linear map to a maximal cone on which it is linear and the unique linear extension of this restriction to the ambient vector space.

Proposition 4.3.

Let be a seed. Let be the root of , any other vertex. Consider the Fock–Goncharov tropicalization of . Its restriction to each cluster chamber is a linear isomorphism onto the corresponding chamber . The linear map

induces an isomorphism

These glue to give an isomorphism of positive spaces .

In view of Proposition 4.3, we can view as independent of the seed in a given mutation class.

Proof.

It is enough to treat the case where is adjacent to via an edge labeled with , so that , as in general is the inverse of a composition of mutations . Note in this special case by Proposition 2.4, the definition of in Appendix A, and the formula for the -cluster mutation (see, e.g., Reference GHK13, (2.5)). So

is the isomorphism determined by the linear map . Proposition 4.3 amounts to showing commutativity of the diagram, for , , ,

where in the left column indicates wall crossings in while in the right column the wall crossings are in .

If and are on the same side of the wall , then commutativity follows immediately from Theorem 1.24. So we can assume that and are adjacent cluster chambers separated by the wall , and further without loss of generality that is nonnegative on . Now by Remark 1.29 there is only one wall of () contained in , with support itself and attached function (resp. ). Now it is a simple calculation:

This gives the desired commutativity.

Next we explain how to identify with .

Recall for each vertex of there is an associated cluster chamber in the cluster complex. While the atlas for is parameterized by chambers of , we can use a more redundant atlas indexed by vertices of , equating with . The open sets and the gluing maps in this redundant atlas are the same as in the original, but in the redundant atlas a given open set might be repeated many times.

Theorem 4.4.

Fix a seed . Let be the root of , and let be any other vertex. Let be the linear map . Let be the associated map of tori. These glue to give an isomorphism of positive spaces

Furthermore, the diagram

is commutative, where the right-hand vertical map is the isomorphism of Proposition 4.3, the left-hand vertical map the isomorphism given in Equation 4.2, and the horizontal maps are the isomorphisms just described.

Proof.

Let . The desired isomorphism is equivalent to commutativity of the diagram

where the right-hand vertical arrow is given by wall crossings in between the cluster chambers for . For this we may assume there is an oriented path from to in and, thus, that .

The commutativity of Equation 4.5 is equivalent to the commutativity of

where the right-hand vertical map is the restriction of the isomorphism of Proposition 4.3. We argue the commutativity of Equation 4.7 first and then show that this implies the commutativity of Equation 4.6.

Each map in Equation 4.7 is an isomorphism, induced by the restrictions of tropicalizations of various to various cluster chambers. Explicitly, on character lattices, we have the corresponding diagram

which is obviously commutative as tropicalization is functorial and .

Now for the commutativity of Equation 4.6. It is enough to check the case when there is an oriented edge from to in labeled by . We claim we may also assume . Indeed, assume we have proven commutativity in this case. We draw a cube, whose back vertical face is the diagram Equation 4.6, and whose front vertical face is the analogous diagram for , which is commutative by assumption. The top and bottom horizontal faces are instances of Equation 4.7, and the right-hand vertical face is the commutative diagram of atlas tori giving the isomorphism of Proposition 4.3. Finally, the left-hand vertical face consists of equality of charts or birational maps coming from inclusions of these tori in or , and thus is commutative. Now the commutativity of the back vertical face Equation 4.6 follows.

Finally, to show Equation 4.6 when , i.e., , we note is automatically the identity, and is also the identity, by Definition 1.22, and the identification of as Fock–Goncharov tropicalization of the birational map of tori . Thus the commutativity amounts to showing that the wall-crossing automorphism of fraction fields, given by crossing the wall from the negative to the positive side, is the pullback on rational functions of the birational mutation . Note the only scattering function on the wall is , so this follows from the coordinate free formula for the birational mutation; see, e.g., Reference GHK13, (2.6).

We can now make precise the relationship between theta functions and cluster monomials mentioned at the end of §3.

Definition 4.8.

Given fixed and initial data , if a seed is given, with the dual basis and , a cluster monomial in this seed is a monomial on of the form with and the nonnegative for . By the Laurent phenomenon Reference FZ02b, such a monomial always extends to a regular function on . A cluster monomial on is a regular function which is a cluster monomial in some seed.

Theorem 4.9.

Let be fixed data satisfying the injectivity assumption, and let be an initial seed. Let and for some . Then is a positive Laurent polynomial which expresses a cluster monomial of in the initial seed . Further, all cluster monomials can be expressed in this way.

Proof.

By Theorem 4.4 and Proposition 4.3 we have a canonical isomorphism of positive spaces . Let be the root of and let be any vertex of . Then we have , and the cluster monomials for the seed are just the monomials on with . By Theorem 4.4, this is identified with the monomial on , as takes to by Proposition 4.3. So the cluster monomials for the chart indexed by in are of the form with . Furthermore, if for each vertex of , is a general basepoint, we have for by Corollary 3.9. By the definition of in Construction 4.1, the corresponding rational function on the open set is , where is a path from to lying in the support of . But by Theorem 3.5. Finally, is a positive Laurent series by Theorem 1.13 and the definition of broken lines. By the Laurent phenomenon, it is also a polynomial.

We can now remove the injectivity assumption to prove:

Theorem 4.10 (Positivity of the Laurent phenomenon).

Each cluster variable of an -cluster algebra is a Laurent polynomial with nonnegative integer coefficients in the cluster variables of any given seed.

Proof.

Since, as explained in Proposition B.11, each cluster variable lifts canonically from to , we can replace the initial data with , for which the injectivity assumption holds. The result then immediately follows from Theorem 4.9.

Remark 4.11.

When fixed and initial data , have frozen variables, there is a partial compactification of cluster varieties ; see Construction B.9. We have an analogous partial compactification , given by an atlas of toric varieties . The choice of fans is forced by the identifications of Proposition 4.3: for the root of , ( as in Construction B.9) and then . Now Proposition 4.3 and Theorem 4.4 (and their proofs) extend to the partial compactifications without change. One checks easily that all mutations in the positive spaces , and all the linear isomorphisms between corresponding tori in the atlases for preserve the monomials , (these are the frozen cluster variables), so that all the spaces come with canonical projection to , preserved by the isomorphisms between these positive spaces. We shall see in the next section that in the special case of the partial compactification of , the relevant fans are particularly well-behaved.

5. Sign coherence of - and -vectors

We begin with some philosophy concerning log Calabi–Yau varieties following on from the discussion of Reference GHK13, §1. Suppose are both log Calabi–Yau and is a Zariski open subset of , both having maximal boundary (Reference GHK13, Definition 1.5). The tropical sets (which are expected to parameterize the theta function basis of functions on the mirror) of and are canonically equal, and we expect the mirror to degenerate to the mirror . In particular when is an algebraic torus, we expect a canonical degeneration of to the dual torus , under which the theta functions degenerate to monomials (i.e., characters). When is an -cluster variety and is a cluster torus, it turns out this degeneration has a purely cluster construction: the choice of seed determines a canonical partial compactification of ; see Proposition B.2 and Remark B.10. The main point of this section is to show that ; see Corollary 5.3. This degeneration is central to what follows in this paper. For example, we prove linear independence of theta functions by showing they restrict to different characters on , and the Fock–Goncharov conjecture, false in general, is true in a formal neighborhood of this fiber. There are analogous degenerations (identified with this one when the Fock–Goncharov conjecture holds) for, e.g., , and here they are even more central, being the main tool we have for proving properties of this algebra (e.g., that its spectrum gives a Gorenstein log Calabi–Yau of the right dimension); see Theorem 8.32. The equality , while not at all obvious from the cluster atlas, is immediate using the alternative description of the previous section, as we now explain. Further, there are some immediate benefits, such as sign coherence of -vectors.

For the remainder of the paper, the only scattering diagram we will ever consider is ; see Construction 2.11. So we will often omit the superscript from the notation.

Construction 5.1.

Fix a seed for fixed data . By Construction 4.1, the scattering diagram gives an atlas for the space . (Technically, we should write to indicate we are constructing something isomorphic to ; however, this will make the notation even less readable.) This was constructed by attaching a copy of the torus to each cluster chamber , and (compositions of) wall-crossing automorphisms give the birational maps between them. By Theorem 4.4 this space is canonically identified with : has an atlas of tori parameterized by vertices of , and we have canonical isomorphisms for each vertex which induce the isomorphism .

In what follows, if is a vertex of , we write for the seed obtained by mutating (see Equation B.1) via the sequence of mutations dictated by the path from the root of to . As described in Remark B.10, the initial seed determines the partial compactification , given by the atlas of toric varieties

where is the cone generated by the subset of basis vectors of corresponding to the second copy of .

By Remark 4.11, the seed also determines a partial compactification (the superscript, thus the seed close to the overline in the notation, is responsible for the partial compactification), given by an atlas of toric varieties. Explicitly, if is a vertex of , the fan yields the partial compactification of in , and this is identified with via under the isomorphism of Theorem 4.4. Thus the fan giving the partial compactifaction of is

In fact, this fan is easily calculated:

Lemma 5.2.

The cones , and thus the toric varieties in the atlas for the partial compactification , are the same for all . Each is equal to the cone spanned by the vectors , where and denotes the dual basis.

Proof.

is the given cone, by definition of the seed . By Construction B.9 the other fans are given by applying the geometric tropicalization of the birational gluing of the tori in the atlas for . These birational maps are given by wall crossings in . But for each wall between cluster chambers, the wall crossing is a standard mutation (notation as in §2), for some . The attached scattering function is for some in the convex hull of , and . But then . Thus the geometric tropicalization fixes all the by Equation 2.3, and so the fan is constant.

Corollary 5.3.

Fix a seed , and let be the root of . The following hold:

(1)

The fiber of over is see Proposition B.2 for the definition of .

(2)

The mutation maps

for the atlas of toric varieties defining are isomorphisms in a neighborhood of the fiber over .

(3)

For the partial compactification with atlas corresponding to cluster chambers of , the corresponding mutation map between two charts (which by Lemma 5.2 has the same domain and range) is an isomorphism in a neighborhood of the fiber and restricts to the identity on this fiber.

Proof.

It is clear that (3) implies (2) implies (1).

For (3), the scattering diagram is trivial modulo the (which pulls back to ), because this holds for the initial walls, with attached functions . Now for any adjacent vertices , the birational gluing map is given on the level of monomials by for a regular function on and some and any , and by the above is identically when restricted to the torus where the are zero. On the other hand, this birational map gives an isomorphism between the open subsets of and where is nonzero. In particular, the gluing maps are isomorphisms in the neighborhood of the fiber where all vanish and are the identity on that fiber.

The proof of the corollary shows the utility of constructing as the positive space associated to the cluster chambers in the scattering diagram . Next we show that sign coherence of -vectors follows easily from the corollary.

In what follows, given a seed obtained via mutation from , we write for the exchange matrix for this seed, with

The -vectors of this seed are the rows of the right-hand submatrix.

Corollary 5.5 (Sign coherence of -vectors).

For any vertex of and fixed satisfying , either the entries , are all nonpositive or these entries are all nonnegative.

Proof.

The result follows directly from Corollary 5.3 by writing down the mutation in cluster coordinates. Following the notation given in Appendix B, we have the fixed seed which determines and the family . The corresponding initial seed for is

and the coordinate on pulls back to on . These are the frozen cluster variables for . Note that where is the dual basis to the basis of .

A vertex corresponds to a seed for with corresponding seed for , with obtained from by a sequence of mutations. The are no longer necessarily given by the . Write , for the corresponding basis of . The cluster variables on the corresponding torus are . Say is a vertex of adjacent to along an edge labeled by . Then

and the cluster coordinates are . Since the last cluster variables are frozen, , .

The fan determining a toric variety in the atlas for consists of a single cone spanned by , and

Similarly

The mutation is given by the exchange relation Reference FZ07, (2.15) (see Reference GHK13, (2.8) in our notation) which is, with ,

where

Now fails to be an isomorphism exactly along the vanishing locus of

This locus is disjoint from the central fiber by Corollary 5.3. On the other hand it is disjoint from the central fiber if and only if exactly one of is the empty product, i.e., the constant monomial . Sign coherence is the statement that at least one of , is the empty product.

Recall from Definition 4.8 the notion of cluster monomial, and also note from Proposition B.2(2) the -action on .

Definition 5.6.

By Proposition B.11, the choice of seed provides a canonical extension of each cluster monomial on to a cluster monomial on . Each cluster monomial on is a -eigenfunction under the above action. The -vector with respect to a seed (see Reference FZ07, (6.4)) associated to a cluster monomial of is the -weight of its lift determined by .

We now give an alternative description of -vectors, which will lead to a more intrinsic definition of -vector (Definition 5.8). This in turn generalizes to all the different flavors of cluster varieties (Definition 5.10).

Proposition 5.7.

Fix a seed , giving the partial compactification and -equivariant . The central fiber is a -torsor. Let be a cluster monomial on and let be the corresponding lifted cluster monomial on . This restricts to a regular nonvanishing -eigenfunction along and so canonically determines an element of (its weight). This is the -vector associated to .

Proof.

Let determine the seed in which is defined as a monomial. By Corollary 5.3, all mutations are isomorphisms near the central fiber of , so it is enough to check that is regular on the toric variety , and it restricts to a character on its central fiber. But this is true by construction: if the seed is , then the cluster variables for the seed on the torus are and is the fan with rays spanned by the . Thus the lift of is regular on , and hence is regular in a neighborhood of . Furthermore, it is nonzero on since the canonical lift only involves monomials , which are nonvanishing on the strata of . The final statement follows since the restriction of the variable to the central fiber will have the same -weight, as the map is -equivariant, and fixes .

Definition 5.8.

Writing , let be a cluster monomial of the form on a chart , . Note that for all , so after identifying with , yields a point in the Fock–Goncharov cluster chamber , as defined in Lemma 2.10. We define to be this point of .

Corollary 5.9.

Let be a cluster monomial on , and fix a seed giving an identification . Then under this identification, is the -vector of the cluster monomial with respect to .

Proof.

We first note that if is a monomial on the chart with , , then the image of under the identification is , where as usual is the rational map induced by the inclusions .

The choice of the seed gives the lift of to a cluster monomial on . Using the identification of with , is identified with a monomial of the form on the chart (or , depending on how one chooses to parameterize charts of ). Let be the root of . By Lemma 5.2, the corresponding chart of is the toric variety defined by the fan . By Proposition 5.7, is a regular function on which is nonvanishing along . The -weight is the -vector. Since is the cone spanned by in , where , one sees that .

Thus to show the corollary, it is enough to show that . Note however a similar statement is already true at the level of . Indeed, in the chart of , the monomial takes the form for some , and lies in the positive chamber of . But is the image of this positive chamber under the map , where now is the map induced by the inclusions . Now by Theorem 4.4, and , so we see that .

Now because there is a well-defined map by Proposition B.2(4), with given by projection onto , this projection is compatible with the tropicalizations and , i.e., . Thus , as desired.

This corollary shows us how to generalize the notion of -vector to any cluster variety:

Definition 5.10.

Let be a cluster variety, suppose that is a global monomial (see Definition 0.1) on , and let be a seed such that is the character , . Define the -vector of to be the image of under the identifications of §2:

We write the -vector of as .

Note that the definition as given is not clearly independent of the choice of seed , but for a cluster variety of type, the previous corollary shows this. This independence will be shown in general in Lemma 7.10.

By Reference NZ, the sign coherence for -vectors (proved in Corollary 5.5 here), implies a sign coherence for -vectors. Here we give a much shorter proof using the above description of -vectors.

Theorem 5.11 (Sign coherence of -vectors).

Fix initial seed , with as usual. Given any mutation equivalent seed , the th-coordinates of the -vectors for the cluster variables of this seed, expressed in the basis , are either all nonnegative or all nonpositive.

Proof.

By Corollary 5.9, the -vectors in question are the generators of a chamber in the cluster complex of , defined as the images of the cluster chambers of under the projection , by Theorem 2.13. The hyperplanes are thus walls in the cluster complex. In particular, is either nonnegative everywhere on a chamber or nonpositive everywhere on a chamber. The theorem follows.

For future reference, we record the relationship between -vectors and the cluster chambers in the case of no frozen variables. Fix a seed . By Lemma 2.10, each mutation equivalent seed has an associated cluster chamber . This is a full-dimensional strictly simplicial cone, generated by a basis of consisting of -vectors of the cluster variables of . The facets of are thus in natural bijection with the elements of (or the indices in ).

Lemma 5.12.

The facet of corresponding to is the intersection of with the orthogonal complement of the -vector for the corresponding element of (the corresponding mutation of the Langlands dual seed ; see Appendix A). Furthermore, each -vector for is nonnegative on .

Proof.

This is the content of Reference NZ, Theorem 1.2 with the condition Reference NZ, (1.8) holding by our Corollary 5.5. The -vectors used in Reference NZ are precisely the -vectors of the cluster variables .

6. The formal Fock–Goncharov conjecture

In this section we associate in a canonical way to every universal Laurent polynomial on a formal sum , , which, roughly speaking, converges to at infinity in each partial compactification . To give such an expression for a single is quite easy; see Propositions 6.4 and 6.5. The crucial point, remarkably, is that these coefficients are independent of ; see Theorem 6.8, our alternative to the Fock–Goncharov conjecture (which fails in general). This establishes the connection between and , and is key to one of our main technical results; see the proof of Proposition 8.22.

Choose a seed . We let , let , set

and write the map induced by also as

We use the notation for a variety , so that, e.g., is the upper cluster algebra. We define

Note that for any , for some monomial in the . This induces a canonical inclusion

where is the monoid generated by . Let be the projection, and set

Recall that a choice of seed determines a scattering diagram with initial walls for . We let be the monoid generated by . We have the cluster complex of cones in , with cones for each vertex of .

Similarly to the above discussion, induces maps

The isomorphism between and discussed in Construction 5.1 restricts to give an isomorphism between and . Furthermore, as is described by gluing charts isomorphic to with the cone generated by for every chart by Lemma 5.2, in fact is described by gluing charts parameterized by isomorphic to

Note for , the birational map between the charts of indexed by and restrict to isomorphisms ; this is implied by Corollary 5.3(3).

We choose a generic basepoint for each . Then for any , by Proposition 3.4, we obtain as a sum over broken lines a well-defined series

satisfying by Theorem 3.5

The following definition will yield the structure constants for the theta functions:

Definition-Lemma 6.2.

Let . Let be chosen generally. There are at most finitely many pairs of broken lines with , and (see Definition 3.3 for this notation). We can then define

The integers are nonnegative.

Proof.

By definition of scattering diagram for , all walls have . Note also that because comes from a strictly convex cone, any element of can only be written as a finite sum of elements in in a finite number of ways. In particular, as , we can write for . Thus we have

and there are only a finite number of possible , . So with fixed, there are only finitely many possible monomial decorations that can occur on either . From this, finiteness is clear; cf. the proof of Proposition 3.4. The nonnegativity statement follows from Theorem 1.13, which implies for any broken line .

Definition 6.3.

For a monoid a lattice, we write for the subset of elements which can be written as a sum of noninvertible elements of .

Proposition 6.4.

Notation as above. The following hold:

(1)

For , is a regular function on , and the as varies glue to give a canonically defined function .

(2)

For each and , we have , and thus the for canonically define

The are linearly independent; i.e., we have a canonical inclusion of -vector spaces

(3)

for chosen sufficiently close to . In particular, is independent of the choice of sufficiently near , and we define

for chosen sufficiently close to .

(4)

restrict to bases of as a -vector space and -module, respectively.

Proof.

Using the isomorphism with , the basic compatibility Theorem 3.5 gives the gluing statement (1). To prove (4), using the -linearity, it is enough to prove the given restrict to basis as -module. By Corollary 5.3, the central fiber of is the torus . If , the only broken lines with and are straight lines. Thus these restrict to the basis of characters on the central fiber. Now the result follows from the nilpotent Nakayama lemma (see Reference Ma89, pg. 58, Theorem 8.4).

(2) follows immediately from (1) and (4).

For (3), it is enough to prove the equality in for each . The argument is the same as the proof of the multiplication rule in Reference GHK11, Theorem 2.38, which, as it is very short, we recall for the reader’s convenience. We work with the scattering diagram modulo , which has only finitely many walls with nontrivial attached function. Expressing in the basis of (4), we examine the coefficient of . We choose a general point very close to , so that lie in the closure of the same connected component of (where denotes the finitely many walls nontrivial modulo ). By definition of ,

Now observe first that there is only one broken line with endpoint and : this is the broken line whose image is . Indeed, the final segment of such a is on this ray, and this ray meets no walls, other than walls containing , so the broken line cannot bend. Thus the coefficient of can be read off by looking at the coefficient of the monomial on the right-hand side of the above equation. This gives the desired formula to order . The finiteness argument of Definition-Lemma 6.2 then shows that for any given , chosen sufficiently close to will work for all .

By the proposition, each has a unique expression as a convergent formal sum , with coefficients . This immediately implies:

Proposition 6.5.

Notation as in Proposition 6.4. There is a unique inclusion

given by

We have for all .

Definition 6.6.

For , write on the torus chart of corresponding to a seed . We also have a formal expansion as for some . Set

If is the monoid generated by , it is easy to check from the construction of theta functions that

Remark 6.7.

Note that is constructed by gluing together various via isomorphisms, and hence . Thus by Proposition 6.4(4), a collection of the yields a basis for regular functions on with the property that this basis restricts to a monomial basis on the underlying reduced space. It remains a mystery about theta functions in general whether they satisfy some other interesting characterizing properties, such as the heat equation satisfied by theta functions on abelian varieties.

The main point of the following theorem is that on , is independent of the seed .

Theorem 6.8.

There is a unique function

with all the following properties:

(1)

is compatible with the -module structure on and the -translation action on in the sense that

for all , , .

(2)

For each choice of seed , the formal sum converges to in .

(3)

If then unless , and

and the coefficients are the coefficients for the expansion of viewed as an element of in the basis of theta functions from Proposition 6.4.

(4)

For any seed obtained via mutations from , is the composition of the inclusions

given by Equation 6.1 and Proposition 6.5. This sends a cluster monomial to the delta function for its -vector .

In the notation of Definition 6.6, for any seed . In particular the sets of Definition 6.6 are independent of the seed, depending only on .

Proof.

It is easy to see from Proposition 6.4 that is the unique function which satisfies conditions (1)–(3) of Theorem 6.4 for the given seed . Moreover, it satisfies condition (4) for . Thus it is enough to show that is independent of the choice of seed.

The basic idea is that expresses as a sum of theta functions. As the theta functions are linearly independent, the expression is unique. But as the sums can be infinite, we make the argument in the appropriate formal neighborhood.

For a seed we write for the fan in with rays spanned by the . We write for the fan in with rays spanned by the .

Clearly, for the invariance it is enough to consider two adjacent seeds, say and obtained, without loss of generality, by mutation of the first basis vector .

We consider the union of the two tori in the atlas for , glued by the mutation , which we recall is given by

where ; see Reference GHK13, (2.6). We will partially compactify this union by gluing the toric varieties

writing

under this gluing. Note this union of toric varieties is not part of the atlas for either or (for either of these, the fans determining the atlases for the toric compactifications are related by geometric tropicalization of the birational mutation, but here , and thus ).

Note for , while , (see, e.g., Reference GHK13, (2.3)). Thus the two cones share a codimension 1 face and form a fan, . Let . By construction the rational map is regular, and the same holds for the seed . Observe that commutes with the second projection . From this it follows that is regular. Note the toric boundary has a unique complete one-dimensional stratum and two zero strata , , whose complements in the we write as , respectively. We write, e.g., for the th-order neighborhood of this curve, and, e.g., for the scheme-theoretic inverse image . Finally, let

We will show that theta functions give a basis of functions on these formal neighborhoods. To make the computation transparent, we introduce coordinates.

We let , , observing that for all . In particular , . Since there is a map of fans from to the fan defining by dividing out by the subspace spanned by , there is a map . We can pull back to , getting a line bundle with monomial sections pulled back from with in the above notation. The open subset of where is given explicitly up to codimension 2 by the hypersurface

where and .

Note the points

lie in the chambers of corresponding to and respectively, and thus by Proposition 3.8 these points determine theta functions in , which are of course the corresponding cluster monomials .

We have the exactly analogous description for the open subset .

Next observe that all but one of the functions attached to walls in is trivial modulo the ideal . Indeed, the unique nontrivial wall is . It follows from Theorem 1.28 that modulo the scattering diagram has only finitely many nontrivial walls, and is regular on , for a basepoint in the distinguished chamber , so long as , the projection.

Let , . Noting , we can set

Observe is the subscheme of defined by the ideal in the open subset . Note that the open subset of defined by is the union of the two tori .

Claim 6.9.

The following hold:

(1)

The collection , , forms a -basis of the vector space .

(2)

The collection , , forms a basis of as an -module.

(3)

The collection , , forms a -basis of .

Proof.

(2) implies (1) using the -linearity of the scattering diagram and multiplication rule with respect to the -translation. Similarly, (2) implies (3) by inverting .

For the second claim, by Reference GHK11, Lemma 2.30, we need only prove the statement for . To prove linear independence it is enough to show linear independence modulo for all . For this, again by Reference GHK11, Lemma 2.30, it is enough to check just over the fiber . This is the torus and the theta functions restrict to the basis of characters.

So it remains only to show the given theta functions generate modulo . Here we use the explicit description of the open subset of where above. This is an affine variety, and the ring of functions is clearly generated by the as a -algebra. On the other hand, by the explicit description of modulo the ideal , for ,

This shows theta functions generate as an -module, hence the result.

Of course there is an analogous claim for .

Now we can prove for .

By the -translation action on (and the corresponding -linearity of the scattering diagrams), to prove the equality, we are free to multiply by a monomial from the base of . Multiplying by a monomial in the , , we can then assume is a regular function on the open subset of where . Now in the notation of Definition 6.6, for or . It follows now from the fact that is finite modulo for any that each for is a finite Laurent polynomial modulo . Here are basepoints in the chambers indexed by and respectively.

Claim 6.10.

Modulo , the sums , are finite and coincide with in the charts indexed by and , respectively.

Proof.

By symmetry it is enough to treat . We can multiply both sides by a power of , and so we may assume is regular on and for each . Note , thus by construction modulo for any the sum is finite and equal to . By Claim 6.9(1), we have a (finite) expression modulo

Thus, for fixed and arbitrary we have modulo ,

By linear independence these expressions are the same, for all , thus the expressions are the same modulo .

Note that by Theorem 3.5, for , and induce the same regular function on . Thus we have by Claim 6.10 that

Now by Claim 6.9(3) (varying ) the coefficients in the sums are the same.

The theorem implies that the theta functions are a topological basis for a natural topological -algebra completion of :

Corollary 6.11.

For , let

denote precomposition by the action of translation by on . Let

be the vector subspace of functions such that for each seed , there exists for which the restriction of to has finite support for all . Then we have

is a complete topological vector space under the weakest topology so that each inclusion is continuous. Let be the delta function associated to . The are a topological basis for . There is a unique topological -algebra structure on such that with structure constants given by Definition-Lemma 6.2.

7. The middle cluster algebra

In this section, we prove one of the main theorems of the paper, Theorem 0.3. This is done in two steps. First, it follows from the results of the previous section and properties of theta functions in the case. This is easiest since the scattering diagram technology works best for . Second, we descend to the or case and the case, with the -type varieties appearing as fibers of and the variety as a quotient of by , using the case to deduce the result for these other cases.

7.1. The middle algebra for

Recall from Definition 1.32 that is the collection of chambers forming the cluster complex. Abstractly, by Lemma 2.10, this can be viewed as giving a collection of chambers in .

Proposition 7.1.

Choose . If for some generic basepoint there are only finitely many broken lines with and , then the same is true for any generic . In particular, is a positive universal Laurent polynomial.

Proof.

