The positive Dressian equals the positive tropical Grassmannian

By David Speyer and Lauren K. Williams

Abstract

The Dressian and the tropical Grassmannian parameterize abstract and realizable tropical linear spaces; but in general, the Dressian is much larger than the tropical Grassmannian. There are natural positive notions of both of these spaces – the positive Dressian, and the positive tropical Grassmannian (which we introduced roughly fifteen years ago in [J. Algebraic Combin. 22 (2005), pp. 189–210]) – so it is natural to ask how these two positive spaces compare. In this paper we show that the positive Dressian equals the positive tropical Grassmannian. Using the connection between the positive Dressian and regular positroidal subdivisions of the hypersimplex, we use our result to give a new “tropical” proof of da Silva’s 1987 conjecture (first proved in 2017 by Ardila-Rincón-Williams) that all positively oriented matroids are realizable. We also show that the finest regular positroidal subdivisions of the hypersimplex consist of series-parallel matroid polytopes, and achieve equality in Speyer’s -vector theorem. Finally we give an example of a positroidal subdivision of the hypersimplex which is not regular, and make a connection to the theory of tropical hyperplane arrangements.

1. Introduction

The tropical Grassmannian, first studied in Reference HKT06Reference KT06Reference SS04, is the space of realizable tropical linear spaces, obtained by applying the valuation map to Puisseux-series valued elements of the usual Grassmannian. Meanwhile the Dressian is the space of tropical Plücker vectors , first studied by Andreas Dress, who called them valuated matroids. Thinking of each tropical Plücker vector as a height function on the vertices of the hypersimplex , one can show that the Dressian parameterizes regular matroid subdivisions of the hypersimplex Reference Kap93Reference Spe08, which in turn are dual to the abstract tropical linear spaces of the first author Reference Spe08.

There are positive notions of both of the above spaces. The positive tropical Grassmannian, introduced by the authors in Reference SW05, is the space of realizable positive tropical linear spaces, obtained by applying the valuation map to Puisseux-series valued elements of the totally positive Grassmannian Reference PosReference Lus94. The positive Dressian is the space of positive tropical Plücker vectors, and it was recently shown to parameterize the regular positroidal subdivisions of the hypersimplex Reference LPW20Reference AHLS20.⁠Footnote1

1

Although this result did not appear in the literature until recently, it was anticipated by various people including the first author, Nick Early Reference Ear19a, Felipe Rincón, Jorge Olarte.

In general, the Dressian is much larger than the tropical Grassmannian – for example, the dimension of the Dressian grows quadratically is , while the dimension of the tropical Grassmannian is linear in Reference HJJS08. However, the situation for their positive parts is different. The first main result of this paper is the following, see Theorem 3.9.

Theorem.

The positive tropical Grassmannian equals the positive Dressian .⁠Footnote2

2

Our result was announced in Reference LPW20, Theorem 9.6, and subsequently appeared in the independent work Reference AHLS20.

We give several interesting applications of Theorem 3.9. The first application is a new proof of the following 1987 conjecture of da Silva, which was proved in 2017 by Ardila, Rincón and the second author Reference ARW17, using the combinatorics of positroid polytopes.

Theorem (Reference ARW17).

Every positively oriented matroid is realizable.

Reformulating this statement in the language of Postnikov’s 2006 preprint Reference Pos, da Silva’s conjecture says that every positively oriented matroid is a positroid. We give a new proof of this statement, using Theorem 3.9, which we think of as a “tropical version” of da Silva’s conjecture. Interestingly, although the definitions of positively oriented matroid and positroid don’t involve tropical geometry at all, there does not seem to be an easy way to remove the tropical geometry from our proof without making it significantly longer.

There are two natural fan structures on the Dressian: the Plücker fan, and the secondary fan, which were shown in Reference OPS19 to coincide. Our second application of Theorem 3.9 is a description of the maximal cones in the positive Dressian, or equivalently, the finest regular positroidal subdivisions of the hypersimplex. The following result appears as Theorem 6.6.

Theorem.

Let be a positive tropical Plücker vector, and consider the corresponding regular positroidal subdivision . The following statements are equivalent:

(1)

is a finest subdivision.

(2)

Every facet of is the matroid polytope of a series-parallel matroid.

(3)

Every octahedron in is subdivided.

It was shown by the first author in Reference Spe09 that if is a tropical Plücker vector corresponding to a realizable tropical linear space, has at most interior faces of dimension , with equality if and only if all facets of correspond to series-parallel matroids. We refer to this result as the -vector theorem. Combining this result with Theorem 6.6 gives the following elegant result (see Corollary 6.7):

Corollary.

Every finest positroidal subdivision of achieves equality in the -vector theorem. In particular, such a positroidal subdivision has precisely facets (top-dimensional polytopes).

Most of our paper concerns the regular positroidal subdivisions of , which are precisely those induced by positive tropical Plücker vectors. However, it is also natural to consider the set of all positroidal subdivisions of , whether or not they are regular. In light of the various nice realizability results for positroids, one might hope that all positroidal subdivisions of are regular. However, this is not the case. In Section 7, we construct a nonregular positroidal subdivision of , based off a standard example of a nonregular mixed subdivision of . We also make a connection to the theory of tropical hyperplane arrangements and tropical oriented matroids Reference AD09Reference Hor16.

It is interesting to note that the positive tropical Grassmannian and the positive Dressian have recently appeared in the study of scattering amplitudes in SYM Reference DFGK19Reference AHHLT19Reference HP19Reference Ear19bReference LPW20Reference AHLS20, and in certain scalar theories Reference CEGM19Reference BC19. In particular, the second author together with Lukowski and Parisi Reference LPW20 gave striking evidence that the positive tropical Grassmannian controls the regular positroidal subdivisions of the amplituhedron , which was introduced by Arkani-Hamed and Trnka Reference AHT14 to study scattering amplitudes in SYM.

The structure of this paper is as follows. In Section 2 we review the notion of the positive Grassmannian and its cell decomposition, as well as matroid and positroid polytopes. In Section 3, after introducing the notions of the (positive) tropical Grassmannian and (positive) Dressian, we show that the positive tropical Grassmannian equals the positive Dressian. We review the connection between the positive tropical Grassmannian and positroidal subdivisions in Section 4, then give a new proof in Section 5 that every positively oriented matroid is realizable. We give several characterizations of finest positroidal subdivisions of the hypersimplex in Section 6, and show that such subdivisions achieve equality in the -vector theorem. Then in Section 7, we construct a nonregular positroidal subdivision of , and make a connection to the theory of tropical hyperplane arrangements and tropical oriented matroids Reference AD09Reference Hor16. We end our paper with an appendix (Section 8), which reviews some of Postnikov’s technology Reference Pos for studying positroids.

2. The positive Grassmannian and positroid polytopes

Definition 2.1.

The (real) Grassmannian (for ) is the space of all -dimensional subspaces of . An element of can be viewed as a matrix of rank modulo invertible row operations, whose rows give a basis for the -dimensional subspace.

Let denote , and denote the set of all -element subsets of . Given represented by a matrix , for we let be the minor of using the columns . The do not depend on our choice of matrix (up to simultaneous rescaling by a nonzero constant), and are called the Plücker coordinates of .

2.1. The positive Grassmannian and its cells

Definition 2.2 (Reference Pos, Section 3).

We say that is totally nonnegative (respectively, totally positive) if (resp. ) for all . The set of all totally nonnegative is the totally nonnegative Grassmannian and the set of all totally positive is the totally positive Grassmannian . For , let be the set of with the prescribed collection of Plücker coordinates strictly positive (i.e. for all ), and the remaining Plücker coordinates equal to zero (i.e. for all ). If , we call a positroid and its positroid cell.

Each positroid cell is indeed a topological cell Reference Pos, Theorem 6.5, and moreover, the positroid cells of glue together to form a CW complex Reference PSW09.

As shown in Reference Pos, the cells of are in bijection with various combinatorial objects, including decorated permutations on with anti-excedances, -diagrams of type , and equivalence classes of reduced plabic graphs of type . In Section 8 we review these objects and give bijections between them. This gives a canonical way to label each positroid by a decorated permutation, a -diagram, and an equivalence class of plabic graphs; we will correspondingly refer to positroid cells as , , etc.

2.2. Matroid and positroid polytopes

In what follows, we set , where is the standard basis of .

Definition 2.3.

Given a matroid , the (basis) matroid polytope of is the convex hull of the indicator vectors of the bases of :

The dimension of a matroid polytope is determined by the number of connected components of the matroid. Recall that a matroid which cannot be written as the direct sum of two nonempty matroids is called connected.

Proposition 2.4 (Reference Oxl11).

Let be a matroid on . For two elements , we set whenever there are bases of such that . The relation is an equivalence relation, and the equivalence classes are precisely the connected components of .

Proposition 2.5 (Reference BGW03).

For any matroid, the dimension of its matroid polytope is , where is the number of connected components of .

Recall that any full rank matrix gives rise to a matroid , where . Positroids are the matroids associated to matrices with maximal minors all nonnegative. We call the matroid polytope associated to a positroid a positroid polytope.

3. The positive tropical Grassmannian equals the positive Dressian

In this section we review the notions of the tropical Grassmannian, the Dressian, the positive tropical Grassmannian, and the positive Dressian. The main theorem of this section is Theorem 3.9, which says that the positive tropical Grassmannian equals the positive Dressian.

Definition 3.1.

Given , we let denote . Let . For a nonzero polynomial, we denote by the set of all points such that, if we form the collection of numbers for ranging over , then the minimum of this collection is not unique. We say that is the tropical hypersurface associated to .

In our examples, we always consider polynomials with real coefficients. We also have a positive version of Definition 3.1.

Definition 3.2.

Let , and let be a nonzero polynomial with real coefficients which we write as , where all of the coefficients are nonnegative real numbers. We denote by the set of all points such that, if we form the collection of numbers for ranging over , then the minimum of this collection is not unique and furthermore is achieved for some and some . We say that is the positive part of .

The Grassmannian is a projective variety which can be embedded in projective space , and is cut out by the Plücker ideal, that is, the ideal of relations satisfied by the Plücker coordinates of a generic matrix. These relations include the three-term Plücker relations, defined below.

Definition 3.3.

Let and choose a subset which is disjoint from . Then is a three-term Plücker relation for the Grassmannian . Here denotes , etc.

Definition 3.4.

Given as in Definition 3.3, we say that the tropical three-term Plücker relation holds if

or

or

.

And we say that the positive tropical three-term Plücker relation holds if either of the first two conditions above holds.

Definition 3.5.

The tropical Grassmannian is the intersection of the tropical hypersurfaces , where ranges over all elements of the Plücker ideal. The Dressian is the intersection of the tropical hypersurfaces , where ranges over all three-term Plücker relations.

The tropical Grassmannian , first studied in Reference SS04Reference HKT06Reference KT06, parameterizes tropicalizations of ordinary linear spaces, defined over the field of generalized Puisseux series in one variable , with real exponents. More formally, recall that there is a valuation , given by if , where the lowest order term is assumed to have non-zero coefficient . Then lies in the tropical Grassmannian if and only if there is an element whose Plücker coordinates have valuations given by (see Reference Pay09Reference Pay12 for a proof). We will call elements of realizable tropical linear spaces. The tropical Grassmannian is a proper subset of the Dressian,⁠Footnote3 which parameterizes what one might call abstract tropical linear spaces. Moreover, the Dressian has a natural fan structure, whose cones correspond to the regular matroidal subdivisions of the hypersimplex Reference Kap93, Reference Spe08, Proposition 2.2, see Theorem 4.2. Note that the Dressian is the subset of where the tropical three-term Plücker relations hold.

3

Also called the tropical pre-Grassmannian in Reference SS04 and named in Reference HJJS08 for Andreas Dress’ work on valuated matroids.

Definition 3.6.

The positive tropical Grassmannian is the intersection of the positive tropical hypersurfaces , where ranges over all elements of the Plücker ideal. The positive Dressian is the intersection of the positive tropical hypersurfaces , where ranges over all three-term Plücker relations.

The positive tropical Grassmannian was introduced by the authors fifteen years ago in Reference SW05, and was shown to parameterize tropicalizations of ordinary linear spaces that lie in the totally positive Grassmannian (defined over the field of Puiseux series). The positive tropical Grassmannian lies inside the positive Dressian, which controls the regular positroidal subdivisions of the hypersimplex Reference LPW20, see Theorem 4.3. Note that the positive Dressian is the subset of where the positive tropical three-term Plücker relations hold.

Definition 3.7.

We say that a point is a (finite) tropical Plücker vector if it lies in the Dressian , i.e. for every three-term Plücker relation, it lies in the associated tropical hypersurface. And we say that is a positive tropical Plücker vector, if it lies in the positive Dressian , i.e. for every three-term Plücker relation, it lies in the positive part of the associated tropical hypersurface.

Example 3.8.

For , there is only one Plücker relation, . We have that is the set of points , , , , , such that

or

or

.

And is the set of points such that

or

In general, the Dressian is much larger than the tropical Grassmannian – for example, the dimension of the Dressian grows quadratically is , while the dimension of the tropical Grassmannian is linear in Reference HJJS08. However, the situation for their positive parts is different. The main result of this section is the following.

Theorem 3.9.

The positive tropical Grassmannian equals the positive Dressian .

Theorem 3.9 was recently announced in Reference LPW20. It subsequently appeared in independent work of Reference AHLS20.

Before proving Theorem 3.9, we review some results from Reference SW05 which allow one to compute positive tropical varieties.

Remark 3.10.

In Section 8 we describe many parametrizations of cells of , which were given by Postnikov using plabic graphs. Reference SW05, Proposition 2.5 says that if one has a subtraction-free rational map which surjects onto the positive part of a variety (for example a cluster chart), then the tropicalization of this map surjects onto the positive tropical part of the variety. Therefore we can tropicalize each parameterization from Theorem 8.8 – to obtain a parameterization of a positive tropical positroid variety (in particular, ). More specifically, we tropicalize by replacing the positive parameters (with ) with real parameters (with ) – and replacing products with sums and sums with minimums in the expressions for flow polynomials. Then Reference SW05, Proposition 2.5 say that this tropicalized map gives a parameterization of .

For the proof of Theorem 3.9 it is convenient to use one particular plabic graph (corresponding to the directed graph from Reference SW05, Section 3), see Figure 1.

Applying Theorem 8.8 to the graph from Figure 1, we have the following result.

Theorem 3.11.

Label the faces of by indices and let denote the collection of indices. Define the weight of a path in to be the product of parameters where ranges over all face labels to the left of a path. Define the weight of a flow (i.e. a collection of nonintersecting paths) to be the product of the weights of its paths. Let where ranges over all flows from to . Then the map sending to the collection of flow polynomials is a homeomorphism from to the totally positive Grassmannian (realized in its Plücker embedding).

In the case of the graph , we obtain the following parameterization of .

Theorem 3.12.

Label the faces of by indices as before. Define the weight of a path in to be the sum of parameters where ranges over all face labels to the left of a path. Define the weight of a flow (i.e. a collection of nonintersecting paths) to be sum of the weights of its paths. Let where ranges over all flows from to . Then the map sending to the collection of tropical flow polynomials is a bijection from to the tropical positive Grassmannian (realized in its Plücker embedding).

In the case of , we can easily invert the maps and . This was done in Reference SW05; we review the construction here. First, given and labeling horizontal and vertical wires of (i.e. and ), let

If does not correspond to a region of , set .

Definition 3.13.

Let . Then for and labeling horizontal and vertical wires of (i.e. and ), we define

We likewise define the tropical version. Let .

Definition 3.13 gives a way to label each face of by a (tropical) Laurent monomial in (tropical) Plücker coordinates. This is shown in Figure 2.

Proposition 3.14.

The maps and are inverses.

Proposition 3.15 (Reference SW05, Corollary 3.5 and its proof).

The maps and are inverses.

Lemma 3.16.

The collection of Plücker coordinates form a cluster for the cluster algebra structure Reference Sco06 on (the affine cone over the) Grassmannian . We call this the corectangles cluster. In particular, this collection of Plücker coordinates is algebraically independent, and all other Plücker coordinates can be written as Laurent polynomials with positive coefficients in the Plücker coordinates from the collection.

Proof.

Note that for each and as above, is a -element subset of . Moreover, if we identify Young diagrams contained in a rectangle with the labels of the vertical steps in the length- lattice path taking unit steps south and west from to , then the elements precisely correspond to the Young diagrams whose complementary Young diagram is a rectangle. It is not hard to see that the collection is a maximal weakly separated set collection Reference OPS15, and hence form a cluster for the cluster algebra structure Reference Sco06.

Example 3.17.

Figure 2 depicts the map . Since and are inverses, this example shows how to express each of the variables (as shown in Figure 1) in terms of the tropical Plücker coordinates . Note moreover that if we choose a normalization in tropical projective space (e.g. where ), then we can solve for the tropical Plücker coordinates in in terms of the ’s. For example, comparing Figure 1 and Figure 2, we see that if , then , , so , etc. In this example we see that from the collection together with the normalization , we can uniquely determine the Plücker coordinates . As in Lemma 3.16, this collection of Plücker coordinates is a cluster for the cluster algebras structure on the Grassmannian.

It is easy to generalize Example 3.17, obtaining the following result.

Lemma 3.18.

The map sending (with the convention that ) to is an injective map from to .

Proof of Theorem 3.9.

To prove Theorem 3.9, we must show that every point in the positive Dressian also lies in the positive tropical Grassmannian. We will consider any tropical Plücker vector , with the normalization , and compute . This will give an (a priori new) realizable tropical Plücker vector in . We must show that .

Recall that the map depends only on the tropical Plücker coordinates in , mapping them to . Moreover from Lemma 3.18, is an injective map from to . Therefore, since and are inverses, we have that for all with and . But now from Lemma 3.16, the collection is a cluster for the cluster structure on . And by Reference OS17, all Plücker clusters can be obtained from each via three-term Plücker relations. Since every Plücker coordinate lies in a Plücker cluster Reference OPS15, the three-term Plücker relations alone (which we know are satisfied since ) determine all the other values and for , so we must have for all . Therefore and we are done.

Remark 3.19.

One may generalize Theorem 3.9 and its proof to any positroid cell, using the -network associated to a positroid cell and the inverse map from Reference Tal11.

4. The positive tropical Grassmannian and positroidal subdivisions

Recall that denotes the -hypersimplex, defined as the convex hull of the points where runs over . Consider a real-valued function on the vertices of . We define a polyhedral subdivision of as follows: consider the points and take their convex hull. Take the lower faces (those whose outwards normal vector have last component negative) and project them back down to ; this gives us the subdivision . We will omit the subscript when it is clear from context. A subdivision obtained in this manner is called regular.

Remark 4.1.

A lower face of the regular subdivision defined above is determined by some vector whose dot product with the vertices of the is maximized. So if is the matroid polytope of a matroid with bases , this is equivalent to saying that for any two bases and .

Given a subpolytope of , we say that is matroidal if the vertices of , considered as elements of , are the bases of a matroid , i.e. .

The following result is originally due to Kapranov Reference Kap93; it was also proved in Reference Spe08, Proposition 2.2.

Theorem 4.2.

The following are equivalent.

The collection is a tropical Plücker vector.

The one-skeleta of and are the same.

Every face of is matroidal.

Given a subpolytope of , we say that is positroidal if the vertices of , considered as elements of , are the bases of a positroid , i.e. . The positroidal version of Theorem 4.2 was recently proved in Reference LPW20, and independently in Reference AHLS20.

Theorem 4.3.

The following are equivalent.

The collection is a positive tropical Plücker vector.

Every face of is positroidal.

It follows from Theorem 4.3 that the regular subdivisions of consisting of positroid polytopes are precisely those of the form , where is a positive tropical Plücker vector.

5. A new proof that positively oriented matroids are realizable

In 1987, da Silva Reference dS87 conjectured that every positively oriented matroid is realizable. Reformulating this statement in the language of Postnikov’s 2006 preprint Reference Pos, her conjecture says that every positively oriented matroid is a positroid. In 2017, da Silva’s conjecture was proved by Ardila, Rincón and the second author Reference ARW17, using the combinatorics of positroid polytopes. In this section we will give a new proof of the conjecture, using our Theorem 3.9, which we think of as a “tropical version” of da Silva’s conjecture.

Recall that an oriented matroid of rank on can be specified by its chirotope, which is a function from to obeying certain axioms Reference BLVS+99. If is a full rank real matrix, the function taking to the sign of the minor using columns is a chirotope, and the realizable oriented matroids are precisely the chirotopes occurring in this way. Thus, if represents a point of the totally nonnegative Grassmannian, then gives a chirotope with for .

We define a positively oriented matroid to be a chirotope such that for . Since every positroid gives rise to a positively oriented matroid, to prove da Silva’s conjecture, we need to verify that every positively oriented matroid comes from a positroid, or in other words, is realizable.

Theorem 5.1 (Reference ARW17, Theorem 5.1).

Let be a positively oriented matroid of rank on the ground set . Then is realizable.

Before proving Theorem 5.1, we need the following lemma, which was implicit in Reference Spe08, Section 4.

Lemma 5.2.

Suppose that lies in the tropical Grassmannian . Then every face of the matroidal subdivision of corresponds to a realizable matroid.

Proof.

Let be a face of , and let denote the bases of the matroid . Adding an affine linear function to , we may assume that is for ; convexity then implies that for .

Since lies in the tropical Grassmannian, we can choose a -valued matrix whose Plücker coordinates have valuations given by (see the discussion following Definition 3.5). But now if we set , then the matrix has Plücker coordinates which are nonzero for and zero for . Therefore is a realizable matroid.

Proof of Theorem 5.1.

We first use the matroid to construct a point of the Dressian, following the method of Reference Spe08, Proposition 4.4 Namely, let be the rank function of the matroid , and for , set . Then Reference Spe08, Proposition 4.4 implies that is a point of the Dressian, and that the matroid polytope is a face of the subdivision .

Using the fact that is positively oriented, we will show that is in fact a point of the positive Dressian. Indeed, consider any -element subset of and any in . We need to show that

or equivalently, that

Let be the matroid on the ground set . For , , we have . Thus, we need to show that

Now we claim that , being a minor of a positively oriented matroid, is itself a positively oriented matroid. It is easy to verify that the dual of a positively oriented matroid is again a positively oriented matroid, and moreover, Reference ARW17, Lemma 4.11 showed that positively oriented matroids are closed under restriction. An analogous proof shows that positively oriented matroids are closed under contraction. This verifies the claim.

It now remains to verify Equation 5.3 for all positively oriented matroids on four elements, which is routine.

We now know that lies in the positive Dressian, so Theorem 3.9 shows that is in the positive tropical Grassmannian. But now by Lemma 5.2, this implies that every face of the matroidal subdivision of corresponds to a realizable matroid. In particular, we have with equality if and only if is a basis of , so is a face of , and we have shown that is realizable.

Interestingly, although the definitions of “positively oriented matroid” and “positroid” don’t involve tropical geometry at all, there does not seem to be a way to remove the tropical geometry from our proof without making it significantly longer.

6. Finest positroidal subdivisions of the hypersimplex

In this section we show that finest positroidal subdivisions of the hypersimplex achieve equality in the first author’s -vector theorem.

Definition 6.1.

A matroid is called series-parallel if it can be obtained by repeated series-parallel extensions from the matroid corresponding to a generic point of .

See Reference Whi86, Section 6.4 for background on series-parallel matroids.

Theorem 6.2 (Reference Spe09).

Let be a tropical Plücker vector arising as for some . Then has at most interior faces of dimension , with equality if and only if all facets of correspond to series-parallel matroids.

In particular, the number of facets of – that is, the number of matroid polytopes of dimension in – is at most .

The following result can be found in Reference Oxl11, Corollary 11.2.15.