By positivity of the scattering diagram, Theorem 1.13, for any basepoint , has only nonnegative coefficients (though it may have infinitely many terms). Also, we know that for basepoints in different chambers, the are related by wall crossings by Theorem 3.5, which in turn are identified with the mutations of tori in the atlas for . So the determine a universal positive Laurent polynomial if and only if we have finiteness of broken lines ending at any in any chamber of . If we vary in the chamber, does not change. So it is enough to check that if is a polynomial, the same is true of for in an adjacent chamber to close to the wall . We can work in some seed. Let the wall be contained in , , with , and denote the wall-crossing automorphism from to as . Recall that , where for walls between cluster chambers, is some positive Laurent polynomial (in fact it has the form for some ).

Monomials are then divided into three groups, according to the sign of . This sign is preserved by , as takes the same value on each exponent of a monomial term in as takes on .

Monomials with are then invariant under , so these terms in coincide with those in . Hence there are only a finite number of such terms in .

The sum of terms of the form in with , which we know forms a Laurent polynomial, is, by the explicit formula for , sent to a polynomial. So it only remains to show that there are only finitely many terms in with . Suppose the contrary is true. The direction vector of each broken line contributing to such terms at is toward the wall , and so we can extend the final segment of any such broken line to obtain a broken line terminating at some point (depending on ) in the same chamber as . As there are no cancellations because of the positivity of all coefficients and does not depend on the location of inside the chamber by Theorem 3.5, we see that has an infinite number of terms, a contradiction.

Definition 7.2.

Let be the collection of such that for some (or equivalently, by Proposition 7.1, any) generic there are only finitely many broken lines with , .

Definition 7.3.

We call a subset intrinsically closed under addition if and implies .

Lemma 7.4.

Let be intrinsically closed under addition. The image of in (under the identification induced by the seed is closed under addition for any seed . If for some seed is a toric monoid (i.e., the integral points of a convex rational polyhedral cone), then this holds for any seed.

Proof.

Choose a seed . Then straight lines in Definition-Lemma 6.2 show . This gives closure under addition. Now suppose is a toric monoid, generating the convex rational polyhedral cone . Then is a rational polyhedral cone with integral points . As this set of integral points is closed under addition, is convex, and so its integral points are a toric monoid.

Recall from the introduction the definition of global monomial (Definition 0.1).

Theorem 7.5.

Let

be the set of integral points in chambers of the cluster complex. Then

(1)

.

(2)

For

is a finite sum (i.e., for all but finitely many with nonnegative integer coefficients. If , then .

(3)

The set is intrinsically closed under addition. For any seed , the image of is a saturated monoid.

(4)

The structure constants of Definition-Lemma 6.2 make the -vector space with basis indexed by ,

into an associative commutative -algebra. There are canonical inclusions of -algebras

Under the first inclusion a cluster monomial is identified with for its -vector. Under the second inclusion each is identified with a universal positive Laurent polynomial.

Proof.

(1) is immediate from Corollary 3.9. For (2), first note that the coefficients are nonnegative by Definition-Lemma 6.2. Suppose . Take a generic basepoint in some cluster chamber. Then is the product of two Laurent polynomials, thus it is a Laurent polynomial. It is equal to by (3) of Proposition 6.4, and hence this sum must be finite, as it involves a positive linear combination of series with positive coefficients. Further, each appearing with must be a Laurent polynomial for the same reason. Thus by Definition 7.2. (2) then immediately implies is intrinsically closed under addition.

For (4), note each , is a universal positive Laurent polynomial by Proposition 7.1. For , is the corresponding cluster monomial by Theorem 6.8(4). The inclusions of algebras, and the associativity of the multiplication on follow from Proposition 6.4.

Finally, we complete the proof of (3) by checking that is saturated. Assume for some integer . Take to be a generic basepoint in some cluster chamber. We show that the set of final monomials for broken lines with , is finite. By assumption (and the positivity of the scattering diagram), this holds with replaced by . So it is enough to show implies . Indeed, it is easy to see that for every broken line for ending at , there is a broken line for with the same underlying path, such that for every domain of linearity of , the exponents and of the monomial decorations of for and respectively, satisfy . This completes the proof of (3), hence the theorem.

The above theorem immediately implies:

Corollary 7.6.

Theorem 0.3 is true for .

The following shows our theta functions are well behaved with respect to the canonical torus action on .

Proposition 7.7.

Let . Then is an eigenfunction for the natural action on (see Proposition B.2, with weight given by the canonical map (the map being dual to the inclusion . In particular is an eigenfunction for the subtorus with weight where is given by .

Proof.

Pick a seed , giving an identification with . Pick also a general basepoint . We need to show that for any broken line in for with endpoint , is a semi-invariant for the action with weight . The action on the seed torus is given on cocharacters by the natural inclusion . By definition of we have , , so every monomial appearing in any function for is in the kernel of . The result for follows. The statement for now follows from the definitions.

With more work, we will define the middle cluster algebra for or .

7.2. From to and .

We now show how the various structures we have used to understand induce similar structures for and .

By Reference GHK13, §3, each seed (in the , and cases) gives a toric model for . The seed specifies the data of a fan , consisting only of rays (so the boundary of the associated toric variety is a disjoint union of tori). The seed also then specifies a blowup with codimension 2 center, the disjoint union of divisors in each of the disjoint irreducible components . If is the proper transform of , then there is a birational map . This map is an isomorphism outside of codimension 2 between and the upper bound (see Reference GHK13, Remark 3.13 and Reference BFZ05, Def. 1.1) , which we recall is the union of with for the adjacent seeds, . In the case or for very general , the inclusion is an isomorphism outside codimension 2. We have

From these toric models it is easy to determine the global monomials:

Lemma 7.8 (Global monomials).

Notation as immediately above. For , the character on the torus is a global monomial if and only if is regular on the toric variety , which holds if and only if for the primitive generator of each ray in the fan . For -type cluster varieties a global monomial is the same as a cluster monomial, i.e., a monomial in the variables of a single cluster, where the nonfrozen variables have nonnegative exponent.

Proof.

We have a surjection by construction of , and thus a monomial is regular on if and only if its pullback to is regular. Certainly such a function is also regular on . Conversely, suppose is not regular on . Then it has a pole on some toric boundary divisor . Now the support of is given by (resp. ) in the (resp. ) case, as explained in Reference GHK13, §3.2. In particular for , is nonempty, in the case because of the assumption that for stated in Appendix A. As has no zeros on the big torus, the divisor of zeros of will not contain the center . It follows that has a pole along the exceptional divisor over . Since , is not regular on . Thus we conclude that is regular on if and only if is regular on . Of course, is regular on if and only if for all primitive generators of rays of .

Now the rational map to the upper bound is an isomorphism outside codimension 2, so the two varieties have the same global functions. In the (or ) case, the inclusion is an isomorphism outside codimension 2 as well. This gives the theorem for or , and the forward direction for . The reverse direction for follows from the Laurent phenomenon. Indeed, the final statement of the lemma simply describes the monomials regular on , and a monomial of the given form is the same as a cluster monomial and these are global regular functions by the Laurent phenomenon.

Note that by Proposition B.2(4) we have canonical maps and with tropicalizations

Note identifies with the quotient of by the natural -action. Since identifies with the fiber over of , identifies with , where is the weight map given by .

Definition 7.9.

Let be a cluster variety. Define to be the set of -vectors (see Definition 5.10) for global monomials, which are characters on the seed torus , and to be the union of all .

Lemma 7.10.
(1)

For of -type is the set of integral points of the cone in the Fock–Goncharov cluster complex corresponding to the seed .

(2)

In any case is the set of integral points of a rational convex cone , and the relative interiors of as varies are disjoint. The -vector depends only on the function (i.e., if restricts to a character on two different seed tori, the -vectors they determine are the same).

(3)

For , the global monomial on is invariant under the action and thus gives a global function on . This is a global monomial and all global monomials on occur this way, and .

Proof.

(1) In the case, is the Fock–Goncharov cone by Lemmas 2.10 and 7.8. These cones form a fan by Theorem 2.13, and the fan statement implies that depends only on .

The case of (2) immediately follows also from the discussion in §5. The case follows from the -case (applied to ) by recalling that there is a map making into -torsor over ; see Proposition B.2(2). This map is defined on monomials by . Pulling back a monomial for under gives a -invariant global monomial for . Thus there is an inclusion by Proposition 7.7. Conversely, if and for a global monomial on , then there is some seed where is represented by a monomial on . Because lies in it is of the form for some . By Lemma 7.8, is nonnegative on the rays of , hence is nonnegative on the rays of . Hence defines a global monomial on . Thus . Furthermore, one then sees that the Fock–Goncharov cones for yield the cones for by intersecting with . This gives the remaining statements of (2) in the case, as well as (3).

Construction 7.11 (Broken lines for and ).

The case. Note that every function attached to a wall in is a power series in for some , thus is zero on all exponents appearing in these functions. Thus broken lines with both and initial infinite segment lying in remain in . In particular , and all their monomial decorations, e.g., , are in . We define these to be the broken lines in .⁠Footnote2

2

In fact each induces a collection of walls with attached functions, , living in , just by intersecting each wall with and taking the same scattering function. This is a consistent scattering diagram, and we are getting exactly the broken lines for this diagram. We will not use this diagram, as we can get whatever we need from .

The case. We define broken lines in to be images of broken lines in under (applying the derivative to the decorating monomials).

Definition 7.12.

We define

(1)

.

(2)

where the superscript denotes the invariant part under the group action.

Corollary 7.13.

Theorem 0.3 holds for .

Proof.

This follows immediately from the case by taking -invariants.

Moving on to the case, the following is easily checked:

Definition-Lemma 7.14.
(1)

Define

Noting that is invariant under -translation, we have . Furthermore, any choice of section of induces a bijection .

(2)

Define , where is given by . Given a choice of , the collection gives a -module basis for and thus a -vector space basis for . For the basis is independent of the choice of , while for it is independent up to scaling each basis vector (i.e., the decomposition of the vector space into one-dimensional subspaces is canonical).

The variety is a space with the tori glued by birational maps which vary with . It is then not so clear how to dualize these birational maps to obtain as it is not obvious how to deal with these parameters. However, the tropicalizations of these birational maps are all the same (independent of ) and thus the tropical sets should all be canonically identified with . So we just take:

Definition 7.15.

.

Theorem 7.16.

For the following modified statements of Theorem 0.3 hold.

(1)

There is a map

depending on a choice of a section . This function is given by the formula

if this sum is finite; otherwise, we take . This sum is finite whenever .

(2)

There is a canonically defined subset given by such that the restriction of the structure constants give the vector subspace with basis indexed by the structure of an associative commutative -algebra.

(3)

, i.e., contains the -vector of each global monomial.

(4)

For the lattice structure on determined by any choice of seed, is closed under addition. Furthermore is saturated.

(5)

There is a -algebra map which sends for to a multiple of the corresponding global monomial.

(6)

There is no analogue of Theorem 0.3 because the coefficients of the will generally not be integers.

(7)

is injective for very general and for all if the vectors , , lie in a strictly convex cone. When is injective, we have canonical inclusions

Taking gives Theorem 0.3 for the case.

Proof.

For (1), note that for , we have and on

using . Note that the sums are finite because . Restricting to gives the formula of (1).

The remaining statements follow easily from the definitions except for the injectivity of (7). To see this, fix a seed , which gives the second projection . Choose the section of to be . Note the collection of , are a -basis for . By the choice of , the restrict to the basis of monomials on the central fiber of . It follows that for any finite subset there is a Zariski open set such that , restrict to linearly independent elements of , . This gives the injectivity of for very general .

Now suppose the , span a strictly convex cone. We can then pick an such that for all . Now pick such that for , and set ; notation is as in Proposition B.2(2). By construction the second projection lies in the interior of the orthant generated by the dual basis . Take the one-parameter subgroup . Now, by Proposition B.2, the map is -equivariant, where the action on is given by the map of cocharacters . Thus has a one-dimensional orbit whose closure contains . So is in the closure of the orbit for all . In particular for all and all there is some with . Now from the -equivariance of the construction, Proposition 7.7, the linear independence holds for all .

Changing will change the -basis for , multiplying each by some character , . The restrictions to are then multiplied by the values .

Theorem 0.3 for now follows from setting , where for all .

It is natural to wonder:

Question 7.17.

Does the equality always hold?

Our guess is no, but we do not know a counterexample.

Certainly in general, for this implies , which is defined to be , coincides with , while we know that in general has many fewer global functions, see Reference GHK13, §7. So we look for conditions that guarantee , and . We turn to this in the next section.

Example 7.18.

In the cases of Example 1.15, the convex hull of the union of the cones of in is all of . Indeed, the first three quadrants already are part of the cluster complex. It then follows from the fact that is closed under addition and is saturated that .

In the case of Example 2.14, we know that

It then follows again from the fact that is closed under addition that either or . We believe, partly based on calculations in Reference M13, §7.1, that in fact the latter holds.

We show the analogue of Proposition 7.7 for the variety:

Proposition 7.19.

If , then is an eigenfunction for the natural action on .

Proof.

This is essentially the same as the proof of Proposition 7.7, noting that the monomials are invariant under the action, as by definition of .

We end this section by showing that linear independence of cluster monomials follows easily from our techniques. This was pointed out to us by Gregory Muller. In the skew-symmetric case, this was proved in Reference CKLP.

Theorem 7.20.

For any cluster variety, there are no linear relations between cluster monomials and theta functions in . More precisely, if there is a linear relation

in , then for all . In particular the cluster monomials in are linearly independent.

Proof.

Suppose we are given such a relation. We choose a seed and a generic basepoint . The seed gives an identification . We first show that if with , then satisfies the proper Laurent property, i.e., every monomial appearing in has for some .

Indeed, fix a section as in Definition-Lemma 7.14. As the restriction to gives a bijection between the cluster variables for and the cluster variables for , between the theta functions , and the theta functions for , and between the corresponding local expressions , it is enough to prove the claim in the case. This follows immediately from the definition of broken line. Indeed, if is a broken line ending at and with for all , then must be wholly contained in . But the unbounded direction of is parallel to , so it follows that .

We then have the relation

which we rearrange as

The collection of for are exactly the distinct cluster monomials for the seed . In particular all of their exponents are nonnegative. Thus both sides of the equation are zero. Since the cluster monomials for are linearly independent, we conclude for all . Varying the result follows.

8. Convexity in the tropical space

As explained in the introduction, the Fock–Goncharov conjecure is in general false, as a cluster variety has in general too few functions. The conjectured theta functions only exist formally, near infinity, in the sense of §6. The failure of convergence in general manifests itself in the existence of infinitely many broken lines with a given incoming direction and fixed basepoint, and also in nonfiniteness of the multiplication rule (for fixed infinitely many with ). This section and the next is devoted to the question of finding conditions on cluster varieties which guarantee the conjecture holds as stated. One can only expect a theta function basis for in cases when has enough functions. More precisely, a basis should exist when is finitely generated, and the natural map is an open immersion. Note the second condition is automatic by the Laurent phenomenon when is -type. Our main (and simple) idea is that we can replace the assumption of enough functions by the existence of a bounded convex polytope in , cut out by the tropicalization of a regular function. However, our notion of convexity is delicate, as a priori only has a piecewise linear structure. The correct notion is explained in §8.1.

Polytopes will play several roles. The existence of bounded convex polytopes implies various results on convergence of theta functions. We get finiteness of the multiplication rule, and so an algebra structure on ; see Proposition 8.17. For technical reasons (see Remark 8.12) we often have to replace enough global functions by enough global monomials (EGM), and we can make optimal use only of convex polytopes cut out by the tropicalizations of global monomials. EGM implies ; see Propositions 8.18 and 8.22. Convex polytopes give partial (full in the bounded polytope case) compactifications of by copying the familiar toric construction; see §8.5. These compactifications then degenerate to toric compactifications under (the analogue) of the canonical degeneration of to . We use these degenerations to prove is log Calabi-Yau; see Theorem 8.32. Convex cones in are intimately related with partial minimal models , and potential functions; see §9.2 and Corollary 9.17. Finally, our methods give several sufficient conditions to guarantee the full Fock–Goncharov conjecture; see Proposition 8.25 and Corollary 9.18. These statements are weaker, and more technical, than one would hope—the reason is our inability to prove in full generality that EGM (or better, the existence of a bounded convex polytope) implies . We have only optimal control over the subset (the cluster complex), and the technical statements are various ways of saying is sufficiently big.

The first issue is to make sense of the notion of convexity in .

8.1. Convexity conditions

The following is elementary:

Definition-Lemma 8.1.

By a piecewise linear function on a real vector space we mean a continuous function piecewise linear with respect to a finite fan of (not necessarily strictly) convex cones. For a piecewise linear function we say is min-convex if it satisfies one of the following three equivalent conditions:

(1)

There are finitely many linear functions such that for all .

(2)

for all and .

(3)

The differential is decreasing on straight lines. In other words, for a directed straight line with tangent vector , and general, then

where is general and the subscript denotes the point at which the differential is calculated.

We now define convexity for functions on for a cluster variety by generalizing the third condition above, using broken lines instead of straight lines:

Definition 8.2.
(1)

A piecewise linear function is a function which is piecewise linear after fixing a seed to get an identification . If the function is piecewise linear for one seed it is clearly piecewise linear for all seeds.

(2)

Let be piecewise linear, and fix a seed , to view . We say is min-convex for (or just min-convex if is clear from context) if for any broken line for in , is increasing on exponents of the decoration monomials (and thus decreasing on their negatives, which are the velocity vectors of the underlying directed path). We note that this notion is independent of mutation, by the invariance of broken lines, Proposition 3.6, and thus an intrinsic property of a piecewise linear function on .

We have a closely related notion, instead defined using the structure constants for multiplication of theta functions.

Definition 8.3.

We say that a piecewise linear is decreasing if for , with , . Here are the structure constants of Theorem 0.3.

Lemma 8.4.
(1)

If is min-convex, then is decreasing.

(2)

If is decreasing, then for any seed , we have min-convex in the sense of Definition-Lemma 8.1.

Proof.

(1) Let , be broken lines. Assume is min-convex and that very close to is the endpoint of each broken line, with . Then

where . Thus is decreasing.

(2) Suppose is decreasing. For any , and the linear structure on determined by any choice of seed , the contribution of straight lines in Definition-Lemma 6.2 (and item (1) of Theorem 7.16 in the case) shows for all . Thus for all positive integers and . By rescaling, the same is true for all positive rational numbers and and . Min-convexity in the sense of Definition-Lemma 8.1 then follows by continuity of .

Remark 8.5.

We do not know whether the converse of either statement of Lemma 8.4 holds.

We have a closely related concept, capturing the generalization of the notion of a convex polytope. For a closed subset, define the cone of

Note the closure is only necessary if is not compact, in which case is an asymptotic form of . Denote

which we view as a subset of . Note for , agrees with the obvious notion .

Definition 8.6.

We will call a closed subset positive if for any nonnegative integers , any , , and any with , we have .

Note that if is a cone, i.e., invariant under rescaling, then this definition agrees with Definition 7.3.

For a piecewise linear function , let

Definition 8.8.

A convex polytope is a subset of for which there exist a finite collection of affine linear functions with

By Lemma 8.4, if is min-convex in the sense of Definition 8.2 (or more generally, decreasing in the sense of Definition 8.3), then under any identification given by any seed, is a convex polytope.

Lemma 8.9.

If a piecewise linear function is decreasing, then is positive. Furthemore, is compact if and only if is strictly negative away from .

Proof.

Note that . Thus if is decreasing and with , then and thus , so . The second statement is obvious.

Next we discuss a result comparing these convexity conditions on and . From Proposition B.2(4) we have the natural map such that is the canonical projection , the quotient by the translation action.

Proposition 8.10.

Suppose is a positive polytope defined over (i.e., all the functions of Definition 8.8 are rationally defined Then is positive.

Proof.

Suppose , and . We can always write with and by the rationality assumption. Let be a positive integer such that and are all integral for .

We first observe that because , . This follows immediately from Definition-Lemma 6.2 and the argument given in the proof of saturatedness in Theorem 7.5. This latter argument shows that if there is a broken line with , , then there is a broken line with , .

We next observe that this implies that . Indeed, to show this, we need to show a bijection between the following sets of broken lines and their decorations for general and :

Once this is shown, if , then and by Definition-Lemma 6.2.

To get the bijections between the sets, we first recall that every wall of is invariant under the canonical -translation and is contained in a hyperplane for some . Thus acts on broken lines, by translation on the underlying path, keeping the monomial decorations the same. This gives the bijection between the second and third sets.

For the bijection between the first and second sets, we need to translate the decorations on each straight segment of by . This will change the slopes of each line segment. To do this precisely, take in the first set, say with straight decorated segments taken in reverse order, with the infinite segment. Suppose the monomial attached to is with . Say the bends are at points along a wall contained the hyperplane so that is parameterized (in the reverse direction to that of Definition 3.1) by , . Then we define

Observe that . Let be the segment , , with attached monomial . Then form the straight pieces of a broken line in the second set. This gives the desired bijection. We now conclude, as promised, that .

To complete the proof of the proposition, we note that as , positivity of implies and thus .

The chief difficulty now lies in constructing min-convex functions or positive polytopes. We turn to this next.

8.2. Convexity criteria

The following would be a powerful tool for construction min-convex functions on cluster varieties:

Conjecture 8.11.

If is a regular function on a log Calabi–Yau manifold with maximal boundary, then is min-convex. Here for the valuation .

Remark 8.12.

To make sense of the conjecture, one needs a good theory of broken lines, currently constructed in Reference GHK11 in dimension two, and here for cluster varieties of all dimensions. In dimension two, the conjecture has been proven by Travis Mandel Reference M14. Also, it is easy to see that in any case, for each seed and regular function , that (see Equation 2.5) is min-convex in the sense of Definition-Lemma 8.1. Indeed this is the standard (min) tropicalization of a Laurent polynomial. We hope to eventually give a direct geometric description of broken lines (without reference to a scattering diagram) for any log Calabi-Yau, as tropicalizations of some algebraic analogue of holomorphic disks. We expect the conjecture to follow easily from such a description.

In fact, we can prove Conjecture 8.11 for global monomials, which gives our main tool for constructing min-convex functions (our inability to prove the conjecture in general is the main reason we use the condition EGM rather than the more natural condition of enough global functions):

Proposition 8.13.

For a global monomial on , the tropicalization is min-convex, and in particular, by Lemma 8.4, decreasing. Further, if the global monomial is of the form for to be an integral point in the interior of a maximal-dimensional cone (see Definition 7.9), then evaluated on monomial decorations strictly increases at any nontrivial bend of a broken line in .

Proof.

First consider the case . Suppose is a global monomial which is a character on a chart indexed by . The integral points of the cluster chamber

correspond to characters of which extend to global regular functions on . Then by Lemma 7.8 such a character is regular on , i.e., it is a character whose geometric tropicalization Equation 2.6 has nonnegative value on each ray in the fan . These rays are spanned by , , the negatives of the initial scattering monomials for . Thus, because of the sign change between geometric and Fock–Goncharov tropicalization (see Equation 2.5), is nonnegative on the initial scattering monomials. Further, if for an interior point in , is positive on all initial scattering monomials. We now use this to show is min-convex.

Indeed, still working in the seed , we consider a broken line , with two consecutive monomial decorations . Let be points in the domain of in the two segments. Then

Now is some positive multiple of the scattering monomial. Thus since nonnegative on scattering monomials, . This gives min-convexity. Further, if lies in the interior of , then and is strictly decreasing on nontrivial bends.

The same argument then applies in the case. Indeed, recall from §7.2 that for any choice of seed, the scattering monomials in lie in . So it makes sense to evaluate functions defined only on on scattering monomials for .

For the case, a global monomial on pulls back to a global monomial on via the map . The result then follows from the case, as a piecewise linear function on is min-convex if and only if on is min-convex. Indeed, broken lines in are by definition images of broken lines on under .

We now introduce our key assumption, which is necessary for proving strong results about theta functions and the algebras they generate.

Definition 8.14.

We say that has EGM if for any , , there is a global monomial such that .

Lemma 8.15.

Under any of the identifications induced by a choice of seed, the set

is a closed convex subset of . The following are equivalent:

(1)

has EGM.

(2)

is bounded, or equivalently, the intersection of the sets for equals .

(3)

There exists a finite number of points such that

is bounded, or equivalently, the intersection of the sets for equals .

(4)

There is function whose associated polytope is bounded.

Proof.

By Proposition 8.13, is the intersection of closed rational convex polytopes (with respect to any seed) and hence is a closed convex set.

The equivalence of (1) and (2) is immediate from the definitions, while (3) clearly implies (2). For the converse, let be a sphere in centered at the origin. For each there is a global monomial such that , and thus there is an open subset on which is negative. The form a cover of , and hence by compactness there is a finite subcover . Taking gives the desired collection of .

Finally, we show the equivalence of (3) and (4). The , are exactly the global monomials on , thus generators of . Now for any finite collection of functions , , and for the positive universal Laurent polynomials (for example for global monomials), we have equality. Thus given (3), we take . Conversely, an element of is a linear combination of some collection of . Then is contained in , so if the latter is bounded, so is the former.

We note that the property of EGM is preserved by Fock–Goncharov duality in the principal coefficient case:

Proposition 8.16.

Let be fixed data, and let be the Langlands dual data. We write, e.g., for the corresponding lattice for the data as in Appendix A. For each seed , the canonical inclusion

commutes with the tropicalization of mutations and induces an isomorphism

For , the monomial on is a global monomial if and only if on is a global monomial. Finally, has EGM if and only if has EGM.

Proof.

The statement about tropical spaces is immediate from the definitions. (Note that a similar statement does not hold at the level of tori, so there is no isomorphism between and .) The statement about global monomials is immediate from Lemma 7.8. Now the final statement follows from the definition of EGM, the isomorphism of Proposition B.2(1), and the equality of Proposition B.2(3).

8.3. The canonical algebra

In Equation 0.2 we introduced as a -vector space. In the presence of suitable convex objects on , we can in fact put an algebra structure on using the structure constants given by . Further, if the EGM condition holds, then contains and is a finitely generated algebra. This often makes it easier to work with , as it is a more geometric object.

Precisely:

Proposition 8.17.

For or , suppose there is a compact positive polytope . Assume further that is top dimensional, i.e., . Then for , there are at most finitely many with . These give structure constants for an associative multiplication on

If there is a compact positive polytope then the same conclusion holds for the structure constants (which are all finite) and multiplication rule of for all .

Proof.

For or , the structure constants are defined in terms of broken lines. The finiteness is then immediate from Lemma 8.9. Indeed, given , we have , for some , by the fact that is top dimensional, and thus if , then lies in the bounded polytope . The algebra structure is associative by Proposition 6.4(3). The case follows from the case and the definitions of the structure constants and multiplication rule for ; see Theorem 7.16.

Corollary 8.18.

For or assume has EGM. For assume has EGM. Then defines a -algebra structure on .

Proof.

The case of follows from the case of so we may assume is either or . Using the EGM hypothesis and Lemma 8.15, we can find a finite collection such that the intersection of the finite collection of polytopes is bounded. But since is min-convex by Proposition 8.13, each of these polytopes is positive by Lemma 8.9. Thus the result follows from Proposition 8.17.

Finite generation of is a special case of a much more general result.

Theorem 8.19.

Let or , assume has EGM, and let be a positive polytope, which we assume is rationally defined and not necessarily compact. Then

is a finitely generated -subalgebra.

Proof.

Note that is a subalgebra of by the definition of positive polytope.

As in the proof of Corollary 8.18, we can choose so that is a compact positive polytope. Moreover, because boundedness of the intersection is preserved by small perturbation of the functions, we can assume that each is in the interior of some maximal dimensional cluster cone . Note the seed is then uniquely determined by by Lemma 7.10. It follows that is strictly increasing on the monomial decorations at any nontrivial bend of any broken line in by Proposition 8.13.

We define

to be the vector subspace spanned by all , where . Then is a graded subalgebra of by Proposition 8.13 (graded by -degree).

The result then follows from the following claim, noting that there is a natural surjection by sending , .

Claim 8.20.

is a finitely generated -algebra.