Theorem 6.3.

A connected matroid is series-parallel if and only if it has no minor which is the uniform matroid or the graphical matroid associated to the complete graph .

The graphical matroid is not a positroid, and all minors of positroids are positroids Reference ARW16, so we have the following corollary.

Corollary 6.4.

A connected positroid is series-parallel if and only if it has no uniform matroid as a minor.

If is a matroid on the ground set , with matroid polytope , and and are disjoint subsets of , then the the polytope is . So we can phrase Corollary 6.4 as

Corollary 6.5.

Let be a connected positroid. Then is series-parallel if and only if its matroid polytope does not contain any face which is an (unsubdivided) octahedron.

It follows from Proposition 2.5 that in a matroidal subdivision, all facets correspond to connected matroids.

Theorem 6.6.

Let be a positive tropical Plücker vector. For the positroidal subdivision of , the following are equivalent:

(1)

is a finest subdivision.

(2)

Every facet of is the matroid polytope of a series-parallel matroid.

(3)

Every octahedron in is subdivided.

Proof.

Suppose that (3) holds. Let be a facet of this subdivision. Since , the matroid is connected, and by hypothesis is a positroid. Hypothesis (3) says that does not contain any octahedron, so Corollary 6.5 says that is series-parallel. We have shown .

Now suppose that (2) holds. If every facet is series-parallel, then by Theorem 6.2, we get equality in the -vector theorem, and in particular get equality in the term. So we have the maximal number of possible facets, so the positroidal subdivision is finest possible. This implies (1).

Now suppose that (1) holds. To show that every octahedron in is subdivided, we need to show that we never have equality in a tropical 3-term Plücker relation, in other words, we never have

for and disjoint from .

Using the fact that the positive Dressian equals the positive tropical Grassmannian (Theorem 3.9), as well as Remark 3.10, we can use flows in plabic graphs to parameterize the points in the positive Dressian, as in Theorem 8.8. We note that it follows from the technology of Reference PSW09 that a flow is uniquely determined by its weight (compare Definition 4.3 and Table 1, and note that flows are in bijection with almost perfect matchings).

Let us choose a reduced plabic graph for , i.e. a reduced plabic graph with trip permutation , and choose a perfect orientation with sources at . (The fact that we can do so follows from e.g. Proposition 8.4).

Then by Remark 3.10, we can express for some fixed real values labeling the faces of . In particular, the coordinates of can be expressed as where ranges over all flows from to , and is a sum of certain parameters .

Since we are assuming that is finest, we can assume that the parameters are generic: that these parameters are distinct real numbers, and that there are not two different subsets of parameters whose sums coincide.

Let us consider the tropical Plücker coordinate . This equals where ranges over all flows from to ; in this case, the flows are simply collections of vertex-disjoint cycles in (including the empty collection). We now explain how to reduce to the case that the flow achieving the minimum is the empty flow.

Let be the flow achieving the minimum, so is a collection of disjoint cycles. Adjust to a new perfect orientation by reversing the orientation of all edges belonging to . Then is again a perfect orientation (see Reference PSW09, Lemma 4.5) and that (preserving the values of the ) the collection of new Plücker coordinates are all adjusted by the same scalar (the weight of ), preserving the point in tropical projective space which is represented by . Now, in the orientation , the minimum flow for is the empty flow. We therefore assume, from now on, that the minimum flow for is the empty flow. With this reduction, we have .

Meanwhile is the weight of the minimal flow from to , which will be a pair of paths taking to and to (plus possibly some closed loops). is the weight of the minimal flow from to , which will be a single path from to (plus possibly closed loops). And is the weight of the minimal flow from to , which will be a single path from to (plus possibly closed loops), see Figure 3.

But now because our parameters associated to the faces are generic, the only way to get equality is if our minimal flow from to has to its left precisely the same multiset of faces that the pair of flows (which consists of the paths plus possibly some loops) does. This is only possible if and are obtained from and by “switching tails” at an intersection point of and . But then would not be vertex-disjoint and hence not part of a flow.

Combining Theorem 3.9, Theorem 6.2, and Theorem 6.6, we now have the following.

Corollary 6.7.

Every finest positroidal subdivision of achieves equality in the -vector theorem. In particular, such a positroidal subdivision has precisely facets.

7. Nonregular positroidal subdivisions

In this paper we have discussed the positive Dressian, which consists of weight functions on the vertices of the hypersimplex which induce positroidal subdivisions of ; recall that subdivisions induced by weight functions are called regular or coherent. It is also natural to consider the set of all positroidal subdivisions of , whether or not they are regular. (See Reference DLRS10 for background on regular subdivisions.) In this section, we will construct a nonregular positroidal subdivision of , and also make a connection to the theory of tropical hyperplane arrangements and tropical oriented matroids Reference AD09Reference Hor16.

Our strategy for producing the counterexample is as follows. We will start with a standard example of a nonregular rhombic tiling of a hexagon (with side lengths equal to ), and extend it to a nonregular mixed subdivision of ; this mixed subdivision gives rise to a dual arrangement of tropical pseudohyperplanes in . Moreover, the mixed subdivision corresponds, via the Cayley trick, to a polyhedral subdivision of . We then map this polyhedral subdivision to a matroidal subdivision of , and analyze the -dimensional regions of to show that it is a positroidal subdivision of . Note that Reference HJJS08, Example 4.7 used a similar strategy to encode a nonregular matroidal subdivision of . We give a careful exposition here in order to verify that our subdivision is positroidal.

7.1. The product of simplices and the hypersimplex

Let be any -element subset of and let . Let be the convex hull of all points of the form for and ; clearly this set of points is in bijection with . The polytope is isomorphic to , with vertices in bijection with . has dimension and sits inside , which has dimension . We review standard constructions for passing between polyhedral subdivisions of and matroidal subdivisions of . We will be interested in polyhedral subdivisions of all of whose vertices are vertices of , and we will take the phrase “subdivision of to include this condition.

In many references, is standardized to be . However, we will want to keep track of how these standard constructions relate to the property of a matroid being a positroid and, for this purpose, it will be important how sits inside the circularly ordered set , so we do not impose a standard choice of .

Given a matroidal subdivision of , we can intersect with and obtain a polyhedral subdivision of . If is regular, so is .

7.2. From subdivisions of to subdivisions of

Following Reference HJS14, Theorem 7 and Remark 8, as well as Reference Rin13, we will explain how to map each convex hull of vertices of to a matroid polytope inside ; this will be the matroid polytope of a principal transversal matroid.

Let . We define a polytope . We also define a bipartite graph with vertex set and an edge from to if and only if .

Associated to the graph is the principal transversal matroid (see Reference Bru87 and Reference Whi86, Chapter 7), defined as follows: is a basis of if and only if there is a matching of to in the bipartite graph . The matroid is realized by a matrix , with rows labeled by and columns labeled by where:

the values for (where and ) are algebraically independent,

if (where and ),

(where ).

Remark 7.1.

Note that the restriction of to the columns labeled by is the identity matrix.

In terms of polyhedral geometry, the matroid polytope of is the intersection of with . Summarizing, we have the following.

Lemma 7.2.

Each polytope gives rise to the matroid polytope . Abusing notation, we say that maps to .

If is a polyhedral subdivision of , then we can apply to each polytope in .

Proposition 7.3 (Reference HJS14, Theorem 7 and Remark 8 and Reference Rin13).

If we apply to each polytope in a polyhedral subdivision of , then we will obtain a matroid subdivision of . The subdivision is regular if and only if is.

We will eventually be studying triangulations of , so we will want to focus on the case that is an -dimensional simplex.

Lemma 7.4.

Let . The following are equivalent:

(1)

The polytope is an -dimensional simplex.

(2)

The graph is a tree on the vertices .

Proof.

The equivalence of (1) and (2) is simple. The polytope is an -dimensional simplex if and only if it’s the convex hull of affinely independent points. But this is equivalent to the statement that consists of edges and no subset forms a cycle. This means that is a tree on .

Remark 7.5.

The conditions from Lemma 7.4 are additionally equivalent to the condition that the matroid is series-parallel. One can prove this using e.g. Reference Spe08, Proposition 5.1.

We will want to know when the matroids in Lemma 7.4 are positroidal. One direction of Lemma 7.6 comes from Reference Mar19, Theorem 6.3.

Lemma 7.6.

Suppose that is a subset of such that is a tree. The matroid is positroidal if and only if we can embed the tree in a disk so that it is planar, and its vertices lie on the boundary of the disk in the standard circular order on .

Proof.

If can be embedded as a planar tree in a disk as above, then this graph is noncrossing, and by Reference Mar19, Theorem 6.3 the transversal matroid is a positroid.

On the other hand, if it cannot be embedded as a planar tree, then we can find and , such that and lie in , and when we put the numbers at the boundary of a disk in the standard circular order, the two chords and cross each other. Moreover since is a tree, we cannot have both and in . Without loss of generality we can assume that either or . Let us consider the first case. Then if we look at the rows labeled by and the columns labeled by in the matrix , we find that the minors and are nonzero, but the product is zero. This fails to be a positroid on because such conditions are incompatible with finding a non-negative solution to the Plücker relation . Using Remark 7.1, we can now extend this submatrix of to a submatrix of , by adding the rows and columns indexed by The second case is analogous.

7.3. From tropical pseudohyperplane arrangements to subdivisions of the product of simplices

We now explain how to go between tropical pseudohyperplane arrangements and subdivisions of . This section is based on Reference AD09, which initiated the study of tropical oriented matroids and conjectured that they are in bijection with subdivisions of the product of two simplices. Reference AD09 proved their conjecture in the case of , which is all we need here; Reference Hor16 proved their conjecture in general. Consult these sources for more detail.

Let denote tropical projective space , and let be an element of . The tropical hyperplane centered at is the set of points such that is not unique. If is any point of , we let be the set of indices at which is minimized. Figure 4 shows a tropical hyperplane in , where the horizontal and vertical coordinates are and , and each region is labelled with the set for in that region. An arrangement of labelled tropical hyperplanes is a list of tropical hyperplanes in .

A tropical pseudohyperplane is a subset of which is PL-homeomorphic to a tropical hyperplane. Note that the quantity makes sense for a tropical pseudohyperplane in and . An arrangement of labelled tropical pseudohyperplanes is a list of tropical pseudohyperplanes which intersect in “reasonable” ways, see Reference Hor16, Section 5 for details. Our main focus in this section will be on the case of tropical pseudohyperplanes in .

Consider an arrangement of tropical pseudohyperplanes , , …, in . Given a point , we define a subset of where if and only if . We can thus associate to each a polytope , as well as the matroid polytope of the transversal matroid . If we let range over the bounded regions of the tropical pseudohyperplane arrangement, we obtain the interior regions of a subdivision of . Using Reference DS04, Theorem 1 and Reference Hor16, Theorems 1.2 and 1.3, this subdivision is regular if and only if tropical pseudohyperplane arrangement can be realized by genuine tropical hyperplanes.

7.4. Our counterexample

We start with the mixed subdivision of shown in Figure 5. The subdivision of the central hexagon (with each side of length ) is a standard example of a nonregular subdivision of a hexagon into rhombi, originally found by Richter-Gebert, see Reference ER96, Figure 9. Thus, this mixed subdivision of is not regular.

Mixed subdivisions of are dual to arrangements of labeled tropical pseudohyperplanes in . The arrangement of tropical pseudohyperplanes in which is dual to the mixed subdivision from Figure 5 is shown in Figure 6. In this figure we have labeled the coordinates of by – placing the labels at the “ends” of the rays, according to which coordinate is becoming large along the ray – and labelled the tropical pseudohyperplanes by , placing the label at the trivalent point.

Also, by the “Cayley trick” Reference HRS00Reference San05, mixed subdivisions of correspond to polyhedral subdivisions of , with regular mixed subdivisions of corresponding to regular polyhedral subdivisions of . Therefore the mixed subdivision from Figure 5 corresponds to a nonregular polyhedral subdivision of .