Proof. We argue first that is finitely generated. We work on , so that the multiplication rule is defined using broken lines for , as described by Proposition 6.4(3) and Definition-Lemma 6.2. Note is linear on . If , then modulo , unless , for otherwise . By Proposition 8.13, is strictly increasing on monomial decorations at any nontrivial bend of a broken line, and thus the only broken lines that will contribute (modulo ) to are straight, thus modulo ,

(addition here in ). Thus is the monoid ring associated to the rational convex cone

and is thus finitely generated.

The result now follows from the following general fact: If is a graded -algebra, a homogeneous element of degree , and is a finitely generated -algebra, then is also a finitely generated -algebra. Indeed, if are homogeneous elements generating , we claim is generated by . Let be a homogeneous element of degree . Then in we can write for some polynomial and . The assertion now follows by induction on .

Corollary 8.21.

For or , suppose that has EGM. For , assume has EGM. Then is a finitely generated -algebra.

Proof.

In the or cases, apply the theorem with or , which are trivially positive. Then is the degree part of the finitely generated ring with respect to the -grading.

Since is a quotient of by construction of in Theorem 7.16(1), is also finitely generated.

Proposition 8.22.

Assume has EGM. Then for each universal Laurent polynomial on , the function of Theorem 6.8 has finite support (i.e., for all but finitely many , and gives inclusions of -algebras

Proof.

Let be a universal Laurent polynomial on . By Theorem 6.8 the sets of Definition 6.6 are independent of the seed . We claim that for each global monomial on , there is a constant such that

To see that this is sufficient to prove the proposition, note that by Lemma 8.15, there are a finite number of such that the intersection of the sets where is the origin in . Thus, if the claim is true, , the support of , is a finite set. The inclusion of algebras follows by Proposition 6.4. So it is enough to establish the claim.

Let be a global monomial which is a character on the seed torus for . We follow the notation of Definition 6.6. Thus , where itself depends on the seed and . The tropicalization of global monomials which restrict to a character on the seed torus are identified with integer points of the dual cone (i.e., elements nonnegative on each of the initial scattering monomials); see the proof of Proposition 8.13.

Note is linear on . Since is a finite set, for any such , there is constant such that

This completes the proof.

8.4. Conditions implying has EGM and the full Fock–Goncharov conjecture

We begin by showing that some standard conditions in cluster theory, namely acyclicity of the quiver or existence of a maximal green sequence, imply a weaker condition which in turn implies both the EGM condition and the full Fock–Goncharov conjecture. This suggests that this weaker condition is perhaps a more natural one in cluster theory. This point has been explored in Reference Mu15.

Definition 8.23.

We say a cluster variety has large cluster complex if for some seed , is not contained in a half-space.

Proposition 8.24.

Consider the following conditions on a skew-symmetric cluster algebra :

(1)

has an acyclic seed.

(2)

has a seed with a maximal green sequence (for the definition, see Reference BDP, Def. 1.8

(3)

has large cluster complex.

Then implies implies .

Proof.

(1) implies (2) is Reference BDP, Lemma 1.20. For (2) implies (3), let be an initial seed, and let be the seed obtained by mutations in a maximal green sequence. By definition the -vectors for have nonpositive entries. By Lemma 5.12 the -vectors are the equations for the walls of the cluster chamber , viewed as a chamber of . But then . Hence these two chambers coincide and, in particular, has large cluster complex.

Proposition 8.25.

If has large cluster complex, then has EGM, , and the full Fock–Goncharov conjecture (see Definition 0.6) holds for , , very general and, if the convexity condition of Theorem 0.3 holds, for .

Proof.

Assume EGM fails for . Then we have some point such that for all . Take any seed . We can compute by using the corresponding positive Laurent polynomial , for a point in the distinguished chamber of . Thus using Lemma 2.8 (leaving the canonical isomorphism out of the notation),

Here the minimum is over all broken lines contributing to and the final inequality comes from the fact that one of the broken lines is the obvious straight line. Thus is contained in the half-space . Since is the inverse image of under the map , the EGM statement follows. Now since and is saturated and intrinsically closed under addition; see Theorem 7.5. Since satisfies EGM, so does by Proposition 8.16, and the full Fock–Goncharov conjecture for then follows from Corollary 8.18 and Proposition 8.22. The and cases then follow as in the proofs of Corollary 7.13 and Theorem 7.16.

We next give another condition for the EGM condition to hold. While this may appear very technical, it is in fact very important for group-theoretic examples; see Remark 8.28.

Proposition 8.26.
(1)

Let be an affine variety over a field , and let generators of be a -algebra. For each divisorial discrete valuation (where denotes the function field of which does not have center on (or equivalently, for each boundary divisor in any partial compactification , for some .

(2)

Suppose is a cluster variety, is a smooth affine variety, and is an open immersion. Let generate as a -algebra. Then is strictly negative on .

Proof.

(1) Let be an open immersion with complement an irreducible divisor . Suppose each is regular along . Then the inclusion is an equality. Thus the inverse birational map is regular, which implies . Thus (1) follows.

(2) Since the restriction to the open subset is an isomorphism, it follows that has codimension at least 2. Thus itself is log Calabi–Yau by Reference GHK13, Lemma 1.4, and the restriction of the holomorphic volume form is a scalar multiple of . In addition . Now (2) follows from (1).

Proposition 8.27.

If the canonical map

is surjective, then

(1)

is isomorphic to .

(2)

We can choose so that the induced map is an isomorphism.

(3)

The map induced by the choice of in , is finite.

(4)

If furthermore for each we can find a cluster variable with , then (and has EGM. This final condition holds if is finitely generated and is a smooth affine variety.

Proof.

(1) is Lemma B.7. (3) follows from (2). So we assume is surjective, and show that we can choose to have finite cokernel, or equivalently, so is injective. We follow the notation of Reference GHK13, §2.1. By the assumed surjectivity, is injective iff the induced map is injective. We can replace by for any map which vanishes on , i.e., factors through a map . Note by the assumed surjectivity that and have the same rank, and moreover the restriction (which is unaffected by the addition of ) is injective. In particular is injective. Thus we can choose , vanishing on (i.e., factoring through a map ) so that is injective. By viewing the determinant of for an integer as a polynomial in , we see that is injective for all but a finite number of . For sufficiently divisible , extends to under the natural inclusion . Now is injective as required. This shows (2).

For (4), when is a trivial bundle, it follows that

So we have EGM so long as we can find cluster variables on with the given condition. The final statement of (4) follows from Proposition 8.26.

Remark 8.28.

Every double Bruhat cell is an affine variety by Reference BFZ05, Prop. 2.8 and smooth by Reference FZ99, Theorem 1.1. The surjectivity condition in the statement of Proposition 8.27 holds for all double Bruhat cells by Reference BFZ05, Proposition 2.6 (the proposition states that the exchange matrix has full rank, but the proof shows the surjectivity). So by the proposition, has EGM for double Bruhat cells for which the upper and ordinary cluster algebras are the same. This holds for the open double Bruhat cell of and the base affine space ( maximal unipotent) for by Reference BFZ05, Remark 2.20, and is announced in Reference GY13 for all double Bruhat cells of all semisimple .

8.5. Compactifications from positive polytopes

In this subsection, we will use positive polytopes in to create partial compactifications of which fiber over an affine space . The fiber over will be a toric variety, and the general fiber is log Calabi–Yau.

Fix seed data for a cluster variety, let and let be the monoid generated by the . Similarly, let be the cone generated by the . The choice of seed gives an identification and in particular determines a second projection (which depends on the choice of seed). We have the canonical inclusion given in each seed by and canonical translation action of on making into a -module.

Now assume we are given a compact, positive, rationally defined top-dimensional polytope . We let . By Proposition 8.17 and the compactness of , is a -algebra with -algebra structure constants .

Lemma 8.29.

The set is a positive polytope. Denote by the degree part of the ring defined in Theorem 8.19. Then is a finitely generated -algebra.

Proof.

Positivity follows from the fact that for each scattering monomial in . The finite generation statement then follows from Theorem 8.19.

Let and

Then is positive by Proposition 8.10, and as the intersection of two positive sets is positive, is positive. Hence the associated graded rings and (graded by ) defined via Theorem 8.19 are finitely generated. Note that is the set of homogeneous elements of degree in the localization . Thus we have an inclusion of an open subset, with complement the zero locus of . The inclusion of in the degree 0 part of induces a morphism . This morphism is flat, since is a free -module.

Theorem 8.30.

The central fiber of

is the polarized toric variety given by⁠Footnote3 the polyhedron where is the natural map of Proposition B.2.

3

Although is only a rationally defined polyhedron rather than a lattice polyhedron, we can still define .

Proof.

This follows from the multiplication rule. Indeed, since all the scattering monomials project under into the interior of , vanishes modulo the maximal ideal of for any broken line that bends. Thus

with multiplication induced by (addition in ). This is the coordinate ring of

Example 8.31.

Consider the fixed data and seed data given in Example 1.14. The scattering diagram for in this case has three walls, pulled back from the walls of the scattering diagram for as given in Example 1.14, with attached functions , and . Here, with basis of and dual basis of , we have and .

Take to be the pentagon with vertices (with respect to the basis ) , , , , and , which we write as . Then pulls back to to give a polytope . It is easy to see that is a positive polytope. Further, write , . Then it is not difficult to describe the ring determined by as the graded ring generated in degree by , with relations

These equations define a family of projective varieties in , parameterized by . For , we obtain a smooth del Pezzo surface of degree . The boundary (where ) is a cycle of five projective lines. When , we obtain a toric surface with two ordinary double points.

Theorem 8.32.

Assume that has EGM, is given as above, and that is an algebraically closed field of characteristic zero. Let be one of or . We note has a finitely generated -algebra structure by Corollary 8.18. Define .

Define (constructed above) in case , and for , take instead its fiber over (we are not defining in the case), so by construction we have an open immersion . Define . The following hold:

(1)

In all cases is a Gorenstein scheme with trivial dualizing sheaf.

(2)

For , , or for general, is a -trivial Gorenstein log canonical variety.

(3)

For or for general, or all assuming there exists a seed and a strictly convex cone containing all of for , we have is a minimal model. In other words, is a (in the case relative to projective normal variety, is a reduced Weil divisor, is trivial, and is log canonical.

Proof.

First we consider the theorem in the cases . Note that (3) implies (2) by restriction.

We consider the family constructed above, where is the divisor given by with its reduced structure. Using Lemma 8.33, the condition that on a fiber , is a minimal model (in the sense of the statement) is open, and it holds for the central fiber as it is toric by Theorem 8.30. Thus the condition holds for fibers over some nonempty Zariski open subset . This gives (3) for with general. The convexity condition (on the ) implies there is a one-parameter subgroup of which pushes a general point of to (see the proof of Theorem 7.16), and now (3) for for all follows by the -equivariance.

Now note given seed data the convexity assumption holds for the seed data . Thus the final paragraph applies with and so in particular is Gorenstein with trivial dualizing sheaf. The same then holds for the fibers of the flat map , which are (for arbitrary ). This gives (1).

Finally, we consider the case . The graded ring construction above applied with seed data gives a degeneration of a compactification of (which is now a fiber of the family) to a toric compactification of . The torus acts on the family, trivially on the base, and the quotient gives an isotrivial degeneration of an analogously defined compactification of to a toric compactification of . We leave the details of the construction (which is exactly analogous to the construction of above) to the reader. Now exactly the same openness argument applies.

We learned of the following result, and its proof, from J. Kollár.

Lemma 8.33 (Kollár).

Let be an algebraically closed field of characteristic . Let be a proper flat morphism of schemes of finite type over , and let be a closed subscheme which is flat over . Let denote the fiber of over a closed point . Assume that is regular and for the following hold:

(1)

is normal and Cohen–Macaulay.

(2)

is a reduced divisor.

(3)

The pair is log canonical.

(4)

.

(5)

Then the natural morphism is an isomorphism, and there exists a Zariski open neighborhood such that the conditions hold for all . In particular, is a -trivial Gorenstein log canonical variety for all .

Proof.

We are free to replace by an open neighborhood of and will do so during the proof without further comment.

By assumption and is Cohen–Macaulay. So

is Cohen–Macaulay by Reference K13, Corollary 2.71, p. 82. It follows that is Cohen–Macaulay by Reference K13, Corollary 2.63, p. 80.

The base is regular by assumption, so is cut out by a regular sequence. Since and are Cohen–Macaulay and is proper and flat, we may assume that and are Cohen–Macaulay. Now is Cohen–Macaulay by Reference K13, Corollary 2.63, and is Cohen–Macaulay by Reference K13, Corollary 2.71. The relative dualizing sheaf is identified with , so is also Cohen–Macaulay. It follows that is Cohen–Macaulay, and so in particular it satisfies Serre’s condition . The natural map is an isomorphism in codimension (because is smooth in codimension ) and both sheaves are , hence the map is an isomorphism. Now implies that we may assume using .

The conditions (1), (2), and (5) are open conditions on because is proper and flat. So we may assume they hold for all . We established above that is invertible. It follows that condition (3) is also open on by Reference K13, Corollary 4.10, p. 159, and that condition (4) is open on (using (5)).

Remark 8.34.

Note that directly from its definition, with the multiplication rule counting broken lines, it is difficult to prove anything about , e.g., that it is an integral domain or to determine its dimension. But the convexity, i.e., existence of a convex polytope in the intrinsic sense, gives this very simple degeneration from which we get many properties, at least for very general , for free.

There have been many constructions of degenerations of flag varieties and the like to toric varieties; see Reference AB and references therein. We expect these are all instances of Theorem 8.30.

Many authors have looked for a nice compactification of the moduli space of (say) rank 2 vector bundles with algebraic connection on an algebraic curve . We know of no satisfactory solution. For example, in Reference IIS the case of the complement of four points in is considered, a compactification is constructed, but the boundary is rather nasty (it lies in , but this anticanonical divisor is not reduced). This can be explained as follows: has a different algebraic structure, the character variety, (as complex manifolds they are the same). Note is covered by affine lines (the space of connections on a fixed bundle is an affine space), thus it is not log Calabi–Yau. Rather, it is the log version of uniruled, and there is no Mori theoretic reason to expect a natural compactification. however is log Calabi–Yau, and then by Mori theory one expects (infinitely many) nice compactifications, the minimal models; see Reference GHK13, §1, for an introduction to these ideas. When has punctures, is a cluster variety; see Reference FST and Reference FG06. In the case of with four punctures, is the universal family of affine cubic surfaces (the complement of a triangle of lines on a cubic surface in ); see Reference GHK11, Example 6.12. Each affine cubic has an obvious normal crossing minimal model, the cubic surface. This compactification is an instance of the above, for a natural choice of polygon . The same procedure will give a minimal model compactification for any character variety (of a punctured Riemann surface) by the above simple procedure that has nothing to do with Teichmüller theory.

For the remainder of this section we will assume that has EGM. By Lemma 8.15, there are global monomials with such that is min-convex with

being compact. Thus we have seeds (possibly repeated) such that is a character on , so that is linear after making the identification . Furthermore, as in the proof of Theorem 8.19, we can assume is in the interior of the cone . We will now observe that with these assumptions the irreducible components of the boundary in the compactification of induced by are toric.

Note for each there is at least one seed where is linear. We assume the collection of is minimal for defining , and thus is a union of maximal faces of , a nonempty closed subset of codimension .

Writing , let be the graded algebra of Theorem 8.19, again a finitely generated algebra. Then is a projective variety and gives a Cartier (but not necessarily reduced) boundary .

Theorem 8.35.

In the above situation, the irreducible components of are projective toric varieties. More precisely, for each we have a seed such that is a character on . Then

is a bounded polytope. The associated projective toric variety is an irreducible component of , and all irreducible components of occur in this way.

Proof.

For each consider the vector subspace with basis with and for .

Note that is an ideal of . Indeed, the fact that lies in the interior of its cone of the cluster complex for implies by Proposition 8.13 that is strictly increasing at bends on monomial decorations of broken lines. Now if , , and appears in , then , and thus .

Now the definitions imply . So it is enough to show that is the projective toric variety given by the polytope . Now has basis , . By the multiplication rule, and the fact again that is strictly increasing at bends on monomial decorations of broken lines, the only broken line that contributes to is the straight broken line, and the multiplication rule on

is given by lattice addition, i.e., is the projective toric variety given by the polytope .

Remark 8.36.

The result is (at least to us) surprising in that many cluster varieties come with a natural compactification, where the boundary is not at all toric. For example, order the columns of a matrix and consider the open subset where the consecutive Plücker coordinates (the determinant of the first columns, columns , etc.) are nonzero. This is a cluster variety. Its boundary in the given compactification is a union of Schubert cells (which are not toric). This has EGM by Proposition 8.27. Then generic compactifications given by bounded polytopes gives an alternative compactification in which we replace all these Schubert cells by toric varieties. We do not know, e.g., how to produce such a compactification by birational geometric operations beginning with .

9. Partial compactifications and representation-theoretic results

9.1. Partial minimal models

As discussed in the introduction, many basic objects in representation theory, e.g., a semisimple group , are not log Calabi–Yau, and we cannot expect that they have a canonical basis of regular functions. However, in many cases the basic object is a partial minimal model of a log Calabi–Yau variety, i.e., contains a Zariski open log Calabi–Yau subset whose volume form has a pole along all components of the complement. For example, the group will be a partial compactification of an open double Bruhat cell, and this is a partial minimal model. We have a canonical basis of functions on the cluster variety, and from this, we conjecture one can get a canonical basis on the partial compactification (the thing we really care about) in the most naive possible way, by taking those elements in the basis of functions for the open set which extend to regular functions on the compactification. We are only able to prove the conjecture under rather strong assumptions; see Corollary 9.17. Happily these conditions hold in many important examples.

Note that a frozen variable for (or ) canonically determines a valuation, a point of , namely the boundary divisor where that variable becomes zero. See Construction B.9.

While we have myriad (and near optimal) sufficient conditions guaranteeing a canonical basis for , we can only prove our conjecture that is a basis of for a partial minimal model under a much stronger condition (which happily holds in the most important representation theoretic examples):

Definition 9.1.

We say a seed is optimized for if

where

is the composition of canonical identifications defined in §2. If instead , we say is optimized for if for all .

We say is optimized for a frozen index if it is optimized for the corresponding point of .

The reason for defining this notion for both and despite the canonical identification between these two sets is that it is sometimes convenient to think of as specified by a boundary divisor.

For the connection between optimal seeds and our conjecture on see Proposition 9.7 and Conjecture 9.8.

Lemma 9.2.

In the skew-symmetric case, a seed is optimized for a frozen index if and only if in the quiver for this seed all arrows between unfrozen vertices and the given frozen vertex point toward the given frozen vertex.

Proof.

Under the identification (which is just in the skew-symmetric case), the valuation corresponding to the divisor given by the frozen variable indexed by is simply . Thus the seed is optimized for this frozen variable if for all ; this is the number of arrows from to in the quiver, with sign telling us that they are incoming arrows.

Lemma 9.3.
(1)

The seed is optimized for if and only if the monomial on is a global monomial. In this case

and the global monomial is the restriction to of . In the case, for primitive, this holds if and only if each of the initial scattering monomials in is regular along the boundary divisor of corresponding to under the identification .

(2)

has an optimized seed if and only if lies in .

Proof.

For (1), the rays for the fan giving the toric model for are for . Note that ; see Equation 2.6. Now the statement concerning follows from Lemmas 7.8 and 7.10. The additional statement in the case is clear from the definitions. For (2), one notes that the forward implication is given by (1), while for the converse, if , then for some seed , and then is optimized for that seed.

Proposition 9.4.

For the standard cluster algebra structure on (the affine cone over in its Plücker embedding) every frozen variable has an optimized seed.

Proof.

As was pointed out to us by Lauren Willams, for , the initial seed in Reference GSV, Figure 4.4, is optimized for one frozen variable (the special upper right hand vertex for the initial quiver). The result follows from the cyclic symmetry of this cluster structure.

Remark 9.5.

B. LeClerc, and independently L. Shen, gave us an explicit sequence of mutations that shows the proposition holds as well for the cluster structure of Reference BFZ05, Reference GLS on the maximal unipotent subgroup , and the same argument applies to the Fock–Goncharov cluster structure on , . The argument appears in Reference Ma17.

Lemma 9.6.

Let be a lattice, and let be a submonoid with . For any subset and collection of elements such that , the subset is linearly independent over .

Proof.

Suppose

for all nonzero, and suppose is a nonempty finite set. Let be minimal with respect to the partial ordering on given by (where means for some ). The coefficient of in the sum, expressed in the basis of monomials, must be zero. But the minimality of implies the monomial does not appear in any of the , . Thus the coefficient of is just , a contradiction.

Proposition 9.7.

Suppose a valuation has an optimized seed. If , then for all with .

Proof.

Let be optimized for . Let be the strictly convex cone spanned by the exponents of the initial scattering monomials . Let . Take a basepoint in the positive chamber of . For each , by definition where is a finite sum of monomials. By Lemma 9.3(1) we have , and thus by Lemma 2.8(2), .

Let be the minimum of over all with , and suppose . Since , necessarily

Note this is the sum of all the monomial terms in which have the maximal order of pole, , along . This contradicts Lemma 9.6.

We believe the assumption of an optimized seed is not necessary:

Conjecture 9.8.

The proposition holds for any .

Any finite set of primitive elements gives a partial compactification (defined canonically up to codimension 2) , with the boundary divisors of this partial compactification in one-to-one correspondence with the elements of (This is true for any finite collection of divisorial discrete valuations on the function field of a normal variety : there is always an open immersion , with divisorial boundary corresponding to , and is unique up to changes in codimension greater than or equal to 2.)

We then define

and the vector subspace with basis . Similarly, we define to be the subalgebra of generated by those cluster variables that are regular (generically) along all .

Definition 9.9.

Each choice of seed gives a pairing

which is just the dual pairing composed with the identifications

Lemma 9.10.
(1)

The subspace is a subalgebra containing . If then

(2)

Assume each has an optimized seed. Then

If then .

(3)

If each has an optimized seed and is optimized for , the piecewise linear function

is min-convex, and for all ,

where is the global monomial on corresponding to (which exists by Lemma 9.3).

Remark 9.11.

There are pairings

which are much more natural than Definition 9.9. Indeed, gives a canonical function and, since , a valuation on . The analogous statements apply to . So we could define a pairing by either

We conjecture that these two pairings are equal. Lemma 9.10(3) gives the result when one of lies in the cluster complex. One can pose the same symmetry conjecture for mirror pairs of affine log Calabi–Yau varieties (with maximal boundary) in general, the two-dimensional case having been shown in Reference M14. Suppose the symmetry conjecture holds, and furthermore . Then (see the proof of Lemma 9.10) the cone of Equation 0.18 cut out by the tropicalization of the potential function is

If furthermore Conjecture 9.8 holds and , then gives a basis of , canonically determined by the open set (together with its cluster structure, though we conjecture the basis is independent of the cluster structure); see Corollary 9.17.

Proof of Lemma 9.10.

The subalgebra statement of (1) follows from the positivity (both of structure constants and the Laurent polynomials ) just as in the proof of Theorem 7.5. Every cluster variable is a theta function, so the inclusion is clear. Now obviously if then both are equal to .

The intersection expression of (2) for the middle algebra follows from Proposition 9.7. Now obviously if then .

For (3) we work with the scattering diagram . Then is the -vector of the global monomial , with , by Lemma 9.3. Using , one sees that is linear on , so obviously min-convex in the sense of Definition-Lemma 8.1. Since it is the tropicalization of a global monomial, it is also min-convex in the sense of Definition 8.2 by Proposition 8.13.

Now fix a basepoint , and consider , . By Lemma 9.3(1) each scattering function is regular along the boundary divisor corresponding to . By definition , where is a linear combination of monomials with regular along the boundary divisor corresponding to . Thus

by Lemma 2.8. Since is the monomial on ,

This completes the proof of (3).

9.2. Cones cut out by the tropicalized potential

Recall a choice of seed gives a partial compactification and a map . The boundary has irreducible components, primitive elements of , the vanishing loci of the .

Lemma 9.12.

The seed is optimized for each of the boundary divisors of .

Proof.

If , the corresponding seed for is

and the boundary divisors correspond to the . But hence the claim.

We adjust slightly the notation of the previous subsection to this case:

Definition 9.13.

Let

be the subset of points such that remains regular on the partial compactification , i.e., such that

Lemma 9.14.

Under the identification , we have and .

Proof.

is invariant under translation by , and thus .

By Lemma 5.2 we construct from the atlas of toric compactifications

parameterized by the cluster chambers in . Now take , and consider for some basepoint in the cluster complex. This is a positive sum of monomials, so it will be regular on the boundary of iff each summand is. One summand is , so if is regular on then . Thus . But now suppose , some and . Then is regular on the boundary. Since the initial scattering monomials are , any bend in a broken line multiplies the decorating monomial by a monomial regular on the boundary. Thus . This completes the proof.

We define

Recall are the cluster and upper cluster algebras with principal coefficients, respectively, with the frozen variables inverted. On the other hand, are the cluster and upper cluster algebras with principal coefficients, respectively, with the frozen variables not inverted. By Lemma 9.10, is a subalgebra, and .

By Lemmas 9.10 and 9.12, we have

Corollary 9.15.

If then .

Here is another sufficient condition for the full Fock–Goncharov conjecture to hold, which will prove immediately useful below:

Proposition 9.16.

Suppose there is a min-convex function , such that implies for integral, and such that for some . Suppose also that there is a bounded positive polytope in (which holds, for example, if has EGM). Then .

Proof.

Take any and with . Then consider any appearing in , for . By Lemma 8.4

for sufficiently large. In particular , so for each that appears, is a universal positive Laurent polynomial, for any basepoint in the cluster complex. The existence of the bounded positive polytope implies is a finite sum of . Thus the product is also a universal positive Laurent polynomial, and thus by the positivity of the scattering diagram, must be a finite positive Laurent polynomial. Thus .

If there are frozen variables, there is a canonical candidate for in the proposition. When we have frozen variables, this gives a partial compactification . In this case, let us change notation slightly and write a seed as

with and , and the are frozen. In this case the elements give canonical boundary divisors for a partial compactification , and an analogous . An atlas for is given by gluing the partial compactification , where is the fan consisting of the rays .

Corollary 9.17.

Assume that for each , has an optimized seed, . Let be the (Landau–Ginzburg) potential, the sum of the corresponding global monomials on given by Lemma 9.3. Then:

(1)

The piecewise linear function

is min-convex and

is a positive polytope.

(2)

has the alternative description

(3)

The set

parameterizes a canonical basis of

Proof.

This is immediate from Lemma 9.10.

Corollary 9.18.

Assume we have EGM on , and every frozen variable has an optimized seed. Let and be as in Corollary 9.17. If for some seed , is contained in the convex hull of (which itself contains the integral points of the cluster complex then , is finitely generated, and the integer points parameterize a canonical basis of .

Proof.

By definition , and is min-convex by Lemma 9.3. Thus by Proposition 9.16. Now the result follows from the inclusions

of Proposition 8.22.

The corollary applies in important representation-theoretic examples:

Proof of Corollary 0.21.

We check the conditions of Theorem 0.19: The existence of an optimized seed is proven in Reference Ma15, following suggestions of L. Shen and B. LeClerc. The cluster variety has large cluster complex (Definition 8.23) by Reference GS16, Theorems 1.12 and 1.17. This gives the hypotheses of Theorem 0.19(3). The equality of our with the Goncharov–Shen potential is given in Reference Ma17, Theorem 2. It is shown in Reference Ma17, Proposition 22 that the exchange matrix has full rank, in the sense of Lemma B.7. Now the results imply the analogous statements for as in the proof directly above. Magee identifies the action with the -action in Reference Ma17, §4.c, which gives the weight statements as above.

Proof of Corollary 0.20.