It remains to check that this subdivision is positroidal. We need to check that each of the two-dimensional polytopes in Figure 5, or equivalently, each of the zero-dimensional cells of the tropical pseudohyperplane arrangement in Figure 6, corresponds to a positroid. Letting be one of these zero dimensional cells, we must check that is a tree in each case, which can be embedded in a disk as in Lemma 7.6.

For example, let be the crossing which is circled in Figure 6; the dual rhombus is shaded in Figure 5. We have

We draw the corresponding tree in Figure 7.

8. Appendix. Combinatorics of cells of the positive Grassmannian

In Reference Pos, Postnikov defined several families of combinatorial objects which are in bijection with cells of the positive Grassmannian, including decorated permutations, and equivalence classes of reduced plabic graphs. Here we review these objects as well as parameterizations of cells.

Definition 8.1.

A decorated permutation of is a bijection whose fixed points are each colored either black (loop) or white (coloop). We denote a black fixed point by , and a white fixed point by . An anti-excedance of the decorated permutation is an element such that either or .

For example, has a loop in position , and a coloop in position . It has three anti-excedances, in positions . We let denote the number of anti-excedances of .

Postnikov showed that the positroids for are indexed by decorated permutations of with exactly anti-excedances Reference Pos, Section 16.

Definition 8.2.

A plabic graphFootnote4 is an undirected planar graph drawn inside a disk (considered modulo homotopy) with boundary vertices on the boundary of the disk, labeled in clockwise order, as well as some internal vertices. Each boundary vertex is incident to a single edge, and each internal vertex is colored either black or white. If a boundary vertex is incident to a leaf (a vertex of degree ), we refer to that leaf as a lollipop.

4

“Plabic” stands for planar bi-colored.

Definition 8.3.

A perfect orientation of a plabic graph is a choice of orientation of each of its edges such that each black internal vertex is incident to exactly one edge directed away from ; and each white internal vertex is incident to exactly one edge directed towards . A plabic graph is called perfectly orientable if it admits a perfect orientation. Let denote the directed graph associated with a perfect orientation of . The source set of a perfect orientation is the set of which are sources of the directed graph . Similarly, if , then is a sink of .

See Figure 8 for an example.

All perfect orientations of a fixed plabic graph have source sets of the same size , where . Here the sum is over all internal vertices , for a black vertex , and for a white vertex; see Reference Pos. In this case we say that is of type .

As shown in Reference Pos, Section 11, every perfectly orientable plabic graph gives rise to a positroid as follows. (Moreover, every positroid can be realized in this way.)

Proposition 8.4.

Let be a plabic graph of type . Then we have a positroid on whose bases are precisely

where is the set of sources of .

Each positroid cell corresponds to a family of reduced plabic graphs which are related to each other by certain moves; see Reference Pos, Section 12. From a reduced plabic graph , we can read off the corresponding decorated permutation as follows.

Definition 8.5.

Let be a reduced plabic graph of type with boundary vertices . For each boundary vertex , we follow a path along the edges of starting at , turning (maximally) right at every internal black vertex, and (maximally) left at every internal white vertex. This path ends at some boundary vertex . By Reference Pos, Section 13, the fact that is reduced implies that each fixed point of is attached to a lollipop; we color each fixed point by the color of its lollipop. In this way we obtain the decorated permutation of . The decorated permutation will have precisely anti-excedances.

We now explain how to parameterize elements of positroid cells using perfect orientations of reduced plabic graphs.

We will associate a parameter to each face of , letting denote the indexing set for the faces. We require that the product of all parameters equals . A flow from to a set of boundary vertices with is a collection of paths and closed cycles in , all pairwise vertex-disjoint, such that the sources of the paths are and the destinations of the paths are .

Note that each directed path and cycle in partitions the faces of into those which are on the left and those which are on the right of . We define the weight of each such path or cycle to be the product of parameters , where ranges over all face labels to the left of the path. And we define the weight of a flow to be the product of the weights of all paths and cycles in the flow.

Fix a perfect orientation of a reduced plabic graph . Given , we define the flow polynomial

where ranges over all flows from to .

Example 8.7.

Consider the graph from Figure 8. There are two flows from to , and . There is one flow from to , and

The following result is a combination of Reference Pos, Theorem 12.7 and Reference Tal08, Theorem 1.1.

Theorem 8.8.

Let be a reduced plabic graph of type , and choose a perfect orientation with source set . Then the map sending to the collection of flow polynomials is a homemorphism from to the corresponding positroid cell (realized in its Plücker embedding).

Acknowledgments

This material is based upon work supported by the National Science Foundation under agreement No. DMS-1855135, No. DMS-1854225, No. DMS-1854316 and No. DMS-1854512. Any opinions, findings and conclusions or recommendations expressed in this material are those of the authors and do not necessarily reflect the views of the National Science Foundation.

Table of Contents

  1. Abstract
  2. 1. Introduction
    1. Theorem.
    2. Theorem (ARW17).
    3. Theorem.
    4. Corollary.
  3. 2. The positive Grassmannian and positroid polytopes
    1. Definition 2.1.
    2. 2.1. The positive Grassmannian and its cells
    3. Definition 2.2 (Pos, Section 3).
    4. 2.2. Matroid and positroid polytopes
    5. Definition 2.3.
    6. Proposition 2.4 (Oxl11).
    7. Proposition 2.5 (BGW03).
  4. 3. The positive tropical Grassmannian equals the positive Dressian
    1. Definition 3.1.
    2. Definition 3.2.
    3. Definition 3.3.
    4. Definition 3.4.
    5. Definition 3.5.
    6. Definition 3.6.
    7. Definition 3.7.
    8. Example 3.8.
    9. Theorem 3.9.
    10. Theorem 3.11.
    11. Theorem 3.12.
    12. Definition 3.13.
    13. Proposition 3.14.
    14. Proposition 3.15 (SW05, Corollary 3.5 and its proof).
    15. Lemma 3.16.
    16. Example 3.17.
    17. Lemma 3.18.
  5. 4. The positive tropical Grassmannian and positroidal subdivisions
    1. Theorem 4.2.
    2. Theorem 4.3.
  6. 5. A new proof that positively oriented matroids are realizable
    1. Theorem 5.1 (ARW17, Theorem 5.1).
    2. Lemma 5.2.
  7. 6. Finest positroidal subdivisions of the hypersimplex
    1. Definition 6.1.
    2. Theorem 6.2 (Spe09).
    3. Theorem 6.3.
    4. Corollary 6.4.
    5. Corollary 6.5.
    6. Theorem 6.6.
    7. Corollary 6.7.
  8. 7. Nonregular positroidal subdivisions
    1. 7.1. The product of simplices and the hypersimplex
    2. 7.2. From subdivisions of to subdivisions of
    3. Lemma 7.2.
    4. Proposition 7.3 (HJS14, Theorem 7 and Remark 8 and Rin13).
    5. Lemma 7.4.
    6. Lemma 7.6.
    7. 7.3. From tropical pseudohyperplane arrangements to subdivisions of the product of simplices
    8. 7.4. Our counterexample
  9. 8. Appendix. Combinatorics of cells of the positive Grassmannian
    1. Definition 8.1.
    2. Definition 8.2.
    3. Definition 8.3.
    4. Proposition 8.4.
    5. Definition 8.5.
    6. Example 8.7.
    7. Theorem 8.8.
  10. Acknowledgments

Figures

Figure 1.

for and . If is the path on the right-hand side, then and

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=.75] \node at (2.2,6.6) {$X_{16}$}; \node at (3.7,6.7) {$X_{15}$}; \node at (5.7,6.7) {$X_{14}$}; \node at (1.8,4.8) {$X_{26}$}; \node at (3.7,4.8) {$X_{25}$}; \node at (5.7,4.9) {$X_{24}$}; \node at (1.8,2.9) {$X_{36}$}; \node at (3.5,2.9) {$X_{35}$}; \node at (5.7,2.9) {$X_{34}$}; \node at (7,7.5) {\Large1}; \node at (7,6) {\Large2}; \node at (7,4) {\Large3}; \node at (4.5,1.9) {\Large4}; \node at (2.5,1.9) {\Large5}; \node at (1,1.9) {\Large6}; \begin{scope}[thick, every node/.style={sloped,allow upside down}] \draw(2.35,7.5) arc (120:160:3); \draw[-triangle 60] (1.52,6.8) -- (1.5,6.78); \draw(4.35,7.5)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (2.65,7.5); \draw(6.6,7.5)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (4.35,7.5); \draw(6.6,6)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (5,6); \draw(6.6,4)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (5,4); \draw(1,6)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (1,4.1); \draw(2.5,5.35)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (3,4.1); \draw(4.5,5.35)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (5,4.1); \draw(1,4)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (1,2.3); \draw(2.5,3.35)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (2.5,2.3); \draw(4.5,3.35)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (4.5,2.3); \draw(4.9,5.9)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (4.6,5.6); \draw(2.9,5.9)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (2.6,5.6); \draw(4.9,3.9)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (4.6,3.6); \draw(2.9,3.9)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (2.6,3.6); \draw(4.4,5.6)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (3,6); \draw(2.4,5.6)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (1,6); \draw(4.4,3.6)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (3,4); \draw(2.4,3.6)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (1,4); \draw(2.5,7.5)-- (3,6); \draw(4.5,7.5)-- (5,6); \end{scope} \filldraw[fill=white] (2.5,7.5) circle (0.15cm); \filldraw[fill=white] (4.5,7.5) circle (0.15cm); \filldraw[fill=black] (1,6) circle (0.15cm); \filldraw[fill=black] (3,6) circle (0.15cm); \filldraw[fill=black] (5,6) circle (0.15cm); \filldraw[fill=white] (2.5,5.5) circle (0.15cm); \filldraw[fill=white] (4.5,5.5) circle (0.15cm); \filldraw[fill=black] (1,4) circle (0.15cm); \filldraw[fill=black] (3,4) circle (0.15cm); \filldraw[fill=black] (5,4) circle (0.15cm); \filldraw[fill=white] (2.5,3.5) circle (0.15cm); \filldraw[fill=white] (4.5,3.5) circle (0.15cm); \draw[rounded corners, gray, line width=4mm] (16.8,6) -- (15,6) -- (14.5,5.5) -- (13,6) -- (12.5,5.5) -- (13,4) -- (12.5,3.5) -- (11,4) -- (11,2.18); \node at (12.2,6.6) {$X_{16}$}; \node at (13.7,6.7) {$X_{15}$}; \node at (15.7,6.7) {$X_{14}$}; \node at (11.8,4.8) {$X_{26}$}; \node at (13.7,4.8) {$X_{25}$}; \node at (15.7,4.9) {$X_{24}$}; \node at (11.8,2.9) {$X_{36}$}; \node at (13.5,2.9) {$X_{35}$}; \node at (15.7,2.9) {$X_{34}$}; \node at (17,7.5) {\Large1}; \node at (17,6) {\Large2}; \node at (17,4) {\Large3}; \node at (14.5,1.9) {\Large4}; \node at (12.5,1.9) {\Large5}; \node at (11,1.9) {\Large6}; \begin{scope}[thick, every node/.style={sloped,allow upside down}] \draw(12.35,7.5) arc (120:160:3); \draw[-triangle 60] (11.52,6.8)-- (11.5,6.78); \draw(14.35,7.5)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (12.65,7.5); \draw(16.6,7.5)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (14.35,7.5); \draw(16.6,6)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (15,6); \draw(16.6,4)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (15,4); \draw(11,6)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (11,4.1); \draw(12.5,5.35)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (13,4.1); \draw(14.5,5.35)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (15,4.1); \draw(11,4)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (11,2.3); \draw(12.5,3.35)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (12.5,2.3); \draw(14.5,3.35)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (14.5,2.3); \draw(14.9,5.9)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (14.6,5.6); \draw(12.9,5.9)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (12.6,5.6); \draw(14.9,3.9)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (14.6,3.6); \draw(12.9,3.9)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (12.6,3.6); \draw(14.4,5.6)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (13,6); \draw(12.4,5.6)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (11,6); \draw(14.4,3.6)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (13,4); \draw(12.4,3.6)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (11,4); \draw(12.5,7.5)-- (13,6); \draw(14.5,7.5)-- (15,6); \end{scope} \filldraw[fill=white] (12.5,7.5) circle (0.15cm); \filldraw[fill=white] (14.5,7.5) circle (0.15cm); \filldraw[fill=black] (11,6) circle (0.15cm); \filldraw[fill=black] (13,6) circle (0.15cm); \filldraw[fill=black] (15,6) circle (0.15cm); \filldraw[fill=white] (12.5,5.5) circle (0.15cm); \filldraw[fill=white] (14.5,5.5) circle (0.15cm); \filldraw[fill=black] (11,4) circle (0.15cm); \filldraw[fill=black] (13,4) circle (0.15cm); \filldraw[fill=black] (15,4) circle (0.15cm); \filldraw[fill=white] (12.5,3.5) circle (0.15cm); \filldraw[fill=white] (14.5,3.5) circle (0.15cm); \end{tikzpicture}
Figure 2.

Inverting the map.

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale = 0.8] \node at (2.8,5.1) {\LARGE$\frac{p_{\scaleto{456}{5pt}}p_{\scaleto{134}{5pt}}p_{\scaleto{125}{5pt}}}{p_{\scaleto{345}{5pt}}p_{\scaleto{156}{5pt}}p_{\scaleto{124}{5pt}}}$}; \node at (5.8,5.5) {\LARGE$\frac{p_{\scaleto{345}{5pt}}p_{\scaleto{124}{5pt}}}{p_{\scaleto{234}{5pt}}p_{\scaleto{145}{5pt}}}$}; \node at (2.5,2.2) {\LARGE$\frac{p_{\scaleto{156}{5pt}}p_{\scaleto{124}{5pt}}}{p_{\scaleto{126}{5pt}}p_{\scaleto{145}{5pt}}}$}; \node at (5.5,2.2) {\LARGE$\frac{p_{\scaleto{145}{5pt}}p_{\scaleto{123}{5pt}}}{p_{\scaleto{125}{5pt}}p_{\scaleto{134}{5pt}}}$}; \node at (8.25,5.5) {\LARGE$\frac{p_{\scaleto{234}{5pt}}}{p_{\scaleto{134}{5pt}}}$}; \node at (8.2,2.3) {\LARGE$\frac{p_{\scaleto{134}{5pt}}}{p_{\scaleto{124}{5pt}}}$}; \node at (2.3,-0.5) {\LARGE$\frac{p_{\scaleto{126}{5pt}}}{p_{\scaleto{125}{5pt}}}$}; \node at (5.25,-0.5) {\LARGE$\frac{p_{\scaleto{125}{5pt}}}{p_{\scaleto{124}{5pt}}}$}; \node at (8.2,-0.5) {\LARGE$\frac{p_{\scaleto{124}{5pt}}}{p_{\scaleto{123}{5pt}}}$}; \node at (10,7) {\huge1}; \node at (10,4) {\huge2}; \node at (10,1) {\huge3}; \node at (6.7,-2) {\huge4}; \node at (3.7,-2) {\huge5}; \node at (1,-2) {\huge6}; \begin{scope}[thick, every node/.style={sloped,allow upside down}] \draw(4,7) arc (90:180:3); \draw(7,7)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (4,7); \draw(9.6,7)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (7,7); \draw(9.6,4)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (7.5,4); \draw(9.6,1)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (7.5,1); \draw(4,7)-- (4.5,4); \draw(7,7)-- (7.5,4); \draw(1,4)-- (1,1); \draw(1,1)-- (1,-1.6); \draw(3.7,0.3)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (3.7,-1.6); \draw(6.7,0.3)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (6.7,-1.6); \draw(3.7,3.2)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (4.5,1); \draw(6.7,3.2)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (7.5,1); \draw(4.5,4)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (3.7,3.2); \draw(7.5,4)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (6.7,3.2); \draw(4.5,1)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (3.7,0.3); \draw(7.5,1)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (6.7,0.3); \draw(3.7,3.2)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (1,4); \draw(6.7,3.2)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (4.5,4); \draw(3.7,0.3)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (1,1); \draw(6.7,0.3)-- node {\tikz\draw[-triangle 60] (0,0) -- +(.05,0);} (4.5,1); \end{scope} \filldraw[fill=white] (4,7) circle (0.25cm); \filldraw[fill=white] (7,7) circle (0.25cm); \filldraw[fill=black] (1,4) circle (0.25cm); \filldraw[fill=black] (4.5,4) circle (0.25cm); \filldraw[fill=black] (7.5,4) circle (0.25cm); \filldraw[fill=white] (3.7,3.2) circle (0.25cm); \filldraw[fill=white] (6.7,3.2) circle (0.25cm); \filldraw[fill=black] (1,1) circle (0.25cm); \filldraw[fill=black] (4.5,1) circle (0.25cm); \filldraw[fill=black] (7.5,1) circle (0.25cm); \filldraw[fill=white] (3.7,0.3) circle (0.25cm); \filldraw[fill=white] (6.7,0.3) circle (0.25cm); \end{tikzpicture}
Figure 3.

Flows used to compute .

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=0.7] \filldraw[fill=white, thick] (0,5) circle (1.5cm); \node at (0,2) {\LARGE$P_{Sab}$}; \filldraw[fill=black] (-0.2,6.5) circle (0.08cm); \filldraw[fill=black] (0.9,6.2) circle (0.08cm); \filldraw[fill=black] (1.01,3.9) circle (0.08cm); \filldraw[fill=black] (-0.5,3.6) circle (0.08cm); \node at (-0.2,6.8) {\large$a$}; \node at (1.13,6.4) {\large$b$}; \node at (1.21,3.8) {\large$c$}; \node at (-0.5,3.3) {\large$d$}; \node at (2.5,5) {\huge+}; \filldraw[fill=white, thick] (5,5) circle (1.5cm); \node at (5,2) {\LARGE$P_{Scd}$}; \filldraw[fill=black] (4.8,6.5) circle (0.08cm); \filldraw[fill=black] (5.9,6.2) circle (0.08cm); \filldraw[fill=black] (6.01,3.9) circle (0.08cm); \filldraw[fill=black] (4.5,3.6) circle (0.08cm); \node at (4.8,6.8) {\large$a$}; \node at (6.13,6.4) {\large$b$}; \node at (6.21,3.8) {\large$c$}; \node at (4.5,3.3) {\large$d$}; \draw[thick] plot [smooth] coordinates { (4.8,6.5) (4.9,6.2) (4.6,5.7) (4.8,5.2) (4.4,4.6) (4.5,3.8)}; \draw[-triangle 60] (4.45,4.1) -- (4.5,3.76); \draw[thick] plot [smooth] coordinates { (5.9,6.2) (5.7,5.7) (5.9,5.3) (5.8,4.9) (6.1,4.5) (6.01,4.1)}; \draw[-triangle 60] (6.05,4.25) -- (6,4.05); \node at (4.2,5) {\large$w_1$}; \node at (5.5,5.4) {\large$w_2$}; \draw[thick] plot [smooth cycle] coordinates { (5.1,5) (5.25,4.95) (5.3,4.7) (5.2,4.4) (5.3,4.1) (5.2,3.9) (5.1,3.9) (4.9,4.4)}; \draw[-triangle 60] (4.9,4.4) -- (4.95,4.65); \node at (7.5,5) {\huge=}; \node at (7.5,5.5) {\LARGE?}; \filldraw[fill=white, thick] (10,5) circle (1.5cm); \node at (10,2) {\LARGE$P_{Sad}$}; \filldraw[fill=black] (9.8,6.5) circle (0.08cm); \filldraw[fill=black] (10.9,6.2) circle (0.08cm); \filldraw[fill=black] (11.01,3.9) circle (0.08cm); \filldraw[fill=black] (9.5,3.6) circle (0.08cm); \node at (9.8,6.8) {\large$a$}; \node at (11.13,6.4) {\large$b$}; \node at (11.21,3.8) {\large$c$}; \node at (9.5,3.3) {\large$d$}; \draw[thick] plot [smooth] coordinates { (10.9,6.2) (10.6,5.9) (10.7,5.1) (10.4,4.8) (9.8,4.4) (9.55,3.75)}; \draw[-triangle 60] (9.56,3.79) -- (9.54,3.73); \node at (10.5,4.4) {\large$w_3$}; \draw[thick] plot [smooth cycle] coordinates { (9.2,5.67) (9.2,5.07) (9.5,4.97) (9.8,5.27) (10.2,5.37) (10.25,5.77) (9.7,5.92)}; \draw[-triangle 60] (9.8,5.92) -- (9.76,5.93); \node at (12.5,5) {\huge+}; \filldraw[fill=white, thick] (15,5) circle (1.5cm); \node at (15,2) {\LARGE$P_{Sbc}$}; \filldraw[fill=black] (14.8,6.5) circle (0.08cm); \filldraw[fill=black] (15.9,6.2) circle (0.08cm); \filldraw[fill=black] (16.01,3.9) circle (0.08cm); \filldraw[fill=black] (14.5,3.6) circle (0.08cm); \node at (14.8,6.8) {\large$a$}; \node at (16.13,6.4) {\large$b$}; \node at (16.21,3.8) {\large$c$}; \node at (14.5,3.3) {\large$d$}; \draw[thick] plot [smooth] coordinates { (14.8,6.5) (14.8,5.8) (15.8,4.8) (16,4.07)}; \draw[-triangle 60] (16,4.1) -- (16.01,4.05); \node at (15.3,5.85) {\large$w_4$}; \end{tikzpicture}
Figure 4.

The labeling of the regions of a tropical hyperplane.

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=0.8] \draw[->, ultra thick] (0,0) -- (4.5,0); \draw[->, ultra thick] (0,0) -- (-3,-2); \draw[->, ultra thick] (0,0) -- (0,4.5); \node at (-2,1.5) {\huge1}; \node at (1,-2) {\huge2}; \node at (2,2) {\huge3}; \node at (0.4,0.3) {\large123}; \node at (-1,-1.1) {\large12}; \node at (0.3,2.3) {\large13}; \node at (2.3,-0.3) {\large23}; \end{tikzpicture}
Figure 5.

A nonregular subdivision of .

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=0.5] \filldraw[fill=gray] (3,8) rectangle ++(1,1); \draw[thick] (0,9) -- (9,9); \draw[thick] (9,9) -- (9,0); \draw[thick] (0,9) -- (9,0); \draw[thick] (6,9) -- (9,6); \draw[thick] (5,9) -- (6,8); \draw[thick] (7,7) -- (9,5); \draw[thick] (4,9) -- (5,8); \draw[thick] (6,7) -- (8,5); \draw[thick] (4,8) -- (6,6); \draw[thick] (7,5) -- (8,4); \draw[thick] (2,9) -- (3,8); \draw[thick] (4,7) -- (6,5); \draw[thick] (7,4) -- (9,2); \draw[thick] (1,9) -- (5,5); \draw[thick] (6,4) -- (9,1); \draw[thick] (1,8) -- (4,8); \draw[thick] (5,8) -- (9,8); \draw[thick] (2,7) -- (4,7); \draw[thick] (5,7) -- (7,7); \draw[thick] (8,7) -- (9,7); \draw[thick] (4,6) -- (5,6); \draw[thick] (6,6) -- (8,6); \draw[thick] (5,5) -- (7,5); \draw[thick] (8,5) -- (9,5); \draw[thick] (5,4) -- (7,4); \draw[thick] (8,4) -- (9,4); \draw[thick] (6,3) -- (9,3); \draw[thick] (3,9) -- (3,6); \draw[thick] (4,9) -- (4,7); \draw[thick] (4,6) -- (4,5); \draw[thick] (5,8) -- (5,6); \draw[thick] (5,5) -- (5,4); \draw[thick] (6,8) -- (6,7); \draw[thick] (6,6) -- (6,4); \draw[thick] (7,9) -- (7,7); \draw[thick] (7,6) -- (7,4); \draw[thick] (7,3) -- (7,2); \draw[thick] (8,9) -- (8,6); \draw[thick] (8,5) -- (8,1); \filldraw[fill=black] (0,9) circle (0.08cm); \filldraw[fill=black] (1,9) circle (0.08cm); \filldraw[fill=black] (2,9) circle (0.08cm); \filldraw[fill=black] (3,9) circle (0.08cm); \filldraw[fill=black] (4,9) circle (0.08cm); \filldraw[fill=black] (5,9) circle (0.08cm); \filldraw[fill=black] (6,9) circle (0.08cm); \filldraw[fill=black] (7,9) circle (0.08cm); \filldraw[fill=black] (8,9) circle (0.08cm); \filldraw[fill=black] (9,8) circle (0.08cm); \filldraw[fill=black] (9,7) circle (0.08cm); \filldraw[fill=black] (9,6) circle (0.08cm); \filldraw[fill=black] (9,5) circle (0.08cm); \filldraw[fill=black] (9,4) circle (0.08cm); \filldraw[fill=black] (9,3) circle (0.08cm); \filldraw[fill=black] (9,2) circle (0.08cm); \filldraw[fill=black] (9,1) circle (0.08cm); \end{tikzpicture}
Figure 6.