The hyptheses of Theorem 0.19 are proven in Reference Ma15, using Proposition 9.16 applied to the tropicalization of our potential . The agreement of with the Berenstein–Kazhdan potential is given in Reference Ma17, §5. Theorem 0.19 is stated for . But in this case it is shown in Reference Ma15 that the exchange matrix has full rank, i.e., the equivalent conditions of Lemma B.7 hold. Now the results for imply the analogous result for using equivariance, as in the proof of Theorem 7.16(7). Magee identifies the -action with the action of on , and the various statements about -weights now follow immediately from the equivariance, Proposition 7.7.

10. Links with quiver representations and work of Reineke

We briefly make a connection with work of Reineke Reference R10 in the acyclic skew-symmetric case. In this case has a natural interpretation in terms of moduli of quiver representations. Consider skew-symmetric fixed and initial data with no frozen variables. Set , and let be the completion of the polynomial ring with respect to the maximal monomial ideal. Let be a monoid as in §1 containing all , so that , the pronilpotent group of §1.1 (in the case), acts by automorphisms of as usual. Note there is an embedding given by . The action of on then induces an action on . Indeed, one checks immediately that an automorphism (for )

induces the automorphism on given by

Proposition 10.1.

Suppose we are given fixed skew-symmetric data with no frozen variables along with an acyclic seed . Let be the associated quiver.⁠Footnote4 Each gives a stability in the sense of Reference R10. Assume there is a unique primitive with . For each let

4

Note that because of the assumption made in Appendix A that for any , the quiver has no isolated vertex.

where is the framed moduli space (framed by the vector spaces with unless , in which case of semistable representations of with dimension vector and -slope (see Reference R10, §5.1), and denotes topological Euler characteristic. Let for some . Then

depends only on and (i.e., is independent of the vertex . Furthermore, for arbitrary , (see Lemma 1.9) acts on by

and on by

Proof.

The equality of the follows from the argument in the proof of Reference R10, Lemma 3.6.

If for some then one checks easily that is a point and for or . Thus and the formula for holds by Remark 1.29.

Let be the pronilpotent group of §1.1 (in the case) associated to the completion of the Lie algebra

The group acts faithfully on but need not act faithfully via restriction on . It turns out, however, that there is a subgroup which does act faithfully on and that all automorphisms attached to walls in are elements of . We see this as follows.

Consider the subspace

By the commutator formula (Equation 1.1) we have , and in particular is a Lie subalgebra of . Let denote the associated pronilpotent subgroup. Because of the assumption that for any unfrozen , all automorphisms associated to initial walls of lie in , and thus by the inductive construction of the scattering diagram in §C.1, all automorphisms associated to outgoing walls of also lie in .

Let denote the pronilpotent group acting faithfully on associated to the completion of the Lie algebra

Then the restriction of the action of on to is given by the group homomorphism associated to the Lie algebra homomorphism

This homomorphism restricts to an isomorphism , and in particular the restriction of the action on to is faithful.

Assume now that the indices are ordered so that has arrows from the vertex with index to the vertex with index only if . We compute , the automorphism associated to a path from the positive to the negative chamber, in two different ways.

First, there is a sequence of chambers connecting to via the mutations , . Indeed, it is easy to check that the -vectors obtained by mutating are precisely , and the chamber corresponding to this sequence of mutations is precisely the dual of the cone generated by the -vectors; see Lemma 5.12. Thus in particular, we can find a path from to which only crosses the walls in order. Note that the element of attached to the wall acts on by , which agrees with the automorphism in Reference R10 written as (noting that Reference R10 uses the opposite sign convention for the skew form associated to the quiver). From this we conclude that , the left-hand side of the equality of Theorem 2.1 of Reference R10.

On the other hand, choose a stability condition and consider the path from to parameterized by , with , with domain sufficiently large so the initial and final endpoints lie in the positive and negative orthants, respectively. Then a dimension vector has -slope if and only if it has -slope . Thus if the description in the statement of the theorem of is correct, then coincides with the right-hand side of the equality of Theorem 2.1 of Reference R10. By the uniqueness of the factorization of from the proof of Theorem 1.17, and the faithful action of on shown above, we obtain the result.

Because nonnegativity of Euler characteristics for the quiver moduli spaces appearing in the above statement is known (see Reference R14), this gives an alternate proof of positivity of the scattering diagram in this case.

Remark 10.2.

Since the initial version of this paper was released, Bridgeland Reference Bri developed a Hall algebra version of scattering diagrams in the context of quiver representation theory, and the above result follows conceptually from his results.

Example 10.3.

Let be a quiver given by an orientation of the Dynkin diagram of a simply laced finite-dimensional simple Lie algebra. Then the dimension vectors of the indecomposable complex representations of are the positive roots of the associated root system (Gabriel’s theorem). Moreover, for each positive root , there is a unique indecomposable representation with dimension vector , and ; see, e.g., Reference BGP73.

The cluster variety associated to is the cluster variety of finite type associated to the root system Reference FZ03a. Using Proposition 10.1, we can give an explicit description of the scattering diagram for as follows.

First we observe that a representation of that contributes to is a direct sum of copies of an indecomposable representation. Let be a primitive vector, and let be such that . Suppose is an -semistable representation of with dimension vector a multiple of , and consider the decomposition of into indecomposable representations. By -semistability and our assumption , each factor must have dimension vector a multiple of . By Gabriel’s theorem, we see that is a positive root and is a direct sum of copies of the associated indecomposable representation.

We see that the walls of are in bijection with the positive roots of . Let be a positive root, and let be the indecomposable representation with dimension vector . Let be the locus of such that is -semistable of -slope zero; that is, and for the dimension vector of any subrepresentation of . Then is a rational polyhedral cone in , and is nonempty of real codimension . Indeed, there exists such that is -stable by Reference K94, Remark 4.5 and Reference S92, Theorem 6.1, and this is an open condition on . Now let be a point such that . Then the -semistable representations of -slope zero are the direct sums of copies of .

Let us now examine the moduli space . An object in this moduli space is a direct sum of copies of the unique indecomposable representation of dimension vector , along with the framing, a choice of a vector . Such an object is stable if and only if is not contained in a proper subrepresentation of of the form for some subspace . In order for this to be the case, the must be linearly independent elements of , and hence span a -dimensional subspace of . The automorphism group of is , which has the effect of changing the basis of the subspace spanned by . Now it follows easily from the definitions that, for each and such that , the moduli space of -semistable representations with framing at vertex is isomorphic to the Grassmannian .

So, in the notation of Proposition 10.1,

and

Thus the wall of associated to is

For example, suppose is the quiver with vertices , and arrows from 1 to 2 and 2 to 3. This is an orientation of the Dynkin diagram . We have the following isomorphism types of indecomposable representations:

(Here the numbers denote the dimension of the vector space at the vertex, and the symbol over an arrow indicates that the corresponding linear transformation is an isomorphism.) We write and . Then the walls of are

For example, the indecomposable representation with dimension vector has subrepresentations with dimension vectors and . So the associated wall has support defined by the inequalities and . This gives the last wall in the list.

Example 10.4.

Kac generalized Gabriel’s theorem to the case of an arbitrary quiver (without edge loops) as follows (see Reference K80, Reference K82): Let be the Kac–Moody algebra associated to the underlying graph of . Then the dimension vectors of indecomposable complex representations of are the positive roots of .

The roots of are divided into real and imaginary roots. The real roots are the translates of the simple roots under the action of the Weyl group. Let be the asymmetric bilinear form defined by

Then for representations and of with dimension vectors and ,

Let be a positive root. We have if is real and if is imaginary. The indecomposable representations of dimension vector depend on parameters. We say is a Schur root if there exists a representation of with dimension vector such that .

Assume that is acyclic. Let be the scattering diagram for the associated cluster variety. We study using Proposition 10.1.

We show that each wall of is contained in for a primitive Schur root. First, as in Example 10.3, each wall is contained in for a primitive positive root (note that the set of roots is saturated by Reference K80, Proposition 1.2). It remains to show that is necessarily a Schur root. Otherwise, a representation with dimension vector deforms to a decomposable representation Reference K82, Proposition 1(b). Then, for such that , is -unstable, and so is -unstable (as -semistability is an open condition). A representation with dimension vector a multiple of is -unstable for the same reason, using Reference S92, Theorem 3.8. Now by Proposition 10.1 we see that there does not exist a wall of contained in .

For a real Schur root there is a unique wall contained in which can be described explicitly as in Example 10.3. We remark that is a real Schur root iff there is an indecomposable representation of with dimension vector such that and . (Moreover, is uniquely determined by .) Such a representation is an exceptional object in the category of representations of in the sense of Reference B90.

For an imaginary Schur root the associated walls involve contributions from positive-dimensional moduli spaces of semistable representations of . The case of the imaginary root for the quiver with vertices and two arrows from 1 to 2 is described in Reference R10, §6.1. (This is the case of Example 1.15.)

Appendix A. Review of notation and Langlands duality

We first review basic cluster variety notation as adopted in Reference GHK13. None of this is original to Reference GHK13, but we follow that source for consistency of notation.

As in Reference GHK13, §2, fixed data means

A lattice with a skew-symmetric bilinear form

An unfrozen sublattice , a saturated sublattice of . If , we say the fixed data has no frozen variables.

An index set with and a subset with .

Positive integers for with greatest common divisor .

A sublattice of finite index such that and .

, .

Here we modify the definition slightly, and include in the fixed data a mutation class of seed. Recall a seed is a basis of satisfying certain properties; see Reference GHK13, §2, for the precise definitions, including that of mutation. In particular, we write for the dual basis and . We write

We shall also assume that if , then the linear functional is nonzero. (If this happens, then one can view as frozen.)

We have two natural maps defined by :

We also choose a map

such that (a) and (b) the composed map agrees with . Different choices of differ by a choice of map . Further, if there are no frozen variables, i.e., , then is canonically defined.

We also define

Following our conventions in Reference GHK13, let be the infinite oriented rooted tree with outgoing edges from each vertex, labeled by the elements of . Let be the root of the tree. Attach some choice of initial seed to the vertex . (We write if we want to record this choice of initial seed.) Now each simple path starting at determines a sequence of mutations, just mutating at the label attached to the edge. In this way we attach a seed to each vertex of . We write the seed attached to a vertex as , and write etc., for the corresponding tori. Mutations define birational maps between these tori, and the associated Fock–Goncharov , cluster varieties are defined by

This parameterization of torus charts is very redundant, with infinitely many copies of the same chart appearing. In particular, given a vertex of , one can consider the subtree rooted at , with initial seed . This tree can similarly be used to define , and the obvious inclusion between these two versions of is in fact an isomorphism, as can be easily checked.

As one expects the mirror of a variety obtained by gluing charts of the form to be obtained by gluing charts of the form , the mirror of is not , as the latter is obtained by gluing charts of the form . To get the correct mirrors of and , one follows Reference FG09 in defining the Langlands dual cluster varieties. This is done by, given fixed data , defining fixed data to be the fixed data

where

The lattice, with its finite index sublattice, is

and the -valued skew-symmetric form on is

For each , we define

One checks easily that gives a bijection between and .

Note that for skew-symmetric cluster algebras, i.e., when all the multipliers , Langlands duality is the identity, .

Definition A.4 (Fock–Goncharov dual).

We write and .

Note in the skew-symmetric case, that .

One observes the elementary

Proposition A.5.

Given fixed data , the double Langlands dual data is canonically isomorphic to the data via the map given by .

Appendix B. The and -varieties with principal coefficients

We recall briefly the construction of principal fixed data from Reference GHK13, Construction 2.11. For fixed data , the data for the cluster variety with principal coefficients is defined by:

with the skew-symmetric bilinear form

The sublattice is .

The index set is now the disjoint union of two copies of , with the taken to be as in . The set of unfrozen indices is just the original thought of as a subset of the first copy of .

Given an initial seed , we define

We then take the mutation class .

Note that depends on the choice of : it is not true that if is obtained by mutation from then is obtained from the same set of mutations applied to . Nevertheless, the cluster varieties

are defined independently of the seed . This is a very important point, which we shall revisit in Remark B.8.

The following summarizes all of the important relationships between the various varieties which will be made use of in this paper.

Proposition B.2.

Giving fixed data , we have:

(1)

There is a commutative diagram where the dotted arrows are only present if there are no frozen variables (i.e.,

with any point, the identity, and with the left- and right-hand squares cartesian and an isomorphism, canonical if there are no frozen variables.

(2)

There are torus actions

Here is the kernel of the map

Furthermore and act on so that the map is - and -equivariant. The map is a -torsor. There is a map such that the map is also compatible with the actions of these two tori on and respectively, so that

is a -torsor.

(3)

and are isomorphic data, so we can define

(4)

There is a commutative diagram:

Proof.

We consider the diagram of (1). The maps with names are given as follows on cocharacter lattices:

Note is the transpose of the inclusion . In the case there are no frozen variables, the two dotted horizontal lines are just given on cocharacter lattices by again. One checks commutativity from these formulas at the level of individual tori, and one checks the maps are compatible with mutations. Note the left-hand diagram defines ; see Reference GHK13, Definition 2.12. The statements that , and are compatible with mutations are in Reference GHK13, §2, as well as the commutativity of the second square in case of no frozen variables. It is clear that induces an isomorphism of lattices, hence an isomorphism of the relevant tori. This isomorphism is canonical in the no frozen variable case because is well-defined in this case. The fact the right-hand square is cartesian follows from the fact that . Note the signs in the definition of are necessary to be compatible with mutations. This gives (1).

For (2), the first action is specified on the level of cocharacter lattices by

while the last three are given by the inclusions

One checks easily that the induced actions are compatible with mutations. The action of and on are induced by the maps and respectively, in order to achieve the desired equivariance. The map is given by

The other statements are easily checked.

For (3), from the definitions, the lattices playing the role of are

These are isomorphic under the map . Furthermore, the pairings in the two cases are given by

respectively. The isomorphism given preserves the pairings, hence the isomorphism.

(4) is the same as (1), but for the Langlands dual data . For reference, the maps are given as follows:

Remark B.5.

Whenever the lattice appears in dealing with the Langlands dual data, we will always identify this with in the obvious way.

Simple linear algebra gives:

Lemma B.6.

The choice of the map gives an inclusion (see Proposition B.2 given by . We also have (a sublattice of included in via . These inclusions induce an isomorphism .

Lemma B.7.

The map induced by the composition of the inclusion and projection is a split surjection if and only if the map

is surjective. This holds if and only if in some seed , the matrix with entries for , , gives a surjective map . In this case is isomorphic to the trivial bundle .

Proof.

For the first statement, note using Lemma B.6 that the map is surjective if and only if the map given by is surjective, and this is the case if and only if the induced map is surjective. The given matrix is the matrix for in the given bases, so the second equivalence is clear. The final statement follows from the -equivariance of (the trivialization then comes by choosing a splitting of ).

Remark B.8.

In general, a seed is defined to be a basis of the lattice (or ), but to define the seed mutations Reference GHK13, (2.2) and the union of tori Equation A.3, all one needs are elements , (the definitions as given make sense even if the are dependent or fail to span). If one makes the construction in this greater generality, the characters on will not be independent (if the are not), and unless we take a full basis, we cannot define the cluster variables on , as the are defined as the dual basis to the basis for .

In the case of the principal data, given a seed for , we get a seed in this modified sense for the data . We also write this seed as . On the other hand, in Reference GHK13, the seed for is defined in the more traditional sense to be the basis . It is not the case that if is obtained from via a sequence of mutations, then is obtained from by the same sequence of mutations. In particular, the set of seeds mutation equivalent to depends not just on the mutation equivalence class of , but on the original seed . However, using the seed as a seed for in this modified sense, we can build , and this depends only on the mutation class of . Thus does not depend on the initial choice of seed, but only on its mutation equivalence class.

However, as we shall now see, the choice of initial seed does give a partial compactification. This is a more general phenomenon when there are frozen variables.

Construction B.9 (Partial compactifications from frozen variables).

When the cluster data includes frozen variables, comes with a canonical partial compactification , given by partially compactifying each torus chart via , where for , . Thus the dual cone is cut out by the half-spaces , . Note that the monomials , are invariant under mutation. These give a canonical map , where is the number of unfrozen variables. Note that the basis elements for , though they have frozen indices, can change under mutation. What is invariant is the associated boundary divisor with valuation given by . These are the boundary divisors of . We remark that like , is also separated, with the argument given in Reference GHK13, Theorem 3.14 working equally well for .

Here is another way of seeing the same thing. Given any cluster variety and a single fan for a toric partial compactification for some , there is a canonical way to build a partial compactification

We let and , where is the birational map given by the composition

and is the geometric tropicalization; see §2.

Remark B.10.

We now return to the discussion of . Note that the frozen variables for are indexed by in the first copy of , along with all indices in the second copy of . However, we can apply Construction B.9 taking only the second copy of as the set of frozen indices, with the initial choice of seed determining a partial compactification of . In this case, we indicate the partial compactification by . It is important to keep in mind the dependence on . Fixing fixes , and hence cluster variables , . The variables can then take the value in the compactification. In particular, we obtain an extension of to , pulling back to .

Note that the seeds in and are in one-to-one correspondence. Given any seed and seed obtained via the same sequence of mutations, we have for some . These two seeds give rise to coordinates on the chart of indexed by and coordinates on the chart of indexed by . As is the fiber of over the point of with all coordinates , the coordinate on the chart of restricts to the coordinate on the chart of . This gives a one-to-one correspondence between cluster variables on and -type cluster variables on . To summarize:

Proposition B.11.

The cluster variety depends only on the mutation class . But the choice of a seed determines:

(1)

a partial compactification

(2)

the canonical extension of each cluster variable on any chart of to a cluster variable on the corresponding chart of .

Appendix C. Construction of scattering diagrams

This appendix is devoted to giving proofs of Theorems 1.28 and 1.13. The proof of 1.28 is essentially given in Reference GS11, but the special case here is considerably simpler than the general case covered there, and it is likely to be very difficult for the reader to extract the needed results from Reference GS11. In addition, the details of the proof of Theorem 1.28 will be helpful in proving Theorem 1.13.

C.1. An algorithmic construction of scattering diagrams

Construction C.1.

There is a simple order by order algorithm, introduced in Reference KS06 in the two-dimensional case and in Reference GS11 in the higher-dimensional case, for producing the diagram of Theorem 1.21, which we will describe shortly after a bit of preparation. This is useful both from a computational point of view and because a more complicated version of this will be necessary in the remainder of this appendix.

We continue with fixed data , yielding the Lie algebra in §1.1.

We first introduce some additional terminology. For any scattering diagram for , and any we let be the (by definition, finite) set of with nontrivial in . A scattering diagram for , induces a scattering diagram for , in the obvious way, viewing for a wall . We say two scattering diagrams , are equivalent to order if they are equivalent as scattering diagrams for .

Definition-Lemma C.2.

Let be a joint of the scattering diagram . Either every wall containing has direction tangent to (where the direction of a wall contained in is or every wall containing has direction not tangent to . In the first case we call the joint parallel; in the second case, perpendicular.

Proof.

Suppose spans the subspace . Then the direction of any wall containing is of the form for some . If this is tangent to , then for , and hence . From this it follows that for all , and hence the direction of any wall containing is tangent to .

A joint is a codimension 2 convex rational polyhedral cone. Let be the set of integral tangent vectors to . This is a saturated sublattice of . Then we set

This is closed under Lie bracket. If is a parallel joint, then is abelian, since if with , , so . We denote by the corresponding group.

We will build a sequence of finite scattering diagrams , with the property that is equivalent to to order . Taking , we obtain equivalent to . Let denote the subset of consisting of walls which are nontrivial in . We start with

If is a joint of a finite scattering diagram, we write for a simple loop around small enough so that it only intersects walls containing . In particular, for each joint of , . Indeed, is abelian and by the form given for in the statement of Theorem 1.21, all walls containing are hyperplanes. Thus the automorphism associated to crossing each wall and its inverse occur once in , and hence cancel.

Now suppose we have constructed . For every perpendicular joint of , we can write uniquely in

where and . Such an expression holds because all wall-crossing automorphisms for walls containing lie in , so that can be viewed as an element of . Furthermore, by the inductive hypothesis, this element is trivial in . Because is perpendicular, we never have . Now define

where the sign is chosen so that the contribution to crossing the wall indexed by in is . Note the latter element is central in . Thus and

in .

We define

where the union is over all perpendicular joints of .

Lemma C.6.

is equivalent to to order .

Proof.

Consider a perpendicular joint of . If is contained in a joint of , is the unique such joint, and we constructed above. If is not contained in a joint of , we define to be the empty set. There are three types of walls in containing :

(1)

.

(2)

, but . This type of wall does not contribute to , as the associated automorphism is central in , and in addition this wall contributes twice to , with the two contributions inverse to each other.

(3)

and . Since each added wall is of the form for some joint of , where is the direction of the wall, the direction of the wall is parallel to , contradicting being a perpendicular joint. Thus this does not occur.

From this, it is clear that , which is the identity in by Equation C.5. This holds for every perpendicular joint of .

The result then follows from Lemma C.7.

Lemma C.7.

Let and be two scattering diagrams for such that

(1)

and are equivalent to order .

(2)

is consistent to order .

(3)

is the identity for every perpendicular joint of to order .

(4)

and have the same set of incoming walls.

Then and are equivalent to order , and in particular is consistent to order .

Proof.

We work with scattering diagrams in the group . There is a finite scattering diagram with the following properties:

(1) is equivalent to ;

(2) consists only of walls trivial to order but nontrivial to order . Indeed, can be chosen so that for any general point in any , . Note that is finite because the same is true of and .

Thus to show and are equivalent, it is sufficient to show that is equivalent to the empty scattering diagram. To do so, replace with an equivalent scattering diagram with minimal support. Let be a perpendicular joint of . Then in , , since and automorphisms in are central in . However, this implies that for each with and two points in on either side of , the automorphisms associated with crossing in through either or must be the same in order for these two automorphisms to cancel in . From this it is easy to see that is equivalent to a scattering diagram such that for every wall , each facet of is a parallel joint of , i.e., the direction is tangent to every facet of . However, such a wall must be incoming, contradicting, if is nonempty, the fact that and have the same set of incoming walls by assumption.

C.2. The proof of Theorem 1.28

We fix the notation of Theorem 1.28, and in addition we make use of the notation of Definition 1.22 and as in Equation 1.26, the map associated to crossing the slab from to .

We define the Lie algebra

and set , as usual, with the degree function given by . We note that acts on as usual, and if is a scattering diagram in the sense of Definition 1.27, then all automorphisms associated to crossing walls (rather than slabs) lie in .

Besides the Lie algebra just defined, recall we also have as usual. We have the degree map given by , but we also have given by the restriction of . We use the notation and to distinguish between the two possibilities for determined by the two choices of degree map. Then and we define . Note that both act faithfully on , where the completion is with respect to the maximal monomial ideal , and act faithfully on . There are inclusions and . Only the second inclusion holds at finite order, i.e., .

For each of the above Lie algebras , we can now also talk about scattering diagrams for using Definitions 1.4 and 1.6, replacing with in those definitions.

For a joint , we define , as subgroups of , defined analogously to Equation C.3.

Finally, we will need one other group. We define, for a fixed ,

There is an inclusion and a surjection .

We need to understand the interaction between elements of and the automorphism associated to crossing the slab (see Reference GS11, Lemma 2.15). Recall the notation from Construction C.1; this is applied also to the various assorted groups above.

Lemma C.8.

Let (resp. ), and let (resp. ) be an automorphism of the form for . Let . If , then

while if , then

Here, we view or as automorphisms of and or as subgroups of the group of automorphisms of this ring.

Proof.

Let us prove the first statement, the second being similar. It is enough to check that

But, with ,

Noting that and in addition , we see that as a derivation, writing

Of course by definition of , so the derivation lives in (resp. ).

We now proceed with the proof of Theorem 1.28.

Step I. Strategy of the proof. We will first construct using essentially the same algorithm as the one given in Construction C.1, but working with the group . The algorithm is slightly more complex because of the slab, and it needs to be carried out in two steps. To show that the diagram constructed is consistent at each step, we compare it with the scattering diagram for the group which we know exists, using as an intermediary group. Because , we obtain a consistent scattering diagram for and . While is equivalent to as a scattering diagram for by construction, this does not show uniqueness of , as there may be a different choice with wall-crossing automorphisms in but not in , so it cannot be compared with . Thus, the final step involves showing uniqueness directly for the group , again as part of the inductive proof.

We will proceed by induction on , constructing for each a finite scattering diagram for containing such that the following induction hypotheses hold:

(1)

For every joint of , there is a simple loop around small enough so that it only intersects walls and slabs containing and such that , as an automorphism of , lies in and is trivial in , or equivalently, by the inclusion , trivial in .

(2)

If is a scattering diagram for which has the same incoming walls as and satisfies (1) (with replaced by everywhere), then is equivalent to in .

Recall that joints of are either parallel or perpendicular; see Definition-Lemma C.2.

Step II. The base case. For , does the job. Indeed, all walls are trivial in , leaving just the single initial slab, and thus there are no joints.

Step III. From to : adding walls associated to joints not contained in . Now assume we have found satisfying the induction hypotheses. We need to add a finite number of walls to get . We will carry out the construction of in two steps, following Construction C.1.

First, let be a perpendicular joint of with . Let be the set of integral tangent vectors to . If is a simple loop around small enough so that it only intersects walls containing , we note that every wall-crossing automorphism contributing to lies in . Thus as in Equation C.4, in we can write

with , and with as is the identity in by the induction hypothesis. Finally, because the joint is perpendicular. Let

Here makes sense as a power series. The sign is chosen in each wall so that its contribution to is to -order .

We now take

where the union is over all perpendicular joints not contained in . We have only added a finite number of walls.

Step IV. From to : adding walls associated to joints contained in . If we did not have a slab, constructed above would now do the job as in the proof of Lemma C.6. However, the elements of trivial in do not commute with to order as automorphisms of in any reasonable sense. As a consequence, we will need to add some additional walls coming from joints in , some of which have arisen as the intersection of with walls added in Step III.

Consider a perpendicular joint of . Necessarily, the linear span of is for some . Furthermore, we can choose so that any wall containing then has linear span for some nonnegative rational numbers. The direction of such a wall is positively proportional to . We now distinguish between two cases. Note that as the joint is not parallel, so we call the joint positive or negative depending on the sign of . Note that if the joint is positive (negative), then is positive (negative) for all .

If the joint is positive, then choose so that the first wall crossed is , passing from to . We can write

where are compositions of wall-crossing automorphisms. It then follows from for all , and Lemma C.8 that , hence . If the joint is negative, then we use a slightly different loop: without changing the orientation of the loop , change the endpoints so that now starts and ends in , crossing just before its endpoint. Then

and again by Lemma C.8, .

Thus in both cases, and is the identity in . Thus we still have Equation C.10, and we can produce a scattering diagram in the same way as for the joints not contained in . We then set

where the union is over perpendicular joints of contained in .

Step V. Part of the induction hypothesis is satisfied. Consider a perpendicular joint of . First suppose . We proceed as in the proof of Lemma C.6. If is contained in a joint of , there is a unique such joint, say , and we constructed above. If is not contained in a joint of , we define to be the empty set. There are three types of walls in containing :

(1)

.

(2)

, but . This type of wall does not contribute to in . Indeed, the associated automorphism is in the center of and this wall contributes twice to , with the two contributions inverse to each other, so the contribution cancels.

(3)

and . Since each added wall is of the form for some joint of , where is the direction of the wall, the direction of the wall is parallel to , contradicting being a perpendicular joint. Thus this does not occur.

From this we see by construction of that is the identity in .

On the other hand, suppose is a perpendicular joint of contained in . Then since no wall of is contained in , by definition of , in fact is a joint of . Thus we see again by construction of that to order , is the identity for the loop around described in Step IV. Recall the choice of loop depends on whether the joint is positive or negative.

Now we show that satisfies the induction hypothesis (1), using the above existence of a such that for each perpendicular joint . Note that there is a map for any . The slab automorphism can be viewed as an element of for any , and hence can be viewed as a scattering diagram for in the sense of Definition 1.6. We will first show that is consistent as a diagram for inductively on .