The dual arrangement of tropical pseudohyperplanes.

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=0.7] \draw[fill={rgb:black,1;white,3},draw={rgb:black,1;white,3}] (4.1,8.9) circle (0.48cm); \draw[fill=white,draw=white] (4.1,8.9) circle (0.17cm); \node at (6,11) {\huge4}; \node at (11,6) {\huge8}; \node at (2,3) {\huge12}; \node at (2.2,7.2) {\large1}; \filldraw[fill=black] (2,7) circle (0.08cm); \node at (2.2,8.3) {\large2}; \filldraw[fill=black] (2,8) circle (0.08cm); \node at (2.2,9.3) {\large3}; \filldraw[fill=black] (2,9) circle (0.08cm); \node at (7.2,9.3) {\large5}; \filldraw[fill=black] (7,9) circle (0.08cm); \node at (8.2,8.3) {\large6}; \filldraw[fill=black] (8,8) circle (0.08cm); \node at (9.2,7.3) {\large7}; \filldraw[fill=black] (9,7) circle (0.08cm); \node at (9.3,2.3) {\large9}; \filldraw[fill=black] (9,2) circle (0.08cm); \node at (8.2,2.3) {\large10}; \filldraw[fill=black] (8,2) circle (0.08cm); \node at (6.5,2.03) {\large11}; \filldraw[fill=black] (7,2) circle (0.08cm); \draw[{<[length=2mm]}-{>[length=2mm]}, thick] plot [smooth, tension=0.1] coordinates { (-1,8) (0,9) (4,9) (5,8) (7,8) (10,6)}; \draw[{<[length=2mm]}-{>[length=2mm]}, thick] plot [smooth, tension=0.1] coordinates { (0,7) (1,8) (4,8) (6,6) (7,6) (8,5) (10,5)}; \draw[{<[length=2mm]}-{>[length=2mm]}, thick] plot [smooth, tension=0.1] coordinates { (-1,10) (4.6,4.4) (7.4,4.4) (8,3) (10,3)}; \draw[{<[length=2mm]}-{>[length=2mm]}, thick] plot [smooth, tension=0.1] coordinates { (2,5) (3,6.7) (4,6.8) (4,8.7) (4.3,9) (10,9)}; \draw[{<[length=2mm]}-{>[length=2mm]}, thick] plot [smooth, tension=0.2] coordinates { (3.2,4) (4.2,5.3) (5.4,5.4) (5.5,7) (8,7) (8,8) (10,8)}; \draw[{<[length=2mm]}-{>[length=2mm]}, thick] plot [smooth, tension=0.15] coordinates { (3,10) (3,7.5) (5.3,5) (5.3,3.5) (10,-0.5)}; \draw[{<[length=2mm]}-{>[length=2mm]}, thick] plot [smooth, tension=0.1] coordinates { (5.2,2.2) (6,3) (7.5,3) (8,6) (9,6) (9,7) (10,7)}; \draw[{<[length=2mm]}-, thick] plot [smooth, tension=0.1] coordinates { (4.5,10) (5.5,9) (5.5,7.5) (6.7,6.5) (6.7,3.5) (8,2)}; \draw[{<[length=2mm]}-, thick] plot [smooth, tension=0.1] coordinates { (5.4,10) (6.5,8.5) (6.5,7.5) (9,5.5) (9,2)}; \draw[-{>[length=2mm]}, thick] plot [smooth, tension=0.1] coordinates { (9,2) (10,2)}; \draw[-{>[length=2mm]}, thick] plot [smooth, tension=0.2] coordinates { (9,2) (8.7,1.7) (8.6,-0.6) (8.4,-1)}; \draw[-{>[length=2mm]}, thick] plot [smooth, tension=0.5] coordinates { (8,2) (8.4,1.8) (9,1) (10,0.6)}; \draw[-{>[length=2mm]}, thick] plot [smooth, tension=0.5] coordinates { (8,2) (7.7,1.8) (7.6,1) (7.2,0.4)}; \draw[-{>[length=2mm]}, thick] plot [smooth, tension=0.1] coordinates { (7,2) (6.2,1.2)}; \draw[-{>[length=2mm]}, thick] plot [smooth, tension=0.5] coordinates { (9,7) (8.8,7.4) (8.8,10)}; \draw[-{>[length=2mm]}, thick] plot [smooth, tension=0.5] coordinates { (8,8) (7.8,8.4) (7.8,10)}; \draw[-{>[length=2mm]}, thick] plot [smooth, tension=0.5] coordinates { (7,9) (6.7,10)}; \draw[-{>[length=2mm]}, thick] plot [smooth, tension=0.5] coordinates { (2,9) (1.7,10)}; \draw[-{>[length=2mm]}, thick] plot [smooth, tension=0.5] coordinates { (2,8) (0.4,10)}; \draw[-{>[length=2mm]}, thick] plot [smooth, tension=0.5] coordinates { (2,7) (1.1,6.1)}; \end{tikzpicture}
Figure 7.

The planar tree corresponding to the marked point in Figure 6. Elements of are shown in white.

\renewcommand{\arraystretch}{1} \setlength{\unitlength}{1.0pt} \begin{tikzpicture}[scale=0.7] \filldraw[fill=white, thick] (0,0) circle (3.5cm); \node[circle, fill=black, draw=black, inner sep=0pt, minimum size=8pt] (1) at (1.75,3.031) {}; \node[circle, fill=black, draw=black, inner sep=0pt, minimum size=8pt] (2) at (3.031,1.75) {}; \node[circle, fill=black, draw=black, inner sep=0pt, minimum size=8pt] (3) at (3.5,0) {}; \node[circle, fill=white, draw=black, inner sep=0pt, minimum size=8pt] (4) at (3.031,-1.75) {}; \node[circle, fill=black, draw=black, inner sep=0pt, minimum size=8pt] (5) at (1.75,-3.031) {}; \node[circle, fill=black, draw=black, inner sep=0pt, minimum size=8pt] (6) at (0,-3.5) {}; \node[circle, fill=black, draw=black, inner sep=0pt, minimum size=8pt] (7) at (-1.75,-3.031) {}; \node[circle, fill=white, draw=black, inner sep=0pt, minimum size=8pt] (8) at (-3.031,-1.75) {}; \node[circle, fill=black, draw=black, inner sep=0pt, minimum size=8pt] (9) at (-3.5,0) {}; \node[circle, fill=black, draw=black, inner sep=0pt, minimum size=8pt] (10) at (-3.031,1.75) {}; \node[circle, fill=black, draw=black, inner sep=0pt, minimum size=8pt] (11) at (-1.75,3.031) {}; \node[circle, fill=white, draw=black, inner sep=0pt, minimum size=8pt] (12) at (0,3.5) {}; \draw[thick] (12) to[bend right] (1); \draw[thick] (12) to[bend right] (2); \draw[thick] (12) to[bend right] (3); \draw[thick] (12) to[bend left] (11); \draw[thick] (3) to[bend right] (4); \draw[thick] (4) to[bend right] (5); \draw[thick] (5) to[bend right] (8); \draw[thick] (6) to[bend right] (8); \draw[thick] (7) to[bend right] (8); \draw[thick] (8) to[bend right] (10); \draw[thick] (8) to[bend right] (9); \node at (2,3.5) {\Large1}; \node at (3.6,1.9) {\Large2}; \node at (4,0) {\Large3}; \node at (3.5,-1.9) {\Large4}; \node at (2,-3.5) {\Large5}; \node at (0,-4.1) {\Large6}; \node at (-2,-3.5) {\Large7}; \node at (-3.55,-1.9) {\Large8}; \node at (-4,0) {\Large9}; \node at (-3.6,2) {\Large10}; \node at (-2.1,3.5) {\Large11}; \node at (0,4) {\Large12}; \end{tikzpicture}
Figure 8.

A plabic graph with trip permutation , together with a perfect orientation with source set .

Graphic without alt text

Mathematical Fragments

Proposition 2.5 (Reference BGW03).

For any matroid, the dimension of its matroid polytope is , where is the number of connected components of .

Definition 3.1.

Given , we let denote . Let . For a nonzero polynomial, we denote by the set of all points such that, if we form the collection of numbers for ranging over , then the minimum of this collection is not unique. We say that is the tropical hypersurface associated to .

Definition 3.3.

Let and choose a subset which is disjoint from . Then is a three-term Plücker relation for the Grassmannian . Here denotes , etc.

Definition 3.5.

The tropical Grassmannian is the intersection of the tropical hypersurfaces , where ranges over all elements of the Plücker ideal. The Dressian is the intersection of the tropical hypersurfaces , where ranges over all three-term Plücker relations.

Theorem 3.9.

The positive tropical Grassmannian equals the positive Dressian .

Remark 3.10.

In Section 8 we describe many parametrizations of cells of , which were given by Postnikov using plabic graphs. Reference SW05, Proposition 2.5 says that if one has a subtraction-free rational map which surjects onto the positive part of a variety (for example a cluster chart), then the tropicalization of this map surjects onto the positive tropical part of the variety. Therefore we can tropicalize each parameterization from Theorem 8.8 – to obtain a parameterization of a positive tropical positroid variety (in particular, ). More specifically, we tropicalize by replacing the positive parameters (with ) with real parameters (with ) – and replacing products with sums and sums with minimums in the expressions for flow polynomials. Then Reference SW05, Proposition 2.5 say that this tropicalized map gives a parameterization of .

Definition 3.13.

Let . Then for and labeling horizontal and vertical wires of (i.e. and ), we define

We likewise define the tropical version. Let .

Lemma 3.16.

The collection of Plücker coordinates form a cluster for the cluster algebra structure Reference Sco06 on (the affine cone over the) Grassmannian . We call this the corectangles cluster. In particular, this collection of Plücker coordinates is algebraically independent, and all other Plücker coordinates can be written as Laurent polynomials with positive coefficients in the Plücker coordinates from the collection.

Example 3.17.

Figure 2 depicts the map . Since and are inverses, this example shows how to express each of the variables (as shown in Figure 1) in terms of the tropical Plücker coordinates . Note moreover that if we choose a normalization in tropical projective space (e.g. where ), then we can solve for the tropical Plücker coordinates in in terms of the ’s. For example, comparing Figure 1 and Figure 2, we see that if , then , , so , etc. In this example we see that from the collection together with the normalization , we can uniquely determine the Plücker coordinates . As in Lemma 3.16, this collection of Plücker coordinates is a cluster for the cluster algebras structure on the Grassmannian.

Lemma 3.18.

The map sending (with the convention that ) to is an injective map from to .

Theorem 4.2.

The following are equivalent.

The collection is a tropical Plücker vector.

The one-skeleta of and are the same.

Every face of is matroidal.

Theorem 4.3.

The following are equivalent.

The collection is a positive tropical Plücker vector.