The base case is . All walls of are trivial to -order and hence to -order . Now satisfies the main induction hypothesis (1) at order , which implies via the natural map that is consistent as a diagram for . Indeed, as is a finite scattering diagram, it is enough to check that is the identity in for any small loop around any joint . By the hypothesis (1), this is the case for some loop , and hence for all loops. Note that by uniqueness of consistent scattering diagrams with the same incoming walls, we also record for future use:

The induction step follows from Lemma C.7, applied to , , and the group being (a quotient of , so the argument of Lemma C.7 still applies). Indeed, if we assume is consistent in , then it is equivalent to as a scattering diagram in . Furthermore, is consistent to all orders by Theorem 1.12, and has the same set of incoming walls as by construction. Finally, is the identity in for any perpendicular joint , as shown above. Thus and are equivalent in , and in particular is consistent in .

Thus taking the inverse limit, we see that is consistent as a scattering diagram for . This almost completes the proof of the induction hypothesis (1) in degree . Indeed, as is a subgroup of , certainly is the identity for any joint not contained in , including the parallel joints. For a perpendicular joint contained in , if we choose as given in Step IV, lies in and is the identity in by the construction of in Step IV. Finally, for a parallel joint contained in , note that all wall and slab-crossing automorphisms associated to walls containing commute, and in particular the contribution of and in as an automorphism of cancel, so that the latter automorphism lies in . Hence the image of this automorphism in must also be trivial. This gives the induction hypothesis (1).

Step VI. Uniqueness. Suppose we have constructed two scattering diagrams for from which satisfy the inductive hypothesis (1) to -order , but with the group replaced with . By the induction hypothesis (2), these two scattering diagrams are equivalent to -order , and we wish to show they are equivalent to -order . One first constructs a finite scattering diagram consisting only of outgoing walls whose attached functions are of the form with and , with the property that is equivalent to to -order . This is done precisely as in the proof of Lemma C.7. We need to show is equivalent to the empty scattering diagram to -order .

To show this, first note that for any loop which does not cross the slab , to -order implies that to -order . Indeed, all wall-crossing automorphisms of are central in . Now if with , let be the set of walls in with attached functions of the form . Note all wall-crossing automorphisms of , viewed as elements of , lie in , which as a group coincides with the additive group structure on . Thus for any path not crossing , we obtain a unique decomposition from the -grading on , and if is the identity, so is each .

Fixing as above, replace with an equivalent scattering diagram with smallest possible support, and let . So if is a general point, if and only if is not the identity. Assume first that . We shall show . Assume not. Taking a general point , it is not possible for the ray to be contained in . This is because consists of only a finite number of walls, none of which are incoming. Let , and . This makes sense as is in the set over which we are taking the maximum, as we are assuming . Then necessarily is in a joint of , and every wall of containing is contained in . Furthermore, since , , it follows that and is not contained in . Thus given a loop around , is the identity. This implies that in fact to -order ,

In particular, a point for small is contained in precisely those walls of containing . But then , contradicting minimality of . Thus one finds that . Similarly, if , then . In particular, if , , but there are no walls contained in , so in this case .

Now consider a joint of contained in . There are three cases: either is perpendicular and positive, perpendicular and negative, or parallel. Consider the first case. Take a loop around as in the positive case in Step IV. Because of positivity, if a wall of contains , then with chosen so that , we must have and hence is contained in . Thus we have that to -order ,

as in Equation C.11, where and denote the contributions coming from the scattering diagrams and and the pieces of not crossing . The same argument works for negative joints, while a parallel joint cannot contain any wall of (as we showed above that if ) so that trivially. We can now repeat the argument of the previous paragraph, taking for any a general point rather than . This allows us to conclude that for all , proving uniqueness.

Step VII. Finishing the proof of Theorem 1.28. Having completed the induction step, we take . We need to check it satisfies the stated conditions in Theorem 1.28. Certainly conditions (1) and (2) hold by construction.

For (3), first recall that because by construction can be viewed as a scattering diagram for , it can also be viewed as a scattering diagram for via the inclusion , and in addition , so that is viewed as a scattering diagram for in the sense of Definition 1.6, i.e., with no slab. Now as a scattering diagram for , is equivalent to by Equation C.12. By consistency of , is independent of the endpoints of as an element of . Now suppose are two automorphisms of which induce automorphisms of (i.e., for , , giving a map which is an automorphism), and agree as automorphisms of the latter ring. Then agree as automorphisms of . Thus in particular, is independent of the endpoints of as an automorphism of . This gives condition (3).

The uniqueness of with these properties then follows from the induction hypothesis (2). Indeed, if satisfies conditions (1)–(3) of Theorem 1.28, then working by induction on the order , the induction hypothesis (1) holds for (the existence of with only being an issue for joints contained in , and Step IV explains how to choose the loop ). Thus by induction hypothesis (2), and are equivalent to order .

This completes the proof of Theorem 1.28.

C.3. The proof of Theorem 1.13

The key point of the proof is just the positivity of the simplest scattering diagram as described in Example 1.14, which we use to analyze general two-dimensional scattering diagrams. We will consider a somewhat more general setup, but only in two dimensions, than that considered in the rest of this paper. In particular, we will follow the notation of Reference G11, §6.3.1, taking , , and assume given a monoid with a map , . We will consider scattering diagrams for this data as in Reference G11, Def. 6.37, consisting of rays and lines which do not necessarily pass through the origin. Given any scattering diagram , the argument of Kontsevich and Soibelman from Reference KS06 (see Reference G11, Theorem 6.38 for an exposition of this particular case) adds rays to to obtain a scattering diagram such that is the identity for every loop . This diagram is unique up to equivalence.

The fundamental observation involves a kind of universal scattering diagram:

Proposition C.13.

In the above setup, suppose given , , with , and positive integers . Consider the scattering diagram

. We can choose within its equivalence class so that for any given ray , we have

for a positive integer and the nonnegative integers with at least two of them nonzero.

Proof.

Step I. The change of monoid trick. Note that if the generate a rank 1 sublattice of , then all the wall-crossing automorphisms of commute and , so we are done. So assume from now on that the generate a rank 2 sublattice of .

Let , generated by , define a map by , and a map by . We extend to a map , and define, for a scattering diagram for the monoid , . Clearly, if , then . Thus if , then is equivalent to , by uniqueness of up to equivalence. So it is sufficient to show the result with , .

Step II. Everything but the positivity of the exponents. We can construct specifically using the original method of Reference KS06, already explained here in Steps III and IV of the proof of Theorem 1.28: we construct order by order, constructing so that is the identity modulo for a loop around the origin. Given a description

with the primitive and the all distinct, we add a collection of rays

for some . However, inductively, we can show the can be taken to be integers. Indeed, if all rays in have this property, then is in fact an automorphism of , and thus the appearing in Equation C.14 of are also integers.

Next let us show that any exponent is of the form with at least two of the nonzero. The pronilpotent group in which all automorphisms live is given by the Lie algebra

following the notation of Reference G11, pp. 290–291. This contains a subalgebra where the sum is taken over all not proportional to one of the . Then clearly , so the corresponding pronilpotent group is normal in . Furthermore, is abelian, hence so is . For any loop , the image of is thus the identity in , as every wall in contributes twice to , but with inverse automorphisms. Assume inductively that only contains rays whose attached functions have not proportional to any . Then the wall-crossing automorphisms associated to these rays lie in , so is the identity in , i.e., it lies in . Thus the expression of Equation C.14 lies in , hence the inductive step follows.

It remains to show that each wall added is of the form with positive.

Step III. The perturbation trick. We will now show the result for all monoids for all , all choices of , all choices of with , and all positive choices of . (Note by Step I this is a bit more than we need, as we do not take the to necessarily be generators of .) All cases are dealt with simultaneously by induction.

We define for the order , which is the unique such that . For a ray , we write , and say is a ray of order . We will go by induction on the order, showing that a ray in of order for any choice of data has positive. This is obviously the case for , as all elements of have order at least . So assume the induction hypothesis is true for all orders , and we need to show rays added of order have positive exponent.

We will use the perturbation trick repeatedly. Given a scattering diagram for which we would like to compute , choose general for each . Define ; this is the perturbed diagram. We can then run the Kontsevich–Soibelman algorithm for , for example as described in Reference G11, Theorem 6.38. This gives a scattering diagram with the property that is the identity for every loop . This is the case in particular for a very large loop around the origin which contains all singular points of . We can assume as usual that has been constructed only by adding rays of the form .

Then, up to equivalence, can be obtained from by taking the asymptotic scattering diagram of ; i.e., just translate each line of so it passes through the origin and each ray of so its endpoint is the origin. See §1.4 of Reference GPS for more detail. If after performing this translation, we obtain a number of rays with the same support of the form , in some index set, we can replace all these rays with a single ray without affecting the equivalence class. Thus if we want to show positivity of the exponents for , it is enough to show the desired positivity for .

We will typically use an induction hypothesis to show positivity for . Indeed, for each order, we will run the Kontsevich–Soibelman algorithm at each singular point, and the behavior at each singular point is equivalent to a scattering diagram of the general type being considered. Indeed, if is a singular point of some constructed to order , we obtain a local version of the scattering diagram at by replacing each with with , and replacing such translated rays with the line spanned by the ray if the translated ray does not have the origin as its endpoint. As long as all attached functions of rays and lines passing through are of the form with a positive integer, we are back in the original situation of the proposition. We shall write for the set of lines in .

We first observe that using the perturbation trick it is enough to show the induction hypothesis for order when at most two of the have . Indeed, after perturbing, the lines of only intersect pairwise, but as more rays are added as the Kontsevich–Soibelman algorithm is run, one might have more complicated behavior at singular points. However, any ray added has order . Thus we only have to analyze initial scattering diagrams with at most two lines of order .

Next we observe the induction hypothesis allows us to show the result only for , with both lines having order . Indeed, write as and order the so that . Apply Step I, getting a map with . We are trying to prove that rays with -order have positive exponent. But consider a ray which is the image under of a ray appearing in with and at least one of nonzero. Then , so by the induction hypothesis we can assume is positive. On the other hand, rays of the form with for appearing in already appear in , as follows easily by working modulo the ideal in generated by the . Thus we are only concerned about rays which arise from scattering the two order lines. Thus it is sufficient to show the result when .

Step IV. The change of lattice trick. To deal with the case where consists of two lines, we use the change of lattice trick to reduce to a simpler expression for the scattering diagram. By Step I, we can take , . Let be the sublattice generated by . Note as in Step I we can assume that this is a rank sublattice, as otherwise the automorphisms associated to the two lines commute. Then is a superlattice of , with dual basis . In what follows, we will talk about scattering diagrams defined using both the lattice and . Bear in mind that a wall could be interpreted using either lattice, and the automorphism induced by crossing such a wall depends on which lattice we are using, as primitive vectors in differ from primitive vectors in .

To see the relationship between these automorphisms, for , let

Then a wall for induces a wall-crossing automorphism of which is the same as the automorphism induced by the wall for , where is primitive and annihilates .

Consider

as a scattering diagram for the lattice . Let . Let be the scattering diagram for obtained by replacing every wall with . Thus the wall-crossing automorphism for each wall in as a scattering diagram for the lattice is the same automorphism for the corresponding wall in . Then is the identity. Thus by uniqueness of the scattering process up to equivalence, is equivalent to . (Note this implies that also, as only involves integer exponents.)

Thus it is enough to prove the desired positivity for the scattering diagram . To do so, we use a variant of the perturbation trick, factoring the two lines in . We choose general , with , . Define

Again, we initially only have pairwise intersections. The first stage of this algorithm will then only involve points where two lines of the form and intersect. The algorithm only adds one ray in the direction with endpoint the intersection point and attached function , as follows from Example 1.14. This now accounts for all new rays of order . We continue to higher degree, but now we can use the induction hypothesis at every singular point as we did in Step III, because every line in has order except for possibly one or two of the given lines of order , and we have already accounted for all rays produced by collisions of two lines of order .

Corollary C.15.

In the situation of Proposition C.13, suppose instead that

where now , the ground field. Choosing up to equivalence, we can assume that each ray satisfies

for some choice of nonnegative integers and where is a positive integer.

Proof.

This follows easily from from Proposition C.13. First, using the change of monoid trick (Step I of the proof of Proposition C.13), we may assume and . Consider the automorphism defined by . Applying to the function attached to each wall of gives a scattering diagram whose incoming walls are precisely those of , and for a loop around the origin. Thus we can take and the result follows from Proposition C.13.

Proof of Theorem 1.13.

In fact one can use the as constructed explicitly in the algorithm of the proof of Theorem 1.28. The only issue is that we need to know that the walls added at each joint have the desired positivity property. Note that the statement of Theorem 1.13 involves scattering diagrams without slabs, while the proof of Theorem 1.28 given involves a slab. So for the purpose of this discussion, we can ignore all issues concerning the slab in the proof of Theorem 1.28, and the only thing we need to do is look at the procedure for producing in Step II of the proof of Theorem 1.28.

For a perpendicular joint of , we can split , where is a rank 2 lattice. For each wall containing , we can inductively assume that for some positive integer , and split , with and . Because is perpendicular, we have . We will apply Corollary C.15 to the case where the monoid is the one being used in Theorem 1.21, and is the projection. We can then view the computation at the joint as a two-dimensional scattering situation in the lattice over the ground field , the quotient field of . To obtain the relevant two-dimensional scattering diagram, we replace each wall with with in . We are then in the situation of Corollary C.15, and the result follows.

Acknowledgments

We received considerable inspiration from conversations with A. Berenstein, V. Fock, S. Fomin, A. Goncharov, B. Keller, B. Leclerc, J. Kollár, G. Muller, G. Musiker, A. Neitzke, D. Rupel, M. Shapiro, B. Siebert, Y. Soibelman, and D. Speyer.

Table of Contents

  1. Abstract
  2. Introduction
    1. 0.1. Statement of the main results
    2. Definition 0.1.
    3. Theorem 0.3.
    4. Corollary 0.4 (Positivity of the Laurent phenomenon).
    5. Example 0.5.
    6. Definition 0.6.
    7. Proposition 0.7 (Proposition 8.25).
    8. 0.2. Toward the main theorem
    9. Theorem 0.8 (Lemma 2.10 and Theorem 2.13).
    10. Theorem 0.9 (Corollary 5.3(1)).
    11. 0.3. Convexity conditions
    12. Conjecture 0.10.
    13. Definition 0.11.
    14. Theorem 0.12.
    15. Proposition 0.14.
    16. Example 0.15.
    17. Theorem 0.17.
    18. 0.4. Representation-theoretic applications
    19. Theorem 0.19 (Corollaries 9.17 and 9.18).
    20. Corollary 0.20.
    21. Corollary 0.21.
  3. 1. Scattering diagrams and chamber structures
    1. 1.1. Definition and constructions
    2. The injectivity assumption
    3. Definition 1.2.
    4. Lemma 1.3.
    5. Definition 1.4.
    6. Definition 1.6.
    7. Definition 1.8.
    8. Lemma 1.9.
    9. Definition 1.10.
    10. Definition 1.11.
    11. Theorem 1.12.
    12. Theorem 1.13.
    13. Example 1.14.
    14. Example 1.15.
    15. 1.2. Construction of consistent scattering diagrams
    16. Theorem 1.17 (Kontsevich and Soibelman).
    17. Proposition 1.20.
    18. Theorem 1.21.
    19. 1.3. Mutation invariance of the scattering diagram
    20. Definition 1.22.
    21. Theorem 1.24.
    22. Definition 1.25.
    23. Definition 1.27.
    24. Theorem 1.28.
    25. Construction 1.30 (The chamber structure).
    26. Definition 1.32.
  4. 2. Basics on tropicalization and the Fock–Goncharov cluster complex
    1. Proposition 2.4.
    2. Lemma 2.8.
    3. Definition 2.9.
    4. Lemma 2.10.
    5. Construction 2.11.
    6. Theorem 2.13.
    7. Example 2.14.
  5. 3. Broken lines
    1. Definition 3.1.
    2. Definition 3.3.
    3. Proposition 3.4.
    4. Theorem 3.5.
    5. Proposition 3.6.
    6. Proposition 3.8.
    7. Corollary 3.9.
    8. Example 3.10.
  6. 4. Building from the scattering diagram and positivity of the Laurent phenomenon
    1. Construction 4.1.
    2. Proposition 4.3.
    3. Theorem 4.4.
    4. Definition 4.8.
    5. Theorem 4.9.
    6. Theorem 4.10 (Positivity of the Laurent phenomenon).
  7. 5. Sign coherence of - and -vectors
    1. Construction 5.1.
    2. Lemma 5.2.
    3. Corollary 5.3.
    4. Corollary 5.5 (Sign coherence of -vectors).
    5. Definition 5.6.
    6. Proposition 5.7.
    7. Definition 5.8.
    8. Corollary 5.9.
    9. Definition 5.10.
    10. Theorem 5.11 (Sign coherence of -vectors).
    11. Lemma 5.12.
  8. 6. The formal Fock–Goncharov conjecture
    1. Definition-Lemma 6.2.
    2. Definition 6.3.
    3. Proposition 6.4.
    4. Proposition 6.5.
    5. Definition 6.6.
    6. Theorem 6.8.
    7. Claim 6.9.
    8. Claim 6.10.
    9. Corollary 6.11.
  9. 7. The middle cluster algebra
    1. 7.1. The middle algebra for
    2. Proposition 7.1.
    3. Definition 7.2.
    4. Definition 7.3.
    5. Lemma 7.4.
    6. Theorem 7.5.
    7. Corollary 7.6.
    8. Proposition 7.7.
    9. 7.2. From to and .
    10. Lemma 7.8 (Global monomials).
    11. Definition 7.9.
    12. Lemma 7.10.
    13. Construction 7.11 (Broken lines for and ).
    14. Definition 7.12.
    15. Corollary 7.13.
    16. Definition-Lemma 7.14.
    17. Definition 7.15.
    18. Theorem 7.16.
    19. Question 7.17.
    20. Example 7.18.
    21. Proposition 7.19.
    22. Theorem 7.20.
  10. 8. Convexity in the tropical space
    1. 8.1. Convexity conditions
    2. Definition-Lemma 8.1.
    3. Definition 8.2.
    4. Definition 8.3.
    5. Lemma 8.4.
    6. Definition 8.6.
    7. Definition 8.8.
    8. Lemma 8.9.
    9. Proposition 8.10.
    10. 8.2. Convexity criteria
    11. Conjecture 8.11.
    12. Proposition 8.13.
    13. Definition 8.14.
    14. Lemma 8.15.
    15. Proposition 8.16.
    16. 8.3. The canonical algebra
    17. Proposition 8.17.
    18. Corollary 8.18.
    19. Theorem 8.19.
    20. Claim 8.20.
    21. Corollary 8.21.
    22. Proposition 8.22.
    23. 8.4. Conditions implying has EGM and the full Fock–Goncharov conjecture
    24. Definition 8.23.
    25. Proposition 8.24.
    26. Proposition 8.25.
    27. Proposition 8.26.
    28. Proposition 8.27.
    29. 8.5. Compactifications from positive polytopes
    30. Lemma 8.29.
    31. Theorem 8.30.
    32. Example 8.31.
    33. Theorem 8.32.
    34. Lemma 8.33 (Kollár).
    35. Theorem 8.35.
  11. 9. Partial compactifications and representation-theoretic results
    1. 9.1. Partial minimal models
    2. Definition 9.1.
    3. Lemma 9.2.
    4. Lemma 9.3.
    5. Proposition 9.4.
    6. Lemma 9.6.
    7. Proposition 9.7.
    8. Conjecture 9.8.
    9. Definition 9.9.
    10. Lemma 9.10.
    11. 9.2. Cones cut out by the tropicalized potential
    12. Lemma 9.12.
    13. Definition 9.13.
    14. Lemma 9.14.
    15. Corollary 9.15.
    16. Proposition 9.16.
    17. Corollary 9.17.
    18. Corollary 9.18.
  12. 10. Links with quiver representations and work of Reineke
    1. Proposition 10.1.
    2. Example 10.3.
    3. Example 10.4.
  13. Appendix A. Review of notation and Langlands duality
    1. Definition A.4 (Fock–Goncharov dual).
    2. Proposition A.5.
  14. Appendix B. The and -varieties with principal coefficients
    1. Proposition B.2.
    2. Lemma B.6.
    3. Lemma B.7.
    4. Construction B.9 (Partial compactifications from frozen variables).
    5. Proposition B.11.
  15. Appendix C. Construction of scattering diagrams
    1. C.1. An algorithmic construction of scattering diagrams
    2. Construction C.1.
    3. Definition-Lemma C.2.
    4. Lemma C.6.
    5. Lemma C.7.
    6. C.2. The proof of Theorem 1.28
    7. Lemma C.8.
    8. C.3. The proof of Theorem 1.13
    9. Proposition C.13.
    10. Corollary C.15.
  16. Acknowledgments

Figures

Figure 1.1.

Scattering diagram for Example 1.14

\input{scat11.pstex_t}
Figure 1.2.

Scattering diagram for Example 1.15, . The unlabeled rays intersecting the interior of the fourth quadrant have attached functions , , , and in clockwise order.

\input{scat13.pstex_t}
Figure 1.3.

The general appearance of the scattering diagram of Example 1.15 for

\input{diagram33.pstex_t}
Figure 3.1.

Broken lines defining

\input{brokenline11.pstex_t}
Figure 3.2.

Broken lines defining

\input{brokenline22.pstex_t}

Mathematical Fragments

Definition 0.1.

A global monomial on a cluster variety is a regular function on which restricts to a character on some torus in the atlas. For an -type cluster variety, a global monomial is the same as a cluster monomial. One defines the upper cluster algebra associated to by and the ordinary cluster algebra to be the subalgebra of generated by global monomials.

Equation (0.2)
Theorem 0.3.

Let be one of . The following hold:

(1)

There are canonically defined nonnegative structure constants

These are given by counts of broken lines, certain combinatorial objects which we will define. The value is not taken in the or case.

(2)

There is a canonically defined subset with such that the restriction of gives the vector subspace with basis indexed by the structure of an associative commutative -algebra.

(3)

, i.e., contains the -vector of each global monomial.

(4)

For the lattice structure on determined by any choice of seed, is closed under addition. Furthermore, is saturated: for and , if and only if .

(5)

There is a canonical -algebra map which sends for to the corresponding global monomial.

(6)

The image is a universal positive Laurent polynomial (i.e., a Laurent polynomial with nonnegative integral coefficients in the cluster variables for each seed

(7)

is injective for or . Furthermore, is injective for under the additional assumption that there is a seed for which all the covectors , , lie in a strictly convex cone. When is injective, we have canonical inclusions

Example 0.5.

Let . Choose a Borel subgroup of , a maximal torus, and let be the unipotent radical of . These choices determine a cluster variety structure (with frozen variables) on , with , the ring of regular functions on ; see Reference GLS, §10.4.2.

Theorem 0.3 implies that these choices canonically determine a vector space basis . Each basis element is an -eigenfunction for the natural (right) action of on . For each character , is a basis of the weight space . The are the collection of irreducible representations of , each of which thus inherits a basis, canonically determined by the choice of .

We give, combining our results with results of T. Magee, much more precise results; see Corollary 0.20.

Canonical bases for have been constructed by Lusztig. Here we will obtain bases by a procedure very different from Lusztig’s, as a special case of the more general Reference GHK11, Conjecture 0.6, which applies in theory to any variety with the right sort of volume form. See Remark 0.16 for further commentary on this.

Definition 0.6.

We say the full Fock–Goncharov conjecture holds for a cluster variety if the map of Theorem 0.3 is injective,

Note this implies .

Proposition 0.14.

Consider the following conditions on a cluster algebra :

(1)

The exchange matrix has full rank, is generated by finitely many cluster variables, and is a smooth affine variety.

(2)

has an acyclic seed.

(3)

has a seed with a maximal green sequence.

(4)

For some seed, the cluster complex is not contained in a half-space.

(5)

has EGM.

Then implies Proposition 8.27). Furthermore, implies implies implies Propositions 8.24 and 8.25 Finally, implies the full Fock–Goncharov conjecture, for , , or very general , or, under the convexity assumption of Theorem 0.3, for (Proposition 8.25

Remark 0.16.

In general, we conjecture the bases we construct for rings of global functions on cluster varieties or partial compactifications are intrinsic to the underlying log Calabi–Yau variety and do not depend on the particular cluster structure on . This is a nontrivial statement: there exist varieties with multiple cluster structures (in particular different atlases of tori for the same variety). Yan Zhou will show in her PhD thesis that the (principal coefficient version of) the cluster variety associated to the once-punctured torus is an example.

This conjecture is suggested by Reference GHK11, Conjecture 0.6, and the results of Reference GHK11, Reference GHK12, and Reference GHKII prove this in the case of the cluster varieties where the skew-symmetric form has rank 2, which includes the case of the sphere with four punctures. Thus we have the (at least to us) remarkable conclusion that in many cases where bases occur because of some extrinsic interpretation of the spaces, in fact this extrinsic interpretation is irrelevant. For example, the theta functions given by trace functions above, which would appear to depend on the realization of the cluster variety as a moduli space of local systems, are actually intrinsic to the underlying variety. In the case of Example 0.5, where bases may arise from representation theory, our basis does not use the group-theoretic aspects of the spaces. The suggestion that the canonical basis is independent of the cluster structure may surprise some, as understanding the canonical basis was the initial motivation for the Fomin–Zelevinsky definition of cluster algebras.

Equation (0.18)
Theorem 0.19 (Corollaries 9.17 and 9.18).

Assume that each frozen index has an optimized seed. Then:

(1)

and are convex in our sense.

(2)

The set parameterizes a canonical basis of an algebra , and

(3)

Now assume further that we have EGM on . If for some seed , is contained in the convex hull of (which itself contains the convex hull of then , is finitely generated, and the integer points parameterize a canonical basis.

Corollary 0.20.

Let and let be the Fomin–Zelevinsky cluster variety for the basic affine space .

(1)

All the hypotheses, and thus the conclusions, of Theorem 0.19 hold. In particular parameterizes a canonical theta function basis of .

(2)

Our potential agrees with the (representation theoretically defined) potential function of Berenstein and Kazhdan Reference BK07.

(3)

The maximal torus acts canonically on , preserving the open set .

(4)

Each theta function is an -eigenfunction, and there is a canonical map

(the target is the character lattice of linear for the linear structure given by any seed, which sends an integer point to the -weight of the corresponding theta function. The slice

parameterizes a canonical theta function basis of the eigenspace , the corresponding irreducible representation of .

(5)

For a natural choice of seed, the cone is canonically identified with the Gelfand–Tsetlin cone.

Corollary 0.21.

Let be the Fock–Goncharov cluster variety for

(1)

All the hypotheses, and thus the conclusions, of Theorem 0.19 hold. In particular the cone parameterizes a canonical theta function basis of .

(2)

Our potential function agrees with the (representation theoretically defined) potential function of Goncharov and Shen Reference GS13.

(3)

acts canonically on , preserving the open subset .

(4)

Each theta function is an -eigenfunction, and there is a canonical map

linear for the linear structure given by any seed, which sends an integer point to the -weight of the corresponding theta function. The slice

parameterizes a canonical theta function basis of the eigenspace

In particular, the number of integral points in is the corresponding Littlewood–Richardson coefficient.

(5)

For a natural choice of seed, the cone is canonically identified with the Knutson–Tao hive cone.

Equation (1.1)
Definition 1.2.

Let , and . Define to be the automorphism of given by

where is the generator of the monoid .

Lemma 1.3.

For , is the subgroup of automorphisms of the form for as in Definition 1.2 with the given . More specifically, acts as the automorphism with , where is the smallest positive rational number such that .

Definition 1.4.

A wall in (for and ) is a pair such that

(1)

for some primitive .;

(2)

is a -dimensional convex (but not necessarily strictly convex) rational polyhedral cone.

The set is called the support of the wall .

Definition 1.6.

A scattering diagram for and is a set of walls such that for every degree , there are only a finite number of with the image of in not the identity.