Every face of is positroidal.

Theorem 5.1 (Reference ARW17, Theorem 5.1).

Let be a positively oriented matroid of rank on the ground set . Then is realizable.

Lemma 5.2.

Suppose that lies in the tropical Grassmannian . Then every face of the matroidal subdivision of corresponds to a realizable matroid.

Equation (5.3)
Theorem 6.2 (Reference Spe09).

Let be a tropical Plücker vector arising as for some . Then has at most interior faces of dimension , with equality if and only if all facets of correspond to series-parallel matroids.

Corollary 6.4.

A connected positroid is series-parallel if and only if it has no uniform matroid as a minor.

Corollary 6.5.

Let be a connected positroid. Then is series-parallel if and only if its matroid polytope does not contain any face which is an (unsubdivided) octahedron.

Theorem 6.6.

Let be a positive tropical Plücker vector. For the positroidal subdivision of , the following are equivalent:

(1)

is a finest subdivision.

(2)

Every facet of is the matroid polytope of a series-parallel matroid.

(3)

Every octahedron in is subdivided.

Corollary 6.7.

Every finest positroidal subdivision of achieves equality in the -vector theorem. In particular, such a positroidal subdivision has precisely facets.

Remark 7.1.

Note that the restriction of to the columns labeled by is the identity matrix.

Lemma 7.4.

Let . The following are equivalent:

(1)

The polytope is an -dimensional simplex.

(2)

The graph is a tree on the vertices .

Lemma 7.6.

Suppose that is a subset of such that is a tree. The matroid is positroidal if and only if we can embed the tree in a disk so that it is planar, and its vertices lie on the boundary of the disk in the standard circular order on .

Proposition 8.4.

Let be a plabic graph of type . Then we have a positroid on whose bases are precisely

where is the set of sources of .

Theorem 8.8.

Let be a reduced plabic graph of type , and choose a perfect orientation with source set . Then the map sending to the collection of flow polynomials is a homemorphism from to the corresponding positroid cell (realized in its Plücker embedding).