If is a scattering diagram, we write

for the support and singular locus of the scattering diagram. If is a finite scattering diagram, then its support is a finite polyhedral cone complex. A joint is an -dimensional cell of this complex, so that is the union of all joints of .

Lemma 1.9.

Two scattering diagrams are equivalent if and only if for all general .

Theorem 1.12.

There is a scattering diagram satisfying:

(1)

is consistent,

(2)

,

(3)

consists only of outgoing walls.

Moreover, satisfying these three properties is unique up to equivalence.

Theorem 1.13.

The scattering diagram is equivalent to a scattering diagram all of whose walls satisfy for some and a positive integer. In particular, all nonzero coefficients of are positive integers.

Example 1.14.

Take , and the skew-symmetric form given by the matrix , where . Let be the dual basis of , and write , . We get

Then one checks easily that

See Figure 1.1. (See for example Reference GPS, Example 1.6.)

Example 1.15.

Take , with basis , and take to be the sublattice generated by . Further, take , , where are two positive integers, and take the skew-symmetric form to be the same as in the previous example. Then , . Taking as before , , we get

For most choices of and , this is a very complicated scattering diagram. A very similar scattering diagram, with functions and , has been analyzed in Reference GP10, but it is easy to translate this latter diagram to the one considered here by replacing by and using the change of lattice trick, which is given in Step IV of the proof of Proposition C.13. All rays of are contained strictly in the fourth quadrant (i.e., in particular are not contained in an axis). Without giving the details, we summarize the results. There are two linear operators given by the matrices in the basis as

Then is invariant under and , in the sense that if , we have provided is contained strictly in the fourth quadrant. It is also the case that applying to or to gives an element of . Further, contains a discrete series of rays consisting of those rays in the fourth quadrant obtained by applying and alternately to the above rays supported on and . These rays necessarily have functions of the form or for various choices of and . If , we obtain a finite diagram. (Moreover, the corresponding cluster variety is the cluster variety of finite type Reference FZ03a associated to the root system , , or for , , or respectively.) See Figure 1.2 for the case . If , these rays converge to the rays contained in the two eigenspaces of and . These are rays of slope . This gives a complete description of the rays outside of the cone spanned by these two rays. The expectation is that every ray of rational slope appears in the interior of this cone, and the attached functions are in general unknown (see Figure 1.3). However, in the case, it is known Reference R12 that the function attached to the ray of slope is

The chamber structure one sees outside the quadratic irrational cone is very well-behaved and familiar in cluster algebra theory. In particular, the interiors of the first, second, and third quadrants are all connected components of , and there are for an infinite number of connected components in the fourth quadrant. We will see in §2 that this chamber structure is precisely the Fock–Goncharov cluster complex.

On the other hand, it is precisely the rich structure inside the quadratic irrational cone which scattering diagram technology brings into the cluster algebra picture.

Equation (1.16)
Theorem 1.17 (Kontsevich and Soibelman).

The assignment of to gives a one-to-one correspondence between equivalence classes of consistent scattering diagrams and elements .

Equation (1.18)
Equation (1.19)
Proposition 1.20.

is a set bijection.

Theorem 1.21.

Let be a consistent scattering diagram corresponding to . The following hold:

(1)

For each , to any fixed finite order, there is an open neighborhood of such that for all general . Here denotes the component of indexed by .

(2)

is equivalent to a diagram with only one wall in containing for each , and the group element attached to this wall is .

(3)

Set

Then is equivalent to a consistent scattering diagram such that and consists only of outgoing walls. Furthermore, up to equivalence, is the unique consistent scattering diagram with this property.

(4)

The equivalence class of a consistent scattering diagram is determined by its set of incoming walls.

Definition 1.22.

We set

For , define the piecewise linear transformation by, for ,

As we will explain in §2, is the tropicalization of viewed as a birational map between tori. We will write and to be the linear transformations used to define in the regions and , respectively.

Define the scattering diagram to be the scattering diagram obtained by the following:

(1)

For each wall , where , we have one or two walls in given as

throwing out the first or second of these if or , respectively. Here for linear, we write for the formal power series obtained by applying to each exponent in .

(2)

also contains the wall .

Theorem 1.24.

Suppose the injectivity assumption is satisfied. Then is a consistent scattering diagram for and . Furthermore, and are equivalent.

Definition 1.25.

Let be a top-dimensional cone containing , , and , and such that . Set , and .

Equation (1.26)
Definition 1.27.

A wall for and ideal is a pair with as in Definition 1.4, but with , and congruent to mod . The slab for the seed means the pair . Note since this does not qualify as a wall. Now a scattering diagram is a collection of walls and possibly this single slab, with the condition that for each , for all but finitely many walls in .

Theorem 1.28.

There exists a scattering diagram in the sense of Definition 1.27 such that

(1)

,

(2)

consisting only of outgoing walls, and

(3)

as an automorphism of only depends on the endpoints of .

Furthermore, with these properties is unique up to equivalence.

Finally, is also a scattering diagram for the data and, as such, is equivalent to .

Remark 1.29.

Note in particular that the theorem implies does not contain any walls contained in besides . Indeed, no wall of is contained in : only the slab is contained in .

Construction 1.30 (The chamber structure).

Suppose given fixed data satisfying the injectivity assumption and seed data . We then obtain for every seed obtained from via mutation a scattering diagram . In each case we will choose a representative for the scattering diagram with minimal support.

Note by construction and Remark 1.29, irrespective of the representative of used, contains walls whose union of supports is . Furthermore, we have given by Equation 1.16, which can be written more explicitly as

Then are the closures of connected components of . Similarly, we see that taking to be the chambers where all are positive (or negative), we have that is the closure of a connected component of , so that is the closure of a connected component of . Note that the closures of and have a common codimension 1 face given by the intersection with . This gives rise to the following chamber structure for a subset of .

We refer the reader to Appendix A for the definition of the infinite oriented tree (or ) used for parameterizing seeds obtained via mutation of . In particular, for any vertex of , there is a simple path from the root vertex to , indicating a sequence of mutations and hence a piecewise linear transformation

Note that is defined using the basis vector of the seed , not the basis vector of the original seed . By applying Theorem 1.24 repeatedly, we see that

(where applied to the scattering diagram is interpreted as the composition of the actions of each ) and

is the closure of a connected component of .

Note that the map from vertices of to chambers of is never one-to-one. Indeed, if is the vertex obtained by following the edge labeled twice starting at the root vertex, one checks that , even though (see Reference GHK13, Remark 2.5).

Thus we have a chamber structure on a subset of ; in general, the union of the cones do not form a dense subset of .

Since we will often want to compare various aspects of this geometry for different seeds, we will write the short-hand for an object parameterized by a vertex where the root of the tree is labeled with the seed . In particular:

Definition 1.32.

We write for the chamber of corresponding to the vertex . We write for the set of chambers for running over all vertices of . We call elements of cluster chambers.

Equation (2.3)
Proposition 2.4.

defined in Equation 1.23 is the Fock–Goncharov tropicalization of

Equation (2.5)
Equation (2.6)
Equation (2.7)
Lemma 2.8.
(1)

For a positive Laurent polynomial (i.e., , and

where is the canonical isomorphism Equation 2.7.

(2)

If is a divisorial discrete valuation and is any Laurent polynomial (so now then

Lemma 2.10.

Suppose we are given fixed data satisfying the injectivity assumption, and suppose we are given an initial seed. For a seed obtained by mutation from the initial seed, the chamber (also identified with via is the Fock–Goncharov cluster chamber associated to . Hence the Fock–Goncharov cluster chambers are the maximal cones of a simplicial fan (of not necessarily strictly convex cones). In particular is identified with for any choice of seed giving an identification of with .

Construction 2.11.

See Appendix B for a review of the cluster variety with principal coefficients, . Any seed gives rise to a scattering diagram living in

the second equality by Proposition B.2(3). Indeed in this situation, the injectivity assumption is satisfied since the form on is nondegenerate (which is the reason we use instead of or ). Indeed, the vectors are linearly independent. Note by Theorem 1.21, contains the scattering diagram

Recall from Proposition B.2 that we have a canonical map which is defined on cocharacter lattices by the canonical projection ; see B.4. Thus the tropicalization

coincides with this projection, which can be viewed as the quotient of an action of translation by . By Definition 1.4, walls of are of the form for . Thus all walls are invariant under translation by , and thus are inverse images of walls under . So even though may not satisfy the injectivity assumption necessary to build a scattering diagram, we see that is the inverse image of a subset of canonically defined independently of the seed. In particular, note that the Fock–Goncharov cluster chamber in associated to the seed (where for all ) pulls back to the corresponding Fock–Goncharov cluster chamber in .

Theorem 2.13.

For any initial data the Fock–Goncharov cluster chambers in are the maximal cones of a simplicial fan.

Example 2.14.

Consider the rank 3 skew-symmetric cluster algebra given by the matrix

Then projecting the walls of to via , one obtains a collection of walls in a three-dimensional vector space. One can visualize this by intersecting the walls with the affine hyperplane . The collection of resulting rays and lines appears on the first page of Reference FG11. While Fock and Goncharov were not aware of scattering diagrams in this context, in fact there the picture represents the same slice of the cluster complex, and hence coincides with the scattering diagram.

The cluster complex in fact fills up the half-space . There is no path through chambers connecting and .

This example is particularly well known in cluster theory, and gives the cluster algebra associated with triangulations of the once-punctured torus.

Definition 3.1.

Let be a scattering diagram in the sense of Definition 1.6, and let and . A broken line for with endpoint is a piecewise linear continuous proper path with a finite number of domains of linearity. This path comes along with a monomial for each domain of linearity of . This data satisfies the following properties:

(1)

.

(2)

If is the first (and therefore unbounded) domain of linearity of , then .

(3)

For in a domain of linearity , .

(4)

Let be a point at which is not linear, passing from domain of linearity to . Let

Then is a term in the formal power series .

Definition 3.3.

Let be a scattering diagram, and let and . For a broken line for with endpoint , define

(where is for initial),

and

to be the monomial attached to the final (where is for final) domain of linearity of . Define

where the sum is over all broken lines for with endpoint .

For , we define for any endpoint

Proposition 3.4.

.

Theorem 3.5.

Let be a consistent scattering diagram, and let and be two points. Suppose further the coordinates of are linearly independent over , and the same is true for . Then for any path with endpoints and for which is defined, we have

Proposition 3.6.

defines a one-to-one correspondence between broken lines for with endpoint for and broken lines for with endpoint for . This correspondence satisfies, depending on whether or ,

where acts linearly on the exponents. In particular, we have

where the superscript indicates which scattering diagram is used to define the theta function.

Proposition 3.8.

Let be a basepoint, and let . Then .

Corollary 3.9.

Let be a cluster chamber, and let , . Then .

Construction 4.1.

Fix a seed . We use the cluster chambers to build a positive space. We attach a copy of the torus to each cluster chamber .

Given any two cluster chambers of , we can choose a path from to . We then get an automorphism which is independent of choice of path. If we choose the path to lie in the support of the cluster complex, then by Remark 1.29 (which shows in particular that the scattering functions on walls of the cluster complex are polynomials, as opposed to formal power series), the wall crossings give birational maps of the torus, and hence we can view as giving a well-defined map of fields of fractions

This induces a birational map

which is in fact positive.

We can then construct a space by gluing together all the tori , via these birational maps; see Reference GHK13, Proposition 2.4. We call this space (with its atlas of tori) .

We write if we need to make clear which seed is being used.

Equation (4.2)
Proposition 4.3.

Let be a seed. Let be the root of , any other vertex. Consider the Fock–Goncharov tropicalization of . Its restriction to each cluster chamber is a linear isomorphism onto the corresponding chamber . The linear map

induces an isomorphism

These glue to give an isomorphism of positive spaces .

Theorem 4.4.

Fix a seed . Let be the root of , and let be any other vertex. Let be the linear map . Let be the associated map of tori. These glue to give an isomorphism of positive spaces

Furthermore, the diagram

is commutative, where the right-hand vertical map is the isomorphism of Proposition 4.3, the left-hand vertical map the isomorphism given in Equation 4.2, and the horizontal maps are the isomorphisms just described.

Equation (4.6)
Equation (4.7)
Definition 4.8.

Given fixed and initial data , if a seed is given, with the dual basis and , a cluster monomial in this seed is a monomial on of the form with and the nonnegative for . By the Laurent phenomenon Reference FZ02b, such a monomial always extends to a regular function on . A cluster monomial on is a regular function which is a cluster monomial in some seed.

Theorem 4.9.

Let be fixed data satisfying the injectivity assumption, and let be an initial seed. Let and for some . Then is a positive Laurent polynomial which expresses a cluster monomial of in the initial seed . Further, all cluster monomials can be expressed in this way.

Theorem 4.10 (Positivity of the Laurent phenomenon).

Each cluster variable of an -cluster algebra is a Laurent polynomial with nonnegative integer coefficients in the cluster variables of any given seed.

Remark 4.11.

When fixed and initial data , have frozen variables, there is a partial compactification of cluster varieties ; see Construction B.9. We have an analogous partial compactification , given by an atlas of toric varieties . The choice of fans is forced by the identifications of Proposition 4.3: for the root of , ( as in Construction B.9) and then . Now Proposition 4.3 and Theorem 4.4 (and their proofs) extend to the partial compactifications without change. One checks easily that all mutations in the positive spaces , and all the linear isomorphisms between corresponding tori in the atlases for preserve the monomials , (these are the frozen cluster variables), so that all the spaces come with canonical projection to , preserved by the isomorphisms between these positive spaces. We shall see in the next section that in the special case of the partial compactification of , the relevant fans are particularly well-behaved.

Construction 5.1.

Fix a seed for fixed data . By Construction 4.1, the scattering diagram gives an atlas for the space . (Technically, we should write to indicate we are constructing something isomorphic to ; however, this will make the notation even less readable.) This was constructed by attaching a copy of the torus to each cluster chamber , and (compositions of) wall-crossing automorphisms give the birational maps between them. By Theorem 4.4 this space is canonically identified with : has an atlas of tori parameterized by vertices of , and we have canonical isomorphisms for each vertex which induce the isomorphism .

In what follows, if is a vertex of , we write for the seed obtained by mutating (see B.1) via the sequence of mutations dictated by the path from the root of to . As described in Remark B.10, the initial seed determines the partial compactification , given by the atlas of toric varieties

where is the cone generated by the subset of basis vectors of corresponding to the second copy of .

By Remark 4.11, the seed also determines a partial compactification (the superscript, thus the seed close to the overline in the notation, is responsible for the partial compactification), given by an atlas of toric varieties. Explicitly, if is a vertex of , the fan yields the partial compactification of in , and this is identified with via under the isomorphism of Theorem 4.4. Thus the fan giving the partial compactifaction of is

In fact, this fan is easily calculated:

Lemma 5.2.

The cones , and thus the toric varieties in the atlas for the partial compactification , are the same for all . Each is equal to the cone spanned by the vectors , where and denotes the dual basis.

Corollary 5.3.

Fix a seed , and let be the root of . The following hold:

(1)

The fiber of over is see Proposition B.2 for the definition of .

(2)

The mutation maps

for the atlas of toric varieties defining are isomorphisms in a neighborhood of the fiber over .

(3)

For the partial compactification with atlas corresponding to cluster chambers of , the corresponding mutation map between two charts (which by Lemma 5.2 has the same domain and range) is an isomorphism in a neighborhood of the fiber and restricts to the identity on this fiber.

Corollary 5.5 (Sign coherence of -vectors).

For any vertex of and fixed satisfying , either the entries , are all nonpositive or these entries are all nonnegative.

Proposition 5.7.

Fix a seed , giving the partial compactification and -equivariant . The central fiber is a -torsor. Let be a cluster monomial on and let be the corresponding lifted cluster monomial on . This restricts to a regular nonvanishing -eigenfunction along and so canonically determines an element of (its weight). This is the -vector associated to .

Definition 5.8.

Writing , let be a cluster monomial of the form on a chart , . Note that for all , so after identifying with , yields a point in the Fock–Goncharov cluster chamber , as defined in Lemma 2.10. We define to be this point of .

Corollary 5.9.

Let be a cluster monomial on , and fix a seed giving an identification . Then under this identification, is the -vector of the cluster monomial with respect to .

Definition 5.10.

Let be a cluster variety, suppose that is a global monomial (see Definition 0.1) on , and let be a seed such that is the character , . Define the -vector of to be the image of under the identifications of §2:

We write the -vector of as .

Lemma 5.12.

The facet of corresponding to is the intersection of with the orthogonal complement of the -vector for the corresponding element of (the corresponding mutation of the Langlands dual seed ; see Appendix A). Furthermore, each -vector for is nonnegative on .

Equation (6.1)
Definition-Lemma 6.2.

Let . Let be chosen generally. There are at most finitely many pairs of broken lines with , and (see Definition 3.3 for this notation). We can then define

The integers are nonnegative.

Proposition 6.4.

Notation as above. The following hold:

(1)

For , is a regular function on , and the as varies glue to give a canonically defined function .

(2)

For each and , we have , and thus the for canonically define

The are linearly independent; i.e., we have a canonical inclusion of -vector spaces

(3)

for chosen sufficiently close to . In particular, is independent of the choice of sufficiently near , and we define

for chosen sufficiently close to .

(4)

restrict to bases of as a -vector space and -module, respectively.

Proposition 6.5.

Notation as in Proposition 6.4. There is a unique inclusion

given by

We have for all .

Definition 6.6.

For , write on the torus chart of corresponding to a seed . We also have a formal expansion as for some . Set

If is the monoid generated by , it is easy to check from the construction of theta functions that

Theorem 6.8.

There is a unique function

with all the following properties:

(1)

is compatible with the -module structure on and the -translation action on in the sense that

for all , , .

(2)

For each choice of seed , the formal sum converges to in .

(3)

If then unless , and

and the coefficients are the coefficients for the expansion of viewed as an element of in the basis of theta functions from Proposition 6.4.

(4)

For any seed obtained via mutations from , is the composition of the inclusions

given by Equation 6.1 and Proposition 6.5. This sends a cluster monomial to the delta function for its -vector .

In the notation of Definition 6.6, for any seed . In particular the sets of Definition 6.6 are independent of the seed, depending only on .

Claim 6.9.

The following hold:

(1)

The collection , , forms a -basis of the vector space .

(2)

The collection , , forms a basis of as an -module.

(3)

The collection , , forms a -basis of .

Claim 6.10.

Modulo , the sums , are finite and coincide with in the charts indexed by and , respectively.

Proposition 7.1.

Choose . If for some generic basepoint there are only finitely many broken lines with and , then the same is true for any generic . In particular, is a positive universal Laurent polynomial.

Definition 7.2.

Let be the collection of such that for some (or equivalently, by Proposition 7.1, any) generic there are only finitely many broken lines with , .

Definition 7.3.

We call a subset intrinsically closed under addition if and implies .

Theorem 7.5.

Let

be the set of integral points in chambers of the cluster complex. Then

(1)

.

(2)

For

is a finite sum (i.e., for all but finitely many with nonnegative integer coefficients. If , then .

(3)

The set is intrinsically closed under addition. For any seed , the image of is a saturated monoid.

(4)

The structure constants of Definition-Lemma 6.2 make the -vector space with basis indexed by ,

into an associative commutative -algebra. There are canonical inclusions of -algebras

Under the first inclusion a cluster monomial is identified with for its -vector. Under the second inclusion each is identified with a universal positive Laurent polynomial.

Proposition 7.7.

Let . Then is an eigenfunction for the natural action on (see Proposition B.2, with weight given by the canonical map (the map being dual to the inclusion . In particular is an eigenfunction for the subtorus with weight where is given by .

Lemma 7.8 (Global monomials).

Notation as immediately above. For , the character on the torus is a global monomial if and only if is regular on the toric variety , which holds if and only if for the primitive generator of each ray in the fan . For -type cluster varieties a global monomial is the same as a cluster monomial, i.e., a monomial in the variables of a single cluster, where the nonfrozen variables have nonnegative exponent.

Definition 7.9.

Let be a cluster variety. Define to be the set of -vectors (see Definition 5.10) for global monomials, which are characters on the seed torus , and to be the union of all .

Lemma 7.10.
(1)

For of -type is the set of integral points of the cone in the Fock–Goncharov cluster complex corresponding to the seed .

(2)

In any case is the set of integral points of a rational convex cone , and the relative interiors of as varies are disjoint. The -vector depends only on the function (i.e., if restricts to a character on two different seed tori, the -vectors they determine are the same).

(3)

For , the global monomial on is invariant under the action and thus gives a global function on . This is a global monomial and all global monomials on occur this way, and .

Corollary 7.13.

Theorem 0.3 holds for .

Definition-Lemma 7.14.
(1)

Define

Noting that is invariant under -translation, we have . Furthermore, any choice of section of induces a bijection .

(2)

Define , where is given by . Given a choice of , the collection gives a -module basis for and thus a -vector space basis for . For the basis is independent of the choice of , while for it is independent up to scaling each basis vector (i.e., the decomposition of the vector space into one-dimensional subspaces is canonical).

Theorem 7.16.

For the following modified statements of Theorem 0.3 hold.

(1)

There is a map

depending on a choice of a section . This function is given by the formula

if this sum is finite; otherwise, we take . This sum is finite whenever .

(2)

There is a canonically defined subset given by such that the restriction of the structure constants give the vector subspace with basis indexed by the structure of an associative commutative -algebra.

(3)

, i.e., contains the -vector of each global monomial.

(4)

For the lattice structure on determined by any choice of seed, is closed under addition. Furthermore is saturated.

(5)

There is a -algebra map which sends for to a multiple of the corresponding global monomial.

(6)

There is no analogue of Theorem 0.3 because the coefficients of the will generally not be integers.

(7)

is injective for very general and for all if the vectors , , lie in a strictly convex cone. When is injective, we have canonical inclusions

Taking gives Theorem 0.3 for the case.

Example 7.18.

In the cases of Example 1.15, the convex hull of the union of the cones of in is all of . Indeed, the first three quadrants already are part of the cluster complex. It then follows from the fact that is closed under addition and is saturated that .

In the case of Example 2.14, we know that

It then follows again from the fact that is closed under addition that either or . We believe, partly based on calculations in Reference M13, §7.1, that in fact the latter holds.

Theorem 7.20.

For any cluster variety, there are no linear relations between cluster monomials and theta functions in . More precisely, if there is a linear relation

in , then for all . In particular the cluster monomials in are linearly independent.

Definition-Lemma 8.1.

By a piecewise linear function on a real vector space we mean a continuous function piecewise linear with respect to a finite fan of (not necessarily strictly) convex cones. For a piecewise linear function we say is min-convex if it satisfies one of the following three equivalent conditions:

(1)

There are finitely many linear functions such that for all .

(2)

for all and .

(3)

The differential is decreasing on straight lines. In other words, for a directed straight line with tangent vector , and general, then

where is general and the subscript denotes the point at which the differential is calculated.

Definition 8.2.
(1)

A piecewise linear function is a function which is piecewise linear after fixing a seed to get an identification . If the function is piecewise linear for one seed it is clearly piecewise linear for all seeds.

(2)

Let be piecewise linear, and fix a seed , to view . We say is min-convex for (or just min-convex if is clear from context) if for any broken line for in , is increasing on exponents of the decoration monomials (and thus decreasing on their negatives, which are the velocity vectors of the underlying directed path). We note that this notion is independent of mutation, by the invariance of broken lines, Proposition 3.6, and thus an intrinsic property of a piecewise linear function on .

Definition 8.3.

We say that a piecewise linear is decreasing if for , with , . Here are the structure constants of Theorem 0.3.

Lemma 8.4.
(1)

If is min-convex, then is decreasing.

(2)

If is decreasing, then for any seed , we have min-convex in the sense of Definition-Lemma 8.1.

Definition 8.8.

A convex polytope is a subset of for which there exist a finite collection of affine linear functions with

Lemma 8.9.

If a piecewise linear function is decreasing, then is positive. Furthemore, is compact if and only if is strictly negative away from .

Proposition 8.10.

Suppose is a positive polytope defined over (i.e., all the functions of Definition 8.8 are rationally defined Then is positive.

Conjecture 8.11.

If is a regular function on a log Calabi–Yau manifold with maximal boundary, then is min-convex. Here for the valuation .

Remark 8.12.

To make sense of the conjecture, one needs a good theory of broken lines, currently constructed in Reference GHK11 in dimension two, and here for cluster varieties of all dimensions. In dimension two, the conjecture has been proven by Travis Mandel Reference M14. Also, it is easy to see that in any case, for each seed and regular function , that (see Equation 2.5) is min-convex in the sense of Definition-Lemma 8.1. Indeed this is the standard (min) tropicalization of a Laurent polynomial. We hope to eventually give a direct geometric description of broken lines (without reference to a scattering diagram) for any log Calabi-Yau, as tropicalizations of some algebraic analogue of holomorphic disks. We expect the conjecture to follow easily from such a description.

Proposition 8.13.

For a global monomial on , the tropicalization is min-convex, and in particular, by Lemma 8.4, decreasing. Further, if the global monomial is of the form for to be an integral point in the interior of a maximal-dimensional cone (see Definition 7.9), then evaluated on monomial decorations strictly increases at any nontrivial bend of a broken line in .

Lemma 8.15.

Under any of the identifications induced by a choice of seed, the set

is a closed convex subset of . The following are equivalent:

(1)

has EGM.

(2)

is bounded, or equivalently, the intersection of the sets for equals .

(3)

There exists a finite number of points such that

is bounded, or equivalently, the intersection of the sets for equals .

(4)

There is function whose associated polytope is bounded.

Proposition 8.16.

Let be fixed data, and let be the Langlands dual data. We write, e.g., for the corresponding lattice for the data as in Appendix A. For each seed , the canonical inclusion

commutes with the tropicalization of mutations and induces an isomorphism

For , the monomial on is a global monomial if and only if on is a global monomial. Finally, has EGM if and only if has EGM.

Proposition 8.17.

For or , suppose there is a compact positive polytope . Assume further that is top dimensional, i.e., . Then for , there are at most finitely many with . These give structure constants for an associative multiplication on

If there is a compact positive polytope then the same conclusion holds for the structure constants (which are all finite) and multiplication rule of for all .

Corollary 8.18.

For or assume has EGM. For assume has EGM. Then defines a -algebra structure on .

Theorem 8.19.

Let or , assume has EGM, and let be a positive polytope, which we assume is rationally defined and not necessarily compact. Then

is a finitely generated -subalgebra.

Corollary 8.21.

For or , suppose that has EGM. For , assume has EGM. Then is a finitely generated -algebra.

Proposition 8.22.

Assume has EGM. Then for each universal Laurent polynomial on , the function of Theorem 6.8 has finite support (i.e., for all but finitely many , and gives inclusions of -algebras

Definition 8.23.

We say a cluster variety has large cluster complex if for some seed , is not contained in a half-space.

Proposition 8.24.

Consider the following conditions on a skew-symmetric cluster algebra :

(1)

has an acyclic seed.

(2)

has a seed with a maximal green sequence (for the definition, see Reference BDP, Def. 1.8

(3)

has large cluster complex.

Then implies implies .

Proposition 8.25.

If has large cluster complex, then has EGM, , and the full Fock–Goncharov conjecture (see Definition 0.6) holds for , , very general and, if the convexity condition of Theorem 0.3 holds, for .

Proposition 8.26.
(1)

Let be an affine variety over a field , and let generators of be a -algebra. For each divisorial discrete valuation (where denotes the function field of which does not have center on (or equivalently, for each boundary divisor in any partial compactification , for some .

(2)

Suppose is a cluster variety, is a smooth affine variety, and is an open immersion. Let generate as a -algebra. Then is strictly negative on .

Proposition 8.27.

If the canonical map

is surjective, then

(1)

is isomorphic to .

(2)

We can choose so that the induced map is an isomorphism.

(3)

The map induced by the choice of in , is finite.

(4)

If furthermore for each we can find a cluster variable with , then (and has EGM. This final condition holds if is finitely generated and is a smooth affine variety.

Remark 8.28.

Every double Bruhat cell is an affine variety by Reference BFZ05, Prop. 2.8 and smooth by Reference FZ99, Theorem 1.1. The surjectivity condition in the statement of Proposition 8.27 holds for all double Bruhat cells by Reference BFZ05, Proposition 2.6 (the proposition states that the exchange matrix has full rank, but the proof shows the surjectivity). So by the proposition, has EGM for double Bruhat cells for which the upper and ordinary cluster algebras are the same. This holds for the open double Bruhat cell of and the base affine space ( maximal unipotent) for by Reference BFZ05, Remark 2.20, and is announced in Reference GY13 for all double Bruhat cells of all semisimple .

Theorem 8.30.

The central fiber of

is the polarized toric variety given by⁠Footnote3 the polyhedron where is the natural map of Proposition B.2.

3

Although is only a rationally defined polyhedron rather than a lattice polyhedron, we can still define .

Theorem 8.32.

Assume that has EGM, is given as above, and that is an algebraically closed field of characteristic zero. Let be one of or . We note has a finitely generated -algebra structure by Corollary 8.18. Define .

Define (constructed above) in case , and for , take instead its fiber over (we are not defining in the case), so by construction we have an open immersion . Define . The following hold:

(1)

In all cases is a Gorenstein scheme with trivial dualizing sheaf.

(2)

For , , or for general, is a -trivial Gorenstein log canonical variety.

(3)

For or for general, or all assuming there exists a seed and a strictly convex cone containing all of for , we have is a minimal model. In other words, is a (in the case relative to projective normal variety, is a reduced Weil divisor, is trivial, and is log canonical.

Lemma 8.33 (Kollár).

Let be an algebraically closed field of characteristic . Let be a proper flat morphism of schemes of finite type over , and let be a closed subscheme which is flat over . Let denote the fiber of over a closed point . Assume that is regular and for the following hold:

(1)

is normal and Cohen–Macaulay.

(2)

is a reduced divisor.

(3)

The pair is log canonical.

(4)

.

(5)

Then the natural morphism is an isomorphism, and there exists a Zariski open neighborhood such that the conditions hold for all . In particular, is a -trivial Gorenstein log canonical variety for all .

Remark 8.34.

Note that directly from its definition, with the multiplication rule counting broken lines, it is difficult to prove anything about , e.g., that it is an integral domain or to determine its dimension. But the convexity, i.e., existence of a convex polytope in the intrinsic sense, gives this very simple degeneration from which we get many properties, at least for very general , for free.

There have been many constructions of degenerations of flag varieties and the like to toric varieties; see Reference AB and references therein. We expect these are all instances of Theorem 8.30.

Many authors have looked for a nice compactification of the moduli space of (say) rank 2 vector bundles with algebraic connection on an algebraic curve . We know of no satisfactory solution. For example, in Reference IIS the case of the complement of four points in is considered, a compactification is constructed, but the boundary is rather nasty (it lies in , but this anticanonical divisor is not reduced). This can be explained as follows: has a different algebraic structure, the character variety, (as complex manifolds they are the same). Note is covered by affine lines (the space of connections on a fixed bundle is an affine space), thus it is not log Calabi–Yau. Rather, it is the log version of uniruled, and there is no Mori theoretic reason to expect a natural compactification. however is log Calabi–Yau, and then by Mori theory one expects (infinitely many) nice compactifications, the minimal models; see Reference GHK13, §1, for an introduction to these ideas. When has punctures, is a cluster variety; see Reference FST and Reference FG06. In the case of with four punctures, is the universal family of affine cubic surfaces (the complement of a triangle of lines on a cubic surface in ); see Reference GHK11, Example 6.12. Each affine cubic has an obvious normal crossing minimal model, the cubic surface. This compactification is an instance of the above, for a natural choice of polygon . The same procedure will give a minimal model compactification for any character variety (of a punctured Riemann surface) by the above simple procedure that has nothing to do with Teichmüller theory.

Theorem 8.35.

In the above situation, the irreducible components of are projective toric varieties. More precisely, for each we have a seed such that is a character on . Then

is a bounded polytope. The associated projective toric variety is an irreducible component of , and all irreducible components of occur in this way.

Definition 9.1.

We say a seed is optimized for if

where

is the composition of canonical identifications defined in §2. If instead , we say is optimized for if for all .

We say is optimized for a frozen index if it is optimized for the corresponding point of .

Lemma 9.2.

In the skew-symmetric case, a seed is optimized for a frozen index if and only if in the quiver for this seed all arrows between unfrozen vertices and the given frozen vertex point toward the given frozen vertex.

Lemma 9.3.
(1)

The seed is optimized for if and only if the monomial on is a global monomial. In this case

and the global monomial is the restriction to of . In the case, for primitive, this holds if and only if each of the initial scattering monomials in is regular along the boundary divisor of corresponding to under the identification .

(2)

has an optimized seed if and only if lies in .

Remark 9.5.

B. LeClerc, and independently L. Shen, gave us an explicit sequence of mutations that shows the proposition holds as well for the cluster structure of Reference BFZ05, Reference GLS on the maximal unipotent subgroup , and the same argument applies to the Fock–Goncharov cluster structure on , . The argument appears in Reference Ma17.

Lemma 9.6.

Let be a lattice, and let be a submonoid with . For any subset and collection of elements such that , the subset is linearly independent over .

Proposition 9.7.

Suppose a valuation has an optimized seed. If , then for all with .

Conjecture 9.8.

The proposition holds for any .

Definition 9.9.

Each choice of seed gives a pairing

which is just the dual pairing composed with the identifications

Lemma 9.10.
(1)

The subspace is a subalgebra containing . If then

(2)

Assume each has an optimized seed. Then

If then .

(3)

If each has an optimized seed and is optimized for , the piecewise linear function

is min-convex, and for all ,

where is the global monomial on corresponding to (which exists by Lemma 9.3).

Lemma 9.12.

The seed is optimized for each of the boundary divisors of .

Proposition 9.16.

Suppose there is a min-convex function , such that implies for integral, and such that for some . Suppose also that there is a bounded positive polytope in (which holds, for example, if has EGM). Then .

Corollary 9.17.

Assume that for each , has an optimized seed, . Let be the (Landau–Ginzburg) potential, the sum of the corresponding global monomials on given by Lemma 9.3. Then:

(1)

The piecewise linear function

is min-convex and

is a positive polytope.

(2)

has the alternative description

(3)

The set

parameterizes a canonical basis of

Corollary 9.18.

Assume we have EGM on , and every frozen variable has an optimized seed. Let and be as in Corollary 9.17. If for some seed , is contained in the convex hull of (which itself contains the integral points of the cluster complex then , is finitely generated, and the integer points parameterize a canonical basis of .

Proposition 10.1.

Suppose we are given fixed skew-symmetric data with no frozen variables along with an acyclic seed . Let be the associated quiver.⁠Footnote4 Each gives a stability in the sense of Reference R10. Assume there is a unique primitive with . For each let

4

Note that because of the assumption made in Appendix A that for any , the quiver has no isolated vertex.

where is the framed moduli space (framed by the vector spaces with unless , in which case of semistable representations of with dimension vector and -slope (see Reference R10, §5.1), and denotes topological Euler characteristic. Let for some . Then

depends only on and (i.e., is independent of the vertex . Furthermore, for arbitrary , (see Lemma 1.9) acts on by

and on by

Example 10.3.

Let be a quiver given by an orientation of the Dynkin diagram of a simply laced finite-dimensional simple Lie algebra. Then the dimension vectors of the indecomposable complex representations of are the positive roots of the associated root system (Gabriel’s theorem). Moreover, for each positive root , there is a unique indecomposable representation with dimension vector , and ; see, e.g., Reference BGP73.

The cluster variety associated to is the cluster variety of finite type associated to the root system Reference FZ03a. Using Proposition 10.1, we can give an explicit description of the scattering diagram for as follows.

First we observe that a representation of that contributes to is a direct sum of copies of an indecomposable representation. Let be a primitive vector, and let be such that . Suppose is an -semistable representation of with dimension vector a multiple of , and consider the decomposition of into indecomposable representations. By -semistability and our assumption , each factor must have dimension vector a multiple of . By Gabriel’s theorem, we see that is a positive root and is a direct sum of copies of the associated indecomposable representation.

We see that the walls of are in bijection with the positive roots of . Let be a positive root, and let be the indecomposable representation with dimension vector . Let be the locus of such that is -semistable of -slope zero; that is, and for the dimension vector of any subrepresentation of . Then is a rational polyhedral cone in , and is nonempty of real codimension . Indeed, there exists such that is -stable by Reference K94, Remark 4.5 and Reference S92, Theorem 6.1, and this is an open condition on . Now let be a point such that . Then the -semistable representations of -slope zero are the direct sums of copies of .

Let us now examine the moduli space . An object in this moduli space is a direct sum of copies of the unique indecomposable representation of dimension vector , along with the framing, a choice of a vector . Such an object is stable if and only if is not contained in a proper subrepresentation of of the form for some subspace . In order for this to be the case, the must be linearly independent elements of , and hence span a -dimensional subspace of . The automorphism group of is , which has the effect of changing the basis of the subspace spanned by . Now it follows easily from the definitions that, for each and such that , the moduli space of -semistable representations with framing at vertex is isomorphic to the Grassmannian .

So, in the notation of Proposition 10.1,

and

Thus the wall of associated to is

For example, suppose is the quiver with vertices , and arrows from 1 to 2 and 2 to 3. This is an orientation of the Dynkin diagram . We have the following isomorphism types of indecomposable representations:

(Here the numbers denote the dimension of the vector space at the vertex, and the symbol over an arrow indicates that the corresponding linear transformation is an isomorphism.) We write and . Then the walls of are

For example, the indecomposable representation with dimension vector has subrepresentations with dimension vectors and . So the associated wall has support defined by the inequalities and . This gives the last wall in the list.

Equation (A.1)
Equation (A.3)
Equation (B.1)
Proposition B.2.

Giving fixed data , we have:

(1)

There is a commutative diagram where the dotted arrows are only present if there are no frozen variables (i.e.,

with any point, the identity, and with the left- and right-hand squares cartesian and an isomorphism, canonical if there are no frozen variables.

(2)

There are torus actions

Here is the kernel of the map

Furthermore and act on so that the map is - and -equivariant. The map is a -torsor. There is a map such that the map is also compatible with the actions of these two tori on and respectively, so that

is a -torsor.

(3)

and are isomorphic data, so we can define

(4)

There is a commutative diagram:

Equation (B.4)
Lemma B.6.

The choice of the map gives an inclusion (see Proposition B.2 given by . We also have (a sublattice of included in via . These inclusions induce an isomorphism .

Lemma B.7.

The map induced by the composition of the inclusion and projection is a split surjection if and only if the map

is surjective. This holds if and only if in some seed , the matrix with entries for , , gives a surjective map . In this case is isomorphic to the trivial bundle .

Remark B.8.

In general, a seed is defined to be a basis of the lattice (or ), but to define the seed mutations Reference GHK13, (2.2) and the union of tori Equation A.3, all one needs are elements , (the definitions as given make sense even if the are dependent or fail to span). If one makes the construction in this greater generality, the characters on will not be independent (if the are not), and unless we take a full basis, we cannot define the cluster variables on , as the are defined as the dual basis to the basis for .

In the case of the principal data, given a seed for , we get a seed in this modified sense for the data . We also write this seed as . On the other hand, in Reference GHK13, the seed for is defined in the more traditional sense to be the basis . It is not the case that if is obtained from via a sequence of mutations, then is obtained from by the same sequence of mutations. In particular, the set of seeds mutation equivalent to depends not just on the mutation equivalence class of , but on the original seed . However, using the seed as a seed for in this modified sense, we can build , and this depends only on the mutation class of . Thus does not depend on the initial choice of seed, but only on its mutation equivalence class.

However, as we shall now see, the choice of initial seed does give a partial compactification. This is a more general phenomenon when there are frozen variables.

Construction B.9 (Partial compactifications from frozen variables).

When the cluster data includes frozen variables, comes with a canonical partial compactification , given by partially compactifying each torus chart via , where for , . Thus the dual cone is cut out by the half-spaces , . Note that the monomials , are invariant under mutation. These give a canonical map , where is the number of unfrozen variables. Note that the basis elements for , though they have frozen indices, can change under mutation. What is invariant is the associated boundary divisor with valuation given by . These are the boundary divisors of . We remark that like , is also separated, with the argument given in Reference GHK13, Theorem 3.14 working equally well for .

Here is another way of seeing the same thing. Given any cluster variety and a single fan for a toric partial compactification for some , there is a canonical way to build a partial compactification

We let and , where is the birational map given by the composition

and is the geometric tropicalization; see §2.

Remark B.10.

We now return to the discussion of . Note that the frozen variables for are indexed by in the first copy of , along with all indices in the second copy of . However, we can apply Construction B.9 taking only the second copy of as the set of frozen indices, with the initial choice of seed determining a partial compactification of . In this case, we indicate the partial compactification by . It is important to keep in mind the dependence on . Fixing fixes , and hence cluster variables , . The variables can then take the value in the compactification. In particular, we obtain an extension of to , pulling back to .

Note that the seeds in and are in one-to-one correspondence. Given any seed and seed obtained via the same sequence of mutations, we have for some . These two seeds give rise to coordinates on the chart of indexed by and coordinates on the chart of indexed by . As is the fiber of over the point of with all coordinates , the coordinate on the chart of restricts to the coordinate on the chart of . This gives a one-to-one correspondence between cluster variables on and -type cluster variables on . To summarize:

Proposition B.11.

The cluster variety depends only on the mutation class . But the choice of a seed determines:

(1)

a partial compactification

(2)

the canonical extension of each cluster variable on any chart of to a cluster variable on the corresponding chart of .

Construction C.1.

There is a simple order by order algorithm, introduced in Reference KS06 in the two-dimensional case and in Reference GS11 in the higher-dimensional case, for producing the diagram of Theorem 1.21, which we will describe shortly after a bit of preparation. This is useful both from a computational point of view and because a more complicated version of this will be necessary in the remainder of this appendix.

We continue with fixed data , yielding the Lie algebra in §1.1.

We first introduce some additional terminology. For any scattering diagram for , and any we let be the (by definition, finite) set of with nontrivial in . A scattering diagram for , induces a scattering diagram for , in the obvious way, viewing for a wall . We say two scattering diagrams , are equivalent to order if they are equivalent as scattering diagrams for .

Definition-Lemma C.2.

Let be a joint of the scattering diagram . Either every wall containing has direction tangent to (where the direction of a wall contained in is or every wall containing has direction not tangent to . In the first case we call the joint parallel; in the second case, perpendicular.

Proof.

Suppose spans the subspace . Then the direction of any wall containing is of the form for some . If this is tangent to , then for , and hence . From this it follows that for all , and hence the direction of any wall containing is tangent to .

A joint is a codimension 2 convex rational polyhedral cone. Let be the set of integral tangent vectors to . This is a saturated sublattice of . Then we set

This is closed under Lie bracket. If is a parallel joint, then is abelian, since if with , , so . We denote by the corresponding group.

We will build a sequence of finite scattering diagrams , with the property that is equivalent to to order . Taking , we obtain equivalent to . Let denote the subset of consisting of walls which are nontrivial in . We start with

If is a joint of a finite scattering diagram, we write for a simple loop around small enough so that it only intersects walls containing . In particular, for each joint of , . Indeed, is abelian and by the form given for in the statement of Theorem 1.21, all walls containing are hyperplanes. Thus the automorphism associated to crossing each wall and its inverse occur once in , and hence cancel.

Now suppose we have constructed . For every perpendicular joint of , we can write uniquely in

where and . Such an expression holds because all wall-crossing automorphisms for walls containing lie in , so that can be viewed as an element of . Furthermore, by the inductive hypothesis, this element is trivial in . Because is perpendicular, we never have . Now define

where the sign is chosen so that the contribution to crossing the wall indexed by in is . Note the latter element is central in . Thus and

in .

We define

where the union is over all perpendicular joints of .

Lemma C.6.

is equivalent to to order .

Proof.

Consider a perpendicular joint of . If is contained in a joint of , is the unique such joint, and we constructed above. If is not contained in a joint of , we define to be the empty set. There are three types of walls in containing :

(1)

.

(2)

, but . This type of wall does not contribute to , as the associated automorphism is central in , and in addition this wall contributes twice to , with the two contributions inverse to each other.

(3)

and . Since each added wall is of the form for some joint of , where is the direction of the wall, the direction of the wall is parallel to , contradicting being a perpendicular joint. Thus this does not occur.

From this, it is clear that , which is the identity in by C.5. This holds for every perpendicular joint of .

The result then follows from Lemma C.7.

Lemma C.7.

Let and be two scattering diagrams for such that

(1)

and are equivalent to order .

(2)

is consistent to order .

(3)

is the identity for every perpendicular joint of to order .

(4)

and have the same set of incoming walls.

Then and are equivalent to order , and in particular is consistent to order .

Lemma C.8.

Let (resp. ), and let (resp. ) be an automorphism of the form for . Let . If , then

while if , then

Here, we view or as automorphisms of and or as subgroups of the group of automorphisms of this ring.

Equation (C.10)
Equation (C.11)
Equation (C.12)
Proposition C.13.

In the above setup, suppose given , , with , and positive integers . Consider the scattering diagram

. We can choose within its equivalence class so that for any given ray , we have

for a positive integer and the nonnegative integers with at least two of them nonzero.

Equation (C.14)
Corollary C.15.

In the situation of Proposition C.13, suppose instead that

where now , the ground field. Choosing up to equivalence, we can assume that each ray satisfies

for some choice of nonnegative integers and where is a positive integer.

References

Reference [AB]
V. Alexeev and M. Brion, Toric degenerations of spherical varieties, Selecta Math. (N.S.) 10 (2004), no. 4, 453–478, DOI 10.1007/s00029-005-0396-8. MR2134452,
Show rawAMSref \bib{AB}{article}{ label={AB}, author={Alexeev, Valery}, author={Brion, Michel}, title={Toric degenerations of spherical varieties}, journal={Selecta Math. (N.S.)}, volume={10}, date={2004}, number={4}, pages={453--478}, issn={1022-1824}, review={\MR {2134452}}, doi={10.1007/s00029-005-0396-8}, }
Reference [A07]
D. Auroux, Mirror symmetry and -duality in the complement of an anticanonical divisor, J. Gökova Geom. Topol. GGT 1 (2007), 51–91. MR2386535,
Show rawAMSref \bib{A07}{article}{ label={A07}, author={Auroux, Denis}, title={Mirror symmetry and $T$-duality in the complement of an anticanonical divisor}, journal={J. G\"okova Geom. Topol. GGT}, volume={1}, date={2007}, pages={51--91}, issn={1935-2565}, review={\MR {2386535}}, }
Reference [BK00]
A. Berenstein and D. Kazhdan, Geometric and unipotent crystals, Geom. Funct. Anal., Special Volume, Part I, (2000), 188–236, DOI 10.1007/978-3-0346-0422-2_8. GAFA 2000 (Tel Aviv, 1999). MR1826254,
Show rawAMSref \bib{BK00}{article}{ label={BK00}, author={Berenstein, Arkady}, author={Kazhdan, David}, title={Geometric and unipotent crystals}, note={GAFA 2000 (Tel Aviv, 1999)}, journal={Geom. Funct. Anal., Special Volume, Part I, (2000)}, pages={188--236}, issn={1016-443X}, review={\MR {1826254}}, doi={10.1007/978-3-0346-0422-2\_8}, }
Reference [BK07]
A. Berenstein and D. Kazhdan, Geometric and unipotent crystals. II. From unipotent bicrystals to crystal bases, Quantum groups, Contemp. Math., vol. 433, Amer. Math. Soc., Providence, RI, 2007, pp. 13–88, DOI 10.1090/conm/433/08321. MR2349617,
Show rawAMSref \bib{BK07}{article}{ label={BK07}, author={Berenstein, Arkady}, author={Kazhdan, David}, title={Geometric and unipotent crystals. II. From unipotent bicrystals to crystal bases}, conference={ title={Quantum groups}, }, book={ series={Contemp. Math.}, volume={433}, publisher={Amer. Math. Soc., Providence, RI}, }, date={2007}, pages={13--88}, review={\MR {2349617}}, doi={10.1090/conm/433/08321}, }
Reference [BZ01]
A. Berenstein and A. Zelevinsky, Tensor product multiplicities, canonical bases and totally positive varieties, Invent. Math. 143 (2001), no. 1, 77–128, DOI 10.1007/s002220000102. MR1802793,
Show rawAMSref \bib{BZ01}{article}{ label={BZ01}, author={Berenstein, Arkady}, author={Zelevinsky, Andrei}, title={Tensor product multiplicities, canonical bases and totally positive varieties}, journal={Invent. Math.}, volume={143}, date={2001}, number={1}, pages={77--128}, issn={0020-9910}, review={\MR {1802793}}, doi={10.1007/s002220000102}, }
Reference [BFZ05]
A. Berenstein, S. Fomin, and A. Zelevinsky, Cluster algebras. III. Upper bounds and double Bruhat cells, Duke Math. J. 126 (2005), no. 1, 1–52, DOI 10.1215/S0012-7094-04-12611-9. MR2110627,
Show rawAMSref \bib{BFZ05}{article}{ label={BFZ05}, author={Berenstein, Arkady}, author={Fomin, Sergey}, author={Zelevinsky, Andrei}, title={Cluster algebras. III. Upper bounds and double Bruhat cells}, journal={Duke Math. J.}, volume={126}, date={2005}, number={1}, pages={1--52}, issn={0012-7094}, review={\MR {2110627}}, doi={10.1215/S0012-7094-04-12611-9}, }
Reference [BGP73]
I. N. Bernšteĭn, I. M. Gel′fand, and V. A. Ponomarev, Coxeter functors, and Gabriel’s theorem (Russian), Uspehi Mat. Nauk 28 (1973), no. 2(170), 19–33. MR0393065,
Show rawAMSref \bib{BGP73}{article}{ label={BGP73}, author={Bern\v ste\u \i n, I. N.}, author={Gel\cprime fand, I. M.}, author={Ponomarev, V. A.}, title={Coxeter functors, and Gabriel's theorem}, language={Russian}, journal={Uspehi Mat. Nauk}, volume={28}, date={1973}, number={2(170)}, pages={19--33}, issn={0042-1316}, review={\MR {0393065}}, }
Reference [B90]
A. I. Bondal, Helices, representations of quivers and Koszul algebras, Helices and vector bundles, London Math. Soc. Lecture Note Ser., vol. 148, Cambridge Univ. Press, Cambridge, 1990, pp. 75–95, DOI 10.1017/CBO9780511721526.008. MR1074784,
Show rawAMSref \bib{B90}{article}{ label={B90}, author={Bondal, A. I.}, title={Helices, representations of quivers and Koszul algebras}, conference={ title={Helices and vector bundles}, }, book={ series={London Math. Soc. Lecture Note Ser.}, volume={148}, publisher={Cambridge Univ. Press, Cambridge}, }, date={1990}, pages={75--95}, review={\MR {1074784}}, doi={10.1017/CBO9780511721526.008}, }
Reference [Bri]
T. Bridgeland, Scattering diagrams, Hall algebras and stability conditions, preprint, 2016.
Reference [BDP]
T. Brüstle, G. Dupont, and M. Pérotin, On maximal green sequences, Int. Math. Res. Not. IMRN 16 (2014), 4547–4586, DOI 10.1093/imrn/rnt075. MR3250044,
Show rawAMSref \bib{BDP}{article}{ label={BDP}, author={Br\"ustle, Thomas}, author={Dupont, Gr\'egoire}, author={P\'erotin, Matthieu}, title={On maximal green sequences}, journal={Int. Math. Res. Not. IMRN}, date={2014}, number={16}, pages={4547--4586}, issn={1073-7928}, review={\MR {3250044}}, doi={10.1093/imrn/rnt075}, }
Reference [C02]
P. Caldero, Toric degenerations of Schubert varieties, Transform. Groups 7 (2002), no. 1, 51–60, DOI 10.1007/s00031-002-0003-4. MR1888475,
Show rawAMSref \bib{C02}{article}{ label={C02}, author={Caldero, Philippe}, title={Toric degenerations of Schubert varieties}, journal={Transform. Groups}, volume={7}, date={2002}, number={1}, pages={51--60}, issn={1083-4362}, review={\MR {1888475}}, doi={10.1007/s00031-002-0003-4}, }
Reference [CLS]
I. Canakci, K. Lee, and R. Schiffler, On cluster algebras from unpunctured surfaces with one marked point, Proc. Amer. Math. Soc. Ser. B 2 (2015), 35–49, DOI 10.1090/bproc/21. MR3422667,
Show rawAMSref \bib{CLS}{article}{ label={CLS}, author={Canakci, Ilke}, author={Lee, Kyungyong}, author={Schiffler, Ralf}, title={On cluster algebras from unpunctured surfaces with one marked point}, journal={Proc. Amer. Math. Soc. Ser. B}, volume={2}, date={2015}, pages={35--49}, issn={2330-1511}, review={\MR {3422667}}, doi={10.1090/bproc/21}, }
Reference [CPS]
M. Carl, M. Pumperla, and B. Siebert, A tropical view of Landau-Ginzburg models, available at http://www.math.uni-hamburg.de/home/siebert/preprints/LGtrop.pdf
Reference [CKLP]
G. Cerulli Irelli, B. Keller, D. Labardini-Fragoso, and P.-G. Plamondon, Linear independence of cluster monomials for skew-symmetric cluster algebras, Compos. Math. 149 (2013), no. 10, 1753–1764, DOI 10.1112/S0010437X1300732X. MR3123308,
Show rawAMSref \bib{CKLP}{article}{ label={CKLP}, 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--1764}, issn={0010-437X}, review={\MR {3123308}}, doi={10.1112/S0010437X1300732X}, }
Reference [CGMMRSW]
M. W. Cheung, M. Gross, G. Muller, G. Musiker, D. Rupel, S. Stella, and H. Williams, The greedy basis equals the theta basis: a rank two haiku, J. Combin. Theory Ser. A 145 (2017), 150–171, DOI 10.1016/j.jcta.2016.08.004. MR3551649,
Show rawAMSref \bib{snowbird}{article}{ label={CGMMRSW}, author={Cheung, Man Wai}, author={Gross, Mark}, author={Muller, Greg}, author={Musiker, Gregg}, author={Rupel, Dylan}, author={Stella, Salvatore}, author={Williams,~Harold}, title={The greedy basis equals the theta basis: a rank two haiku}, journal={J. Combin. Theory Ser. A}, volume={145}, date={2017}, pages={150--171}, issn={0097-3165}, review={\MR {3551649}}, doi={10.1016/j.jcta.2016.08.004}, }
Reference [CO06]
C.-H. Cho and Y.-G. Oh, Floer cohomology and disc instantons of Lagrangian torus fibers in Fano toric manifolds, Asian J. Math. 10 (2006), no. 4, 773–814, DOI 10.4310/AJM.2006.v10.n4.a10. MR2282365,
Show rawAMSref \bib{CO06}{article}{ label={CO06}, author={Cho, Cheol-Hyun}, author={Oh, Yong-Geun}, title={Floer cohomology and disc instantons of Lagrangian torus fibers in Fano toric manifolds}, journal={Asian J. Math.}, volume={10}, date={2006}, number={4}, pages={773--814}, issn={1093-6106}, review={\MR {2282365}}, doi={10.4310/AJM.2006.v10.n4.a10}, }
Reference [FG06]
V. Fock and A. Goncharov, Moduli spaces of local systems and higher Teichmüller theory, Publ. Math. Inst. Hautes Études Sci. 103 (2006), 1–211, DOI 10.1007/s10240-006-0039-4. MR2233852,
Show rawAMSref \bib{FG06}{article}{ label={FG06}, author={Fock, Vladimir}, author={Goncharov, Alexander}, title={Moduli spaces of local systems and higher Teichm\"uller theory}, journal={Publ. Math. Inst. Hautes \'Etudes Sci.}, number={103}, date={2006}, pages={1--211}, issn={0073-8301}, review={\MR {2233852}}, doi={10.1007/s10240-006-0039-4}, }
Reference [FG09]
V. V. Fock and A. B. Goncharov, Cluster ensembles, quantization and the dilogarithm (English, with English and French summaries), Ann. Sci. Éc. Norm. Supér. (4) 42 (2009), no. 6, 865–930, DOI 10.1007/978-0-8176-4745-2_15. MR2567745,
Show rawAMSref \bib{FG09}{article}{ label={FG09}, author={Fock, Vladimir V.}, author={Goncharov, Alexander B.}, title={Cluster ensembles, quantization and the dilogarithm}, language={English, with English and French summaries}, journal={Ann. Sci. \'Ec. Norm. Sup\'er. (4)}, volume={42}, date={2009}, number={6}, pages={865--930}, issn={0012-9593}, review={\MR {2567745}}, doi={10.1007/978-0-8176-4745-2\_15}, }
Reference [FG11]
V. Fock and A. Goncharov, Cluster -varieties at infinity, preprint, 2011.
Reference [FST]
S. Fomin, M. Shapiro, and D. Thurston, Cluster algebras and triangulated surfaces. I. Cluster complexes, Acta Math. 201 (2008), no. 1, 83–146, DOI 10.1007/s11511-008-0030-7. MR2448067,
Show rawAMSref \bib{FST}{article}{ label={FST}, author={Fomin, Sergey}, author={Shapiro, Michael}, author={Thurston, Dylan}, title={Cluster algebras and triangulated surfaces. I. Cluster complexes}, journal={Acta Math.}, volume={201}, date={2008}, number={1}, pages={83--146}, issn={0001-5962}, review={\MR {2448067}}, doi={10.1007/s11511-008-0030-7}, }
Reference [FZ99]
S. Fomin and A. Zelevinsky, Double Bruhat cells and total positivity, J. Amer. Math. Soc. 12 (1999), no. 2, 335–380, DOI 10.1090/S0894-0347-99-00295-7. MR1652878,
Show rawAMSref \bib{FZ99}{article}{ label={FZ99}, author={Fomin, Sergey}, author={Zelevinsky, Andrei}, title={Double Bruhat cells and total positivity}, journal={J. Amer. Math. Soc.}, volume={12}, date={1999}, number={2}, pages={335--380}, issn={0894-0347}, review={\MR {1652878}}, doi={10.1090/S0894-0347-99-00295-7}, }
Reference [FZ02a]
S. Fomin and A. Zelevinsky, Cluster algebras. I. Foundations, J. Amer. Math. Soc. 15 (2002), no. 2, 497–529, DOI 10.1090/S0894-0347-01-00385-X. MR1887642,
Show rawAMSref \bib{FZ02a}{article}{ label={FZ02a}, author={Fomin, Sergey}, author={Zelevinsky, Andrei}, title={Cluster algebras. I. Foundations}, journal={J. Amer. Math. Soc.}, volume={15}, date={2002}, number={2}, pages={497--529}, issn={0894-0347}, review={\MR {1887642}}, doi={10.1090/S0894-0347-01-00385-X}, }
Reference [FZ02b]
S. Fomin and A. Zelevinsky, The Laurent phenomenon, Adv. in Appl. Math. 28 (2002), no. 2, 119–144, DOI 10.1006/aama.2001.0770. MR1888840,
Show rawAMSref \bib{FZ02b}{article}{ label={FZ02b}, author={Fomin, Sergey}, author={Zelevinsky, Andrei}, title={The Laurent phenomenon}, journal={Adv. in Appl. Math.}, volume={28}, date={2002}, number={2}, pages={119--144}, issn={0196-8858}, review={\MR {1888840}}, doi={10.1006/aama.2001.0770}, }
Reference [FZ03a]
S. Fomin and A. Zelevinsky, Cluster algebras. II. Finite type classification, Invent. Math. 154 (2003), no. 1, 63–121, DOI 10.1007/s00222-003-0302-y. MR2004457,
Show rawAMSref \bib{FZ03a}{article}{ label={FZ03a}, author={Fomin, Sergey}, author={Zelevinsky, Andrei}, title={Cluster algebras. II. Finite type classification}, journal={Invent. Math.}, volume={154}, date={2003}, number={1}, pages={63--121}, issn={0020-9910}, review={\MR {2004457}}, doi={10.1007/s00222-003-0302-y}, }
Reference [FZ07]
S. Fomin and A. Zelevinsky, Cluster algebras. IV. Coefficients, Compos. Math. 143 (2007), no. 1, 112–164, DOI 10.1112/S0010437X06002521. MR2295199,
Show rawAMSref \bib{FZ07}{article}{ label={FZ07}, author={Fomin, Sergey}, author={Zelevinsky, Andrei}, title={Cluster algebras. IV. Coefficients}, journal={Compos. Math.}, volume={143}, date={2007}, number={1}, pages={112--164}, issn={0010-437X}, review={\MR {2295199}}, doi={10.1112/S0010437X06002521}, }
Reference [GLS]
C. Geiss, B. Leclerc, and J. Schröer, Partial flag varieties and preprojective algebras (English, with English and French summaries), Ann. Inst. Fourier (Grenoble) 58 (2008), no. 3, 825–876. MR2427512,
Show rawAMSref \bib{GLS}{article}{ label={GLS}, author={Geiss, Christof}, author={Leclerc, Bernard}, author={Schr\"oer, Jan}, title={Partial flag varieties and preprojective algebras}, language={English, with English and French summaries}, journal={Ann. Inst. Fourier (Grenoble)}, volume={58}, date={2008}, number={3}, pages={825--876}, issn={0373-0956}, review={\MR {2427512}}, }
Reference [GSV]
M. Gekhtman, M. Shapiro, and A. Vainshtein, Cluster algebras and Poisson geometry, Mathematical Surveys and Monographs, vol. 167, American Mathematical Society, Providence, RI, 2010, DOI 10.1090/surv/167. MR2683456,
Show rawAMSref \bib{GSV}{book}{ label={GSV}, author={Gekhtman, Michael}, author={Shapiro, Michael}, author={Vainshtein, Alek}, title={Cluster algebras and Poisson geometry}, series={Mathematical Surveys and Monographs}, volume={167}, publisher={American Mathematical Society, Providence, RI}, date={2010}, pages={xvi+246}, isbn={978-0-8218-4972-9}, review={\MR {2683456}}, doi={10.1090/surv/167}, }
Reference [GS13]
A. Goncharov and L. Shen, Geometry of canonical bases and mirror symmetry, Invent. Math. 202 (2015), no. 2, 487–633, DOI 10.1007/s00222-014-0568-2. MR3418241,
Show rawAMSref \bib{GS13}{article}{ label={GS13}, author={Goncharov, Alexander}, author={Shen, Linhui}, title={Geometry of canonical bases and mirror symmetry}, journal={Invent. Math.}, volume={202}, date={2015}, number={2}, pages={487--633}, issn={0020-9910}, review={\MR {3418241}}, doi={10.1007/s00222-014-0568-2}, }
Reference [GS16]
A. Goncharov and L. Shen, Donaldson-Thomas transformations of moduli spaces of -local systems, preprint 2016, arXiv:1602.06479.
Reference [GY13]
K. R. Goodearl and M. T. Yakimov, Quantum cluster algebras and quantum nilpotent algebras, Proc. Natl. Acad. Sci. USA 111 (2014), no. 27, 9696–9703, DOI 10.1073/pnas.1313071111. MR3263301,
Show rawAMSref \bib{GY13}{article}{ label={GY13}, author={Goodearl, Kenneth R.}, author={Yakimov, Milen T.}, title={Quantum cluster algebras and quantum nilpotent algebras}, journal={Proc. Natl. Acad. Sci. USA}, volume={111}, date={2014}, number={27}, pages={9696--9703}, issn={1091-6490}, review={\MR {3263301}}, doi={10.1073/pnas.1313071111}, }
Reference [G09]
M. Gross, Mirror symmetry for and tropical geometry, Adv. Math. 224 (2010), no. 1, 169–245, DOI 10.1016/j.aim.2009.11.007. MR2600995,
Show rawAMSref \bib{G09}{article}{ label={G09}, author={Gross, Mark}, title={Mirror symmetry for $\mathbb {P}^2$ and tropical geometry}, journal={Adv. Math.}, volume={224}, date={2010}, number={1}, pages={169--245}, issn={0001-8708}, review={\MR {2600995}}, doi={10.1016/j.aim.2009.11.007}, }
Reference [G11]
M. Gross, Tropical geometry and mirror symmetry, CBMS Regional Conference Series in Mathematics, vol. 114, Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI, 2011, DOI 10.1090/cbms/114. MR2722115,
Show rawAMSref \bib{G11}{book}{ label={G11}, author={Gross, Mark}, title={Tropical geometry and mirror symmetry}, series={CBMS Regional Conference Series in Mathematics}, volume={114}, publisher={Published for the Conference Board of the Mathematical Sciences, Washington, DC; by the American Mathematical Society, Providence, RI}, date={2011}, pages={xvi+317}, isbn={978-0-8218-5232-3}, review={\MR {2722115}}, doi={10.1090/cbms/114}, }
Reference [GHK11]
M. Gross, P. Hacking, and S. Keel, Mirror symmetry for log Calabi-Yau surfaces I, Publ. Math. Inst. Hautes Études Sci. 122 (2015), 65–168, DOI 10.1007/s10240-015-0073-1. MR3415066,
Show rawAMSref \bib{GHK11}{article}{ label={GHK11}, author={Gross, Mark}, author={Hacking, Paul}, author={Keel, Sean}, title={Mirror symmetry for log Calabi-Yau surfaces I}, journal={Publ. Math. Inst. Hautes \'Etudes Sci.}, volume={122}, date={2015}, pages={65--168}, issn={0073-8301}, review={\MR {3415066}}, doi={10.1007/s10240-015-0073-1}, }
Reference [GHK12]
M. Gross, P. Hacking, and S. Keel, Moduli of surfaces with an anti-canonical cycle, Compos. Math. 151 (2015), no. 2, 265–291, DOI 10.1112/S0010437X14007611. MR3314827,
Show rawAMSref \bib{GHK12}{article}{ label={GHK12}, author={Gross, Mark}, author={Hacking, Paul}, author={Keel, Sean}, title={Moduli of surfaces with an anti-canonical cycle}, journal={Compos. Math.}, volume={151}, date={2015}, number={2}, pages={265--291}, issn={0010-437X}, review={\MR {3314827}}, doi={10.1112/S0010437X14007611}, }
Reference [GHK13]
M. Gross, P. Hacking, and S. Keel, Birational geometry of cluster algebras, Algebr. Geom. 2 (2015), no. 2, 137–175, DOI 10.14231/AG-2015-007. MR3350154,
Show rawAMSref \bib{P1}{article}{ label={GHK13}, author={Gross, Mark}, author={Hacking, Paul}, author={Keel, Sean}, title={Birational geometry of cluster algebras}, journal={Algebr. Geom.}, volume={2}, date={2015}, number={2}, pages={137--175}, issn={2214-2584}, review={\MR {3350154}}, doi={10.14231/AG-2015-007}, }
Reference [GHKII]
M. Gross, P. Hacking and S. Keel, Mirror symmetry for log Calabi-Yau surfaces II, in preparation.
Reference [GHKS]
M. Gross, P. Hacking, S. Keel, and B. Siebert, Theta functions on varieties with effective anti-canonical class, preprint, 2016.
Reference [GP10]
M. Gross and R. Pandharipande, Quivers, curves, and the tropical vertex, Port. Math. 67 (2010), no. 2, 211–259, DOI 10.4171/PM/1865. MR2662867,
Show rawAMSref \bib{GP10}{article}{ label={GP10}, author={Gross, Mark}, author={Pandharipande, Rahul}, title={Quivers, curves, and the tropical vertex}, journal={Port. Math.}, volume={67}, date={2010}, number={2}, pages={211--259}, issn={0032-5155}, review={\MR {2662867}}, doi={10.4171/PM/1865}, }
Reference [GPS]
M. Gross, R. Pandharipande, and B. Siebert, The tropical vertex, Duke Math. J. 153 (2010), no. 2, 297–362, DOI 10.1215/00127094-2010-025. MR2667135,
Show rawAMSref \bib{GPS}{article}{ label={GPS}, author={Gross, Mark}, author={Pandharipande, Rahul}, author={Siebert, Bernd}, title={The tropical vertex}, journal={Duke Math. J.}, volume={153}, date={2010}, number={2}, pages={297--362}, issn={0012-7094}, review={\MR {2667135}}, doi={10.1215/00127094-2010-025}, }
Reference [GS11]
M. Gross and B. Siebert, From real affine geometry to complex geometry, Ann. of Math. (2) 174 (2011), no. 3, 1301–1428, DOI 10.4007/annals.2011.174.3.1. MR2846484,
Show rawAMSref \bib{GSAnnals}{article}{ label={GS11}, author={Gross, Mark}, author={Siebert, Bernd}, title={From real affine geometry to complex geometry}, journal={Ann. of Math. (2)}, volume={174}, date={2011}, number={3}, pages={1301--1428}, issn={0003-486X}, review={\MR {2846484}}, doi={10.4007/annals.2011.174.3.1}, }
Reference [GS12]
M. Gross and B. Siebert, Theta functions and mirror symmetry, Surveys in differential geometry 2016. Advances in geometry and mathematical physics, Surv. Differ. Geom., vol. 21, Int. Press, Somerville, MA, 2016, pp. 95–138. MR3525095,
Show rawAMSref \bib{GS12}{article}{ label={GS12}, author={Gross, Mark}, author={Siebert, Bernd}, title={Theta functions and mirror symmetry}, conference={ title={Surveys in differential geometry 2016. Advances in geometry and mathematical physics}, }, book={ series={Surv. Differ. Geom.}, volume={21}, publisher={Int. Press, Somerville, MA}, }, date={2016}, pages={95--138}, review={\MR {3525095}}, }
Reference [IIS]
M.-a. Inaba, K. Iwasaki, and M.-H. Saito, Moduli of stable parabolic connections, Riemann-Hilbert correspondence and geometry of Painlevé equation of type VI. I, Publ. Res. Inst. Math. Sci. 42 (2006), no. 4, 987–1089. MR2289083,
Show rawAMSref \bib{IIS}{article}{ label={IIS}, author={Inaba, Michi-aki}, author={Iwasaki, Katsunori}, author={Saito, Masa-Hiko}, title={Moduli of stable parabolic connections, Riemann-Hilbert correspondence and geometry of Painlev\'e equation of type VI. I}, journal={Publ. Res. Inst. Math. Sci.}, volume={42}, date={2006}, number={4}, pages={987--1089}, issn={0034-5318}, review={\MR {2289083}}, }
Reference [K80]
V. G. Kac, Infinite root systems, representations of graphs and invariant theory, Invent. Math. 56 (1980), no. 1, 57–92, DOI 10.1007/BF01403155. MR557581,
Show rawAMSref \bib{K80}{article}{ label={K80}, author={Kac, V. G.}, title={Infinite root systems, representations of graphs and invariant theory}, journal={Invent. Math.}, volume={56}, date={1980}, number={1}, pages={57--92}, issn={0020-9910}, review={\MR {557581}}, doi={10.1007/BF01403155}, }
Reference [K82]
V. G. Kac, Infinite root systems, representations of graphs and invariant theory. II, J. Algebra 78 (1982), no. 1, 141–162, DOI 10.1016/0021-8693(82)90105-3. MR677715,
Show rawAMSref \bib{K82}{article}{ label={K82}, author={Kac, V. G.}, title={Infinite root systems, representations of graphs and invariant theory. II}, journal={J. Algebra}, volume={78}, date={1982}, number={1}, pages={141--162}, issn={0021-8693}, review={\MR {677715}}, doi={10.1016/0021-8693(82)90105-3}, }
Reference [K94]
A. D. King, Moduli of representations of finite-dimensional algebras, Quart. J. Math. Oxford Ser. (2) 45 (1994), no. 180, 515–530, DOI 10.1093/qmath/45.4.515. MR1315461,
Show rawAMSref \bib{K94}{article}{ label={K94}, author={King, A. D.}, title={Moduli of representations of finite-dimensional algebras}, journal={Quart. J. Math. Oxford Ser. (2)}, volume={45}, date={1994}, number={180}, pages={515--530}, issn={0033-5606}, review={\MR {1315461}}, doi={10.1093/qmath/45.4.515}, }
Reference [KT99]
A. Knutson and T. Tao, The honeycomb model of tensor products. I. Proof of the saturation conjecture, J. Amer. Math. Soc. 12 (1999), no. 4, 1055–1090, DOI 10.1090/S0894-0347-99-00299-4. MR1671451,
Show rawAMSref \bib{KT99}{article}{ label={KT99}, author={Knutson, Allen}, author={Tao, Terence}, title={The honeycomb model of ${\rm GL}_n({\bf C})$ tensor products. I. Proof of the saturation conjecture}, journal={J. Amer. Math. Soc.}, volume={12}, date={1999}, number={4}, pages={1055--1090}, issn={0894-0347}, review={\MR {1671451}}, doi={10.1090/S0894-0347-99-00299-4}, }
Reference [KM05]
M. Kogan and E. Miller, Toric degeneration of Schubert varieties and Gelfand-Tsetlin polytopes, Adv. Math. 193 (2005), no. 1, 1–17, DOI 10.1016/j.aim.2004.03.017. MR2132758,
Show rawAMSref \bib{KM05}{article}{ label={KM05}, author={Kogan, Mikhail}, author={Miller, Ezra}, title={Toric degeneration of Schubert varieties and Gelfand-Tsetlin polytopes}, journal={Adv. Math.}, volume={193}, date={2005}, number={1}, pages={1--17}, issn={0001-8708}, review={\MR {2132758}}, doi={10.1016/j.aim.2004.03.017}, }
Reference [K13]
J. Kollár, Singularities of the minimal model program, Cambridge Tracts in Mathematics, vol. 200, Cambridge University Press, Cambridge, 2013. With a collaboration of Sándor Kovács, DOI 10.1017/CBO9781139547895. MR3057950,
Show rawAMSref \bib{K13}{book}{ label={K13}, author={Koll\'ar, J\'anos}, title={Singularities of the minimal model program}, series={Cambridge Tracts in Mathematics}, volume={200}, note={With a collaboration of S\'andor Kov\'acs}, publisher={Cambridge University Press, Cambridge}, date={2013}, pages={x+370}, isbn={978-1-107-03534-8}, review={\MR {3057950}}, doi={10.1017/CBO9781139547895}, }
Reference [KS06]
M. Kontsevich and Y. Soibelman, Affine structures and non-Archimedean analytic spaces, in The unity of mathematics, Progr. Math., vol. 244, Birkhäuser Boston, Boston, MA, 2006, pp. 321–385, DOI 10.1007/0-8176-4467-9_9. MR2181810,
Show rawAMSref \bib{KS06}{article}{ label={KS06}, author={Kontsevich, Maxim}, author={Soibelman, Yan}, title={Affine structures and non-Archimedean analytic spaces}, conference={ title={in The unity of mathematics}, }, book={ series={Progr. Math.}, volume={244}, publisher={Birkh\"auser Boston, Boston, MA}, }, date={2006}, pages={321--385}, review={\MR {2181810}}, doi={10.1007/0-8176-4467-9\_9}, }
Reference [KS13]
M. Kontsevich and Y. Soibelman, Wall-crossing structures in Donaldson-Thomas invariants, integrable systems and mirror symmetry, in Homological mirror symmetry and tropical geometry, Lect. Notes Unione Mat. Ital., vol. 15, Springer, Cham, 2014, pp. 197–308, DOI 10.1007/978-3-319-06514-4_6. MR3330788,
Show rawAMSref \bib{KS13}{article}{ label={KS13}, author={Kontsevich, Maxim}, author={Soibelman, Yan}, title={Wall-crossing structures in Donaldson-Thomas invariants, integrable systems and mirror symmetry}, conference={ title={in Homological mirror symmetry and tropical geometry}, }, book={ series={Lect. Notes Unione Mat. Ital.}, volume={15}, publisher={Springer, Cham}, }, date={2014}, pages={197--308}, review={\MR {3330788}}, doi={10.1007/978-3-319-06514-4\_6}, }
[LS13]
K. Lee and R. Schiffler, Positivity for cluster algebras, Ann. of Math. (2) 182 (2015), no. 1, 73–125, DOI 10.4007/annals.2015.182.1.2. MR3374957,
Show rawAMSref \bib{LS13}{article}{ label={LS13}, author={Lee, Kyungyong}, author={Schiffler, Ralf}, title={Positivity for cluster algebras}, journal={Ann. of Math. (2)}, volume={182}, date={2015}, number={1}, pages={73--125}, issn={0003-486X}, review={\MR {3374957}}, doi={10.4007/annals.2015.182.1.2}, }
Reference [LLZ13]
K. Lee, L. Li, and A. Zelevinsky, Greedy elements in rank 2 cluster algebras, Selecta Math. (N.S.) 20 (2014), no. 1, 57–82, DOI 10.1007/s00029-012-0115-1. MR3147413,
Show rawAMSref \bib{LLZ13}{article}{ label={LLZ13}, author={Lee, Kyungyong}, author={Li, Li}, author={Zelevinsky, Andrei}, title={Greedy elements in rank 2 cluster algebras}, journal={Selecta Math. (N.S.)}, volume={20}, date={2014}, number={1}, pages={57--82}, issn={1022-1824}, review={\MR {3147413}}, doi={10.1007/s00029-012-0115-1}, }
Reference [M14]
T. Mandel, Tropical theta functions and cluster varieties, Ph.D. thesis, UT Austin, 2014.
Reference [Ma15]
T. Magee, Fock–Goncharov conjecture and polyhedral cones for and base affine space , preprint, 2015.
Reference [Ma17]
T. Magee, GHK mirror symmetry, the Knutson-Tao hive cone, and Littlewood–Richardson coefficients, preprint, 2017.
Reference [M13]
J. P. Matherne and G. Muller, Computing upper cluster algebras, Int. Math. Res. Not. IMRN 11 (2015), 3121–3149. MR3373046,
Show rawAMSref \bib{M13}{article}{ label={M13}, author={Matherne, Jacob P.}, author={Muller, Greg}, title={Computing upper cluster algebras}, journal={Int. Math. Res. Not. IMRN}, date={2015}, number={11}, pages={3121--3149}, issn={1073-7928}, review={\MR {3373046}}, }
Reference [Ma89]
H. Matsumura, Commutative ring theory, 2nd ed., Cambridge Studies in Advanced Mathematics, vol. 8, Cambridge University Press, Cambridge, 1989. Translated from the Japanese by M. Reid. MR1011461,
Show rawAMSref \bib{Ma89}{book}{ label={Ma89}, author={Matsumura, Hideyuki}, title={Commutative ring theory}, series={Cambridge Studies in Advanced Mathematics}, volume={8}, edition={2}, note={Translated from the Japanese by M. Reid}, publisher={Cambridge University Press, Cambridge}, date={1989}, pages={xiv+320}, isbn={0-521-36764-6}, review={\MR {1011461}}, }
Reference [Mu15]
G. Muller, The existence of a maximal green sequence is not invariant under quiver mutation, Electron. J. Combin. 23 (2016), no. 2, Paper 2.47, 23. MR3512669,
Show rawAMSref \bib{Muller}{article}{ label={Mu15}, author={Muller, Greg}, title={The existence of a maximal green sequence is not invariant under quiver mutation}, journal={Electron. J. Combin.}, volume={23}, date={2016}, number={2}, pages={Paper 2.47, 23}, issn={1077-8926}, review={\MR {3512669}}, }
Reference [NZ]
T. Nakanishi and A. Zelevinsky, On tropical dualities in cluster algebras, Algebraic groups and quantum groups, Contemp. Math., vol. 565, Amer. Math. Soc., Providence, RI, 2012, pp. 217–226, DOI 10.1090/conm/565/11159. MR2932428,
Show rawAMSref \bib{NZ}{article}{ label={NZ}, author={Nakanishi, Tomoki}, author={Zelevinsky, Andrei}, title={On tropical dualities in cluster algebras}, conference={ title={Algebraic groups and quantum groups}, }, book={ series={Contemp. Math.}, volume={565}, publisher={Amer. Math. Soc., Providence, RI}, }, date={2012}, pages={217--226}, review={\MR {2932428}}, doi={10.1090/conm/565/11159}, }
Reference [R10]
M. Reineke, Poisson automorphisms and quiver moduli, J. Inst. Math. Jussieu 9 (2010), no. 3, 653–667, DOI 10.1017/S1474748009000176. MR2650811,
Show rawAMSref \bib{R10}{article}{ label={R10}, author={Reineke, Markus}, title={Poisson automorphisms and quiver moduli}, journal={J. Inst. Math. Jussieu}, volume={9}, date={2010}, number={3}, pages={653--667}, issn={1474-7480}, review={\MR {2650811}}, doi={10.1017/S1474748009000176}, }
Reference [R12]
M. Reineke, Cohomology of quiver moduli, functional equations, and integrality of Donaldson-Thomas type invariants, Compos. Math. 147 (2011), no. 3, 943–964, DOI 10.1112/S0010437X1000521X. MR2801406,
Show rawAMSref \bib{R12}{article}{ label={R12}, author={Reineke, Markus}, title={Cohomology of quiver moduli, functional equations, and integrality of Donaldson-Thomas type invariants}, journal={Compos. Math.}, volume={147}, date={2011}, number={3}, pages={943--964}, issn={0010-437X}, review={\MR {2801406}}, doi={10.1112/S0010437X1000521X}, }
Reference [R14]
M. Reineke, Personal communcation, 2014.
Reference [S92]
A. Schofield, General representations of quivers, Proc. London Math. Soc. (3) 65 (1992), no. 1, 46–64, DOI 10.1112/plms/s3-65.1.46. MR1162487,
Show rawAMSref \bib{S92}{article}{ label={S92}, author={Schofield, Aidan}, title={General representations of quivers}, journal={Proc. London Math. Soc. (3)}, volume={65}, date={1992}, number={1}, pages={46--64}, issn={0024-6115}, review={\MR {1162487}}, doi={10.1112/plms/s3-65.1.46}, }

Article Information

MSC 2010
Primary: 13F60 (Cluster algebras)
Secondary: 14J33 (Mirror symmetry)
Author Information
Mark Gross
DPMMS, Centre for Mathematical Sciences, Wilberforce Road, Cambridge, CB3 0WB, United Kingdom
mgross@dpmms.cam.ac.uk
MathSciNet
Paul Hacking
Department of Mathematics and Statistics, Lederle Graduate Research Tower, University of Massachusetts, Amherst, Massachusetts 01003-9305
hacking@math.umass.edu
MathSciNet
Sean Keel
Department of Mathematics, 1 University Station C1200, Austin, Texas 78712-0257
keel@math.utexas.edu
MathSciNet
Maxim Kontsevich
IHÉS, Le Bois-Marie 35, route de Chartres, 91440 Bures-sur-Yvette, France
maxim@ihes.fr
MathSciNet
Additional Notes

The first author was partially supported by NSF grant DMS-1262531 and a Royal Society Wolfson Research Merit Award, the second by NSF grants DMS-1201439 and DMS-1601065, and the third by NSF grant DMS-0854747. Some of the research was conducted when the first and third authors visited the fourth at I.H.E.S. during the summers of 2012 and 2013.

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/890
  • MathSciNet Review: 3758151
  • Show rawAMSref \bib{3758151}{article}{ author={Gross, Mark}, author={Hacking, Paul}, author={Keel, Sean}, author={Kontsevich, Maxim}, title={Canonical bases for cluster algebras}, journal={J. Amer. Math. Soc.}, volume={31}, number={2}, date={2018-04}, pages={497-608}, issn={0894-0347}, review={3758151}, doi={10.1090/jams/890}, }

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.