References

Reference [AD09]
Federico Ardila and Mike Develin, Tropical hyperplane arrangements and oriented matroids, Math. Z. 262 (2009), no. 4, 795–816, DOI 10.1007/s00209-008-0400-z. MR2511751,
Show rawAMSref \bib{ArdilaDevelin}{article}{ label={AD09}, author={Ardila, Federico}, author={Develin, Mike}, title={Tropical hyperplane arrangements and oriented matroids}, journal={Math. Z.}, volume={262}, date={2009}, number={4}, pages={795--816}, issn={0025-5874}, review={\MR {2511751}}, doi={10.1007/s00209-008-0400-z}, }
Reference [AHHLT19]
Nima Arkani-Hamed, Song He, Thomas Lam, and Hugh Thomas, Binary Geometries, Generalized Particles and Strings, and Cluster Algebras, 2019.
Reference [AHLS20]
Nima Arkani-Hamed, Thomas Lam, and Marcus Spradlin, Positive configuration space, 2020, Preprint, arXiv:2003.03904.
Reference [AHT14]
Nima Arkani-Hamed and Jaroslav Trnka, The amplituhedron, J. High Energy Phys., (10):33, 2014.
Reference [ARW16]
Federico Ardila, Felipe Rincón, and Lauren Williams, Positroids and non-crossing partitions, Trans. Amer. Math. Soc. 368 (2016), no. 1, 337–363, DOI 10.1090/tran/6331. MR3413866,
Show rawAMSref \bib{ARW}{article}{ label={ARW16}, author={Ardila, Federico}, author={Rinc\'{o}n, Felipe}, author={Williams, Lauren}, title={Positroids and non-crossing partitions}, journal={Trans. Amer. Math. Soc.}, volume={368}, date={2016}, number={1}, pages={337--363}, issn={0002-9947}, review={\MR {3413866}}, doi={10.1090/tran/6331}, }
Reference [ARW17]
Federico Ardila, Felipe Rincón, and Lauren Williams, Positively oriented matroids are realizable, J. Eur. Math. Soc. (JEMS) 19 (2017), no. 3, 815–833, DOI 10.4171/JEMS/680. MR3612868,
Show rawAMSref \bib{ARW2}{article}{ label={ARW17}, author={Ardila, Federico}, author={Rinc\'{o}n, Felipe}, author={Williams, Lauren}, title={Positively oriented matroids are realizable}, journal={J. Eur. Math. Soc. (JEMS)}, volume={19}, date={2017}, number={3}, pages={815--833}, issn={1435-9855}, review={\MR {3612868}}, doi={10.4171/JEMS/680}, }
Reference [BC19]
Francisco Borges and Freddy Cachazo, Generalized planar Feynman diagrams: collections, J. High Energy Phys. 11 (2020), Art. 164-27, DOI 10.1007/jhep11(2020)164. MR4204114,
Show rawAMSref \bib{Borges:2019csl}{article}{ label={BC19}, author={Borges, Francisco}, author={Cachazo, Freddy}, title={Generalized planar Feynman diagrams: collections}, journal={J. High Energy Phys.}, date={2020}, number={11}, pages={Art. 164-27}, issn={1126-6708}, review={\MR {4204114}}, doi={10.1007/jhep11(2020)164}, }
Reference [BGW03]
Alexandre V. Borovik, I. M. Gelfand, and Neil White, Coxeter matroids, Progress in Mathematics, vol. 216, Birkhäuser Boston, Inc., Boston, MA, 2003, DOI 10.1007/978-1-4612-2066-4. MR1989953,
Show rawAMSref \bib{Coxeter}{book}{ label={BGW03}, author={Borovik, Alexandre V.}, author={Gelfand, I. M.}, author={White, Neil}, title={Coxeter matroids}, series={Progress in Mathematics}, volume={216}, publisher={Birkh\"{a}user Boston, Inc., Boston, MA}, date={2003}, pages={xxii+264}, isbn={0-8176-3764-8}, review={\MR {1989953}}, doi={10.1007/978-1-4612-2066-4}, }
Reference [BLVS+99]
Anders Björner, Michel Las Vergnas, Bernd Sturmfels, Neil White, and Günter M. Ziegler, Oriented matroids, 2nd ed., Encyclopedia of Mathematics and its Applications, vol. 46, Cambridge University Press, Cambridge, 1999, DOI 10.1017/CBO9780511586507. MR1744046,
Show rawAMSref \bib{OrientedMatroidBook}{book}{ label={BLVS+99}, author={Bj\"{o}rner, Anders}, author={Las Vergnas, Michel}, author={Sturmfels, Bernd}, author={White, Neil}, author={Ziegler, G\"{u}nter M.}, title={Oriented matroids}, series={Encyclopedia of Mathematics and its Applications}, volume={46}, edition={2}, publisher={Cambridge University Press, Cambridge}, date={1999}, pages={xii+548}, isbn={0-521-77750-X}, review={\MR {1744046}}, doi={10.1017/CBO9780511586507}, }
Reference [Bru87]
Richard A. Brualdi, Transversal matroids, Combinatorial geometries, Encyclopedia Math. Appl., vol. 29, Cambridge Univ. Press, Cambridge, 1987, pp. 72–97. MR921069,
Show rawAMSref \bib{Brualdi}{article}{ label={Bru87}, author={Brualdi, Richard A.}, title={Transversal matroids}, conference={ title={Combinatorial geometries}, }, book={ series={Encyclopedia Math. Appl.}, volume={29}, publisher={Cambridge Univ. Press, Cambridge}, }, date={1987}, pages={72--97}, review={\MR {921069}}, }
Reference [CEGM19]
Freddy Cachazo, Nick Early, Alfredo Guevara, and Sebastian Mizera. Scattering Equations: From Projective Spaces to Tropical Grassmannians. JHEP, 06:039, 2019.
Reference [DFGK19]
James Drummond, Jack Foster, Ömer Gürdogan, and Chrysostomos Kalousios. Algebraic singularities of scattering amplitudes from tropical geometry. 2019.
Reference [DLRS10]
Jesús A. De Loera, Jörg Rambau, and Francisco Santos, Triangulations: Structures for algorithms and applications, Algorithms and Computation in Mathematics, vol. 25, Springer-Verlag, Berlin, 2010, DOI 10.1007/978-3-642-12971-1. MR2743368,
Show rawAMSref \bib{Triangulations}{book}{ label={DLRS10}, author={De Loera, Jes\'{u}s A.}, author={Rambau, J\"{o}rg}, author={Santos, Francisco}, title={Triangulations}, series={Algorithms and Computation in Mathematics}, volume={25}, subtitle={Structures for algorithms and applications}, publisher={Springer-Verlag, Berlin}, date={2010}, pages={xiv+535}, isbn={978-3-642-12970-4}, review={\MR {2743368}}, doi={10.1007/978-3-642-12971-1}, }
Reference [dS87]
Ilda P.F. da Silva, Quelques propriétés des matroides orientés, Ph.D. Dissertation, Université Paris VI, 1987.
Reference [DS04]
Mike Develin and Bernd Sturmfels, Tropical convexity, Doc. Math. 9 (2004), 1–27. MR2054977,
Show rawAMSref \bib{DevelinSturmfels}{article}{ label={DS04}, author={Develin, Mike}, author={Sturmfels, Bernd}, title={Tropical convexity}, journal={Doc. Math.}, volume={9}, date={2004}, pages={1--27}, issn={1431-0635}, review={\MR {2054977}}, }
Reference [Ear19a]
Nick Early, From weakly separated collections to matroid subdivisions, 2019, Preprint, arXiv:1910.11522.
Reference [Ear19b]
Nick Early, Planar kinematic invariants, matroid subdivisions and generalized Feynman diagrams, 2019.
Reference [ER96]
P. H. Edelman and V. Reiner, Free arrangements and rhombic tilings, Discrete Comput. Geom. 15 (1996), no. 3, 307–340, DOI 10.1007/BF02711498. MR1380397,
Show rawAMSref \bib{EdelmanReiner}{article}{ label={ER96}, author={Edelman, P. H.}, author={Reiner, V.}, title={Free arrangements and rhombic tilings}, journal={Discrete Comput. Geom.}, volume={15}, date={1996}, number={3}, pages={307--340}, issn={0179-5376}, review={\MR {1380397}}, doi={10.1007/BF02711498}, }
Reference [HJJS08]
Sven Herrmann, Anders Jensen, Michael Joswig, and Bernd Sturmfels, How to draw tropical planes, Electron. J. Combin. 16 (2009), no. 2, Special volume in honor of Anders Björner, Research Paper 6, 26. MR2515769,
Show rawAMSref \bib{Herrmann2008HowTD}{article}{ label={HJJS08}, author={Herrmann, Sven}, author={Jensen, Anders}, author={Joswig, Michael}, author={Sturmfels, Bernd}, title={How to draw tropical planes}, journal={Electron. J. Combin.}, volume={16}, date={2009}, number={2, Special volume in honor of Anders Bj\"{o}rner}, pages={Research Paper 6, 26}, review={\MR {2515769}}, }
Reference [HJS14]
Sven Herrmann, Michael Joswig, and David E. Speyer, Dressians, tropical Grassmannians, and their rays, Forum Math. 26 (2014), no. 6, 1853–1881, DOI 10.1515/forum-2012-0030. MR3334049,
Show rawAMSref \bib{Dressian}{article}{ label={HJS14}, author={Herrmann, Sven}, author={Joswig, Michael}, author={Speyer, David E.}, title={Dressians, tropical Grassmannians, and their rays}, journal={Forum Math.}, volume={26}, date={2014}, number={6}, pages={1853--1881}, issn={0933-7741}, review={\MR {3334049}}, doi={10.1515/forum-2012-0030}, }
Reference [HKT06]
Paul Hacking, Sean Keel, and Jenia Tevelev, Compactification of the moduli space of hyperplane arrangements, J. Algebraic Geom. 15 (2006), no. 4, 657–680, DOI 10.1090/S1056-3911-06-00445-0. MR2237265,
Show rawAMSref \bib{HKT}{article}{ label={HKT06}, author={Hacking, Paul}, author={Keel, Sean}, author={Tevelev, Jenia}, title={Compactification of the moduli space of hyperplane arrangements}, journal={J. Algebraic Geom.}, volume={15}, date={2006}, number={4}, pages={657--680}, issn={1056-3911}, review={\MR {2237265}}, doi={10.1090/S1056-3911-06-00445-0}, }
Reference [Hor16]
Silke Horn, A topological representation theorem for tropical oriented matroids, J. Combin. Theory Ser. A 142 (2016), 77–112, DOI 10.1016/j.jcta.2016.03.003. MR3499492,
Show rawAMSref \bib{Horn}{article}{ label={Hor16}, author={Horn, Silke}, title={A topological representation theorem for tropical oriented matroids}, journal={J. Combin. Theory Ser. A}, volume={142}, date={2016}, pages={77--112}, issn={0097-3165}, review={\MR {3499492}}, doi={10.1016/j.jcta.2016.03.003}, }
Reference [HP19]
Niklas Henke and Georgios Papathanasiou, How tropical are seven- and eight-particle amplitudes?, J. High Energy Phys. 8 (2020), Art. 005-49, DOI 10.1007/jhep08(2020)005. MR4190219,
Show rawAMSref \bib{Henke:2019hve}{article}{ label={HP19}, author={Henke, Niklas}, author={Papathanasiou, Georgios}, title={How tropical are seven- and eight-particle amplitudes?}, journal={J. High Energy Phys.}, date={2020}, number={8}, pages={Art. 005-49}, issn={1126-6708}, review={\MR {4190219}}, doi={10.1007/jhep08(2020)005}, }
Reference [HRS00]
Birkett Huber, Jörg Rambau, and Francisco Santos, The Cayley trick, lifting subdivisions and the Bohne-Dress theorem on zonotopal tilings, J. Eur. Math. Soc. (JEMS) 2 (2000), no. 2, 179–198, DOI 10.1007/s100970050003. MR1763304,
Show rawAMSref \bib{Cayley1}{article}{ label={HRS00}, author={Huber, Birkett}, author={Rambau, J\"{o}rg}, author={Santos, Francisco}, title={The Cayley trick, lifting subdivisions and the Bohne-Dress theorem on zonotopal tilings}, journal={J. Eur. Math. Soc. (JEMS)}, volume={2}, date={2000}, number={2}, pages={179--198}, issn={1435-9855}, review={\MR {1763304}}, doi={10.1007/s100970050003}, }
Reference [Kap93]
M. M. Kapranov, Chow quotients of Grassmannians. I, I. M. Gel′fand Seminar, Adv. Soviet Math., vol. 16, Amer. Math. Soc., Providence, RI, 1993, pp. 29–110. MR1237834,
Show rawAMSref \bib{Kapranov}{article}{ label={Kap93}, author={Kapranov, M. M.}, title={Chow quotients of Grassmannians. I}, conference={ title={I. M. Gel\cprime fand Seminar}, }, book={ series={Adv. Soviet Math.}, volume={16}, publisher={Amer. Math. Soc., Providence, RI}, }, date={1993}, pages={29--110}, review={\MR {1237834}}, }
Reference [KT06]
Sean Keel and Jenia Tevelev, Geometry of Chow quotients of Grassmannians, Duke Math. J. 134 (2006), no. 2, 259–311, DOI 10.1215/S0012-7094-06-13422-1. MR2248832,
Show rawAMSref \bib{KT}{article}{ label={KT06}, author={Keel, Sean}, author={Tevelev, Jenia}, title={Geometry of Chow quotients of Grassmannians}, journal={Duke Math. J.}, volume={134}, date={2006}, number={2}, pages={259--311}, issn={0012-7094}, review={\MR {2248832}}, doi={10.1215/S0012-7094-06-13422-1}, }
Reference [LPW20]
Tomasz Lukowski, Matteo Parisi, and Lauren K. Williams. The positive tropical grassmannian, the hypersimplex, and the m=2 amplituhedron, 2020. Preprint, arXiv:2002.06164.
Reference [Lus94]
G. Lusztig, Total positivity in reductive groups, Lie theory and geometry, Progr. Math., vol. 123, Birkhäuser Boston, Boston, MA, 1994, pp. 531–568, DOI 10.1007/978-1-4612-0261-5_20. MR1327548,
Show rawAMSref \bib{lusztig}{article}{ label={Lus94}, author={Lusztig, G.}, title={Total positivity in reductive groups}, conference={ title={Lie theory and geometry}, }, book={ series={Progr. Math.}, volume={123}, publisher={Birkh\"{a}user Boston, Boston, MA}, }, date={1994}, pages={531--568}, review={\MR {1327548}}, doi={10.1007/978-1-4612-0261-5\_20}, }
Reference [Mar19]
Cameron Marcott, Basis shape loci and the positive grassmannian, 2019.
Reference [OPS15]
Suho Oh, Alexander Postnikov, and David E. Speyer, Weak separation and plabic graphs, Proc. Lond. Math. Soc. (3) 110 (2015), no. 3, 721–754, DOI 10.1112/plms/pdu052. MR3342103,
Show rawAMSref \bib{OPS}{article}{ label={OPS15}, author={Oh, Suho}, author={Postnikov, Alexander}, author={Speyer, David E.}, title={Weak separation and plabic graphs}, journal={Proc. Lond. Math. Soc. (3)}, volume={110}, date={2015}, number={3}, pages={721--754}, issn={0024-6115}, review={\MR {3342103}}, doi={10.1112/plms/pdu052}, }
Reference [OPS19]
Jorge Alberto Olarte, Marta Panizzut, and Benjamin Schröter, On local Dressians of matroids, Algebraic and geometric combinatorics on lattice polytopes, World Sci. Publ., Hackensack, NJ, 2019, pp. 309–329. MR3971702,
Show rawAMSref \bib{Olarte}{article}{ label={OPS19}, author={Olarte, Jorge Alberto}, author={Panizzut, Marta}, author={Schr\"{o}ter, Benjamin}, title={On local Dressians of matroids}, conference={ title={Algebraic and geometric combinatorics on lattice polytopes}, }, book={ publisher={World Sci. Publ., Hackensack, NJ}, }, date={2019}, pages={309--329}, review={\MR {3971702}}, }
Reference [OS17]
Su Ho Oh and David E. Speyer, Links in the complex of weakly separated collections, J. Comb. 8 (2017), no. 4, 581–592, DOI 10.4310/JOC.2017.v8.n4.a2. MR3682392,
Show rawAMSref \bib{OhSpeyer}{article}{ label={OS17}, author={Oh, Su Ho}, author={Speyer, David E.}, title={Links in the complex of weakly separated collections}, journal={J. Comb.}, volume={8}, date={2017}, number={4}, pages={581--592}, issn={2156-3527}, review={\MR {3682392}}, doi={10.4310/JOC.2017.v8.n4.a2}, }
Reference [Oxl11]
James Oxley, Matroid theory, 2nd ed., Oxford Graduate Texts in Mathematics, vol. 21, Oxford University Press, Oxford, 2011, DOI 10.1093/acprof:oso/9780198566946.001.0001. MR2849819,
Show rawAMSref \bib{Oxley}{book}{ label={Oxl11}, author={Oxley, James}, title={Matroid theory}, series={Oxford Graduate Texts in Mathematics}, volume={21}, edition={2}, publisher={Oxford University Press, Oxford}, date={2011}, pages={xiv+684}, isbn={978-0-19-960339-8}, review={\MR {2849819}}, doi={10.1093/acprof:oso/9780198566946.001.0001}, }
Reference [Pay09]
Sam Payne, Fibers of tropicalization, Math. Z. 262 (2009), no. 2, 301–311, DOI 10.1007/s00209-008-0374-x. MR2504879,
Show rawAMSref \bib{Payne1}{article}{ label={Pay09}, author={Payne, Sam}, title={Fibers of tropicalization}, journal={Math. Z.}, volume={262}, date={2009}, number={2}, pages={301--311}, issn={0025-5874}, review={\MR {2504879}}, doi={10.1007/s00209-008-0374-x}, }
Reference [Pay12]
Sam Payne, Erratum to: Fibers of tropicalization [MR2504879], Math. Z. 272 (2012), no. 3-4, 1403–1406, DOI 10.1007/s00209-012-1080-2. MR2995174,
Show rawAMSref \bib{Payne2}{article}{ label={Pay12}, author={Payne, Sam}, title={Erratum to: Fibers of tropicalization [MR2504879]}, journal={Math. Z.}, volume={272}, date={2012}, number={3-4}, pages={1403--1406}, issn={0025-5874}, review={\MR {2995174}}, doi={10.1007/s00209-012-1080-2}, }
Reference [Pos]
Alexander Postnikov, Total positivity, Grassmannians, and networks, Preprint, http://math.mit.edu/˜apost/papers/tpgrass.pdf.
Reference [PSW09]
Alexander Postnikov, David Speyer, and Lauren Williams, Matching polytopes, toric geometry, and the totally non-negative Grassmannian, J. Algebraic Combin. 30 (2009), no. 2, 173–191, DOI 10.1007/s10801-008-0160-1. MR2525057,
Show rawAMSref \bib{PSW}{article}{ label={PSW09}, author={Postnikov, Alexander}, author={Speyer, David}, author={Williams, Lauren}, title={Matching polytopes, toric geometry, and the totally non-negative Grassmannian}, journal={J. Algebraic Combin.}, volume={30}, date={2009}, number={2}, pages={173--191}, issn={0925-9899}, review={\MR {2525057}}, doi={10.1007/s10801-008-0160-1}, }
Reference [Rin13]
Felipe Rincón, Local tropical linear spaces, Discrete Comput. Geom. 50 (2013), no. 3, 700–713, DOI 10.1007/s00454-013-9519-8. MR3102586,
Show rawAMSref \bib{Felipe}{article}{ label={Rin13}, author={Rinc\'{o}n, Felipe}, title={Local tropical linear spaces}, journal={Discrete Comput. Geom.}, volume={50}, date={2013}, number={3}, pages={700--713}, issn={0179-5376}, review={\MR {3102586}}, doi={10.1007/s00454-013-9519-8}, }
Reference [San05]
Francisco Santos, The Cayley trick and triangulations of products of simplices, Integer points in polyhedra—geometry, number theory, algebra, optimization, Contemp. Math., vol. 374, Amer. Math. Soc., Providence, RI, 2005, pp. 151–177, DOI 10.1090/conm/374/06904. MR2134766,
Show rawAMSref \bib{Cayley2}{article}{ label={San05}, author={Santos, Francisco}, title={The Cayley trick and triangulations of products of simplices}, conference={ title={Integer points in polyhedra---geometry, number theory, algebra, optimization}, }, book={ series={Contemp. Math.}, volume={374}, publisher={Amer. Math. Soc., Providence, RI}, }, date={2005}, pages={151--177}, review={\MR {2134766}}, doi={10.1090/conm/374/06904}, }
Reference [Sco06]
Joshua S. Scott, Grassmannians and cluster algebras, Proc. London Math. Soc. (3) 92 (2006), no. 2, 345–380, DOI 10.1112/S0024611505015571. MR2205721,
Show rawAMSref \bib{Scott}{article}{ label={Sco06}, author={Scott, Joshua S.}, title={Grassmannians and cluster algebras}, journal={Proc. London Math. Soc. (3)}, volume={92}, date={2006}, number={2}, pages={345--380}, issn={0024-6115}, review={\MR {2205721}}, doi={10.1112/S0024611505015571}, }
Reference [Spe08]
David E. Speyer, Tropical linear spaces, SIAM J. Discrete Math. 22 (2008), no. 4, 1527–1558, DOI 10.1137/080716219. MR2448909,
Show rawAMSref \bib{Speyer}{article}{ label={Spe08}, author={Speyer, David E.}, title={Tropical linear spaces}, journal={SIAM J. Discrete Math.}, volume={22}, date={2008}, number={4}, pages={1527--1558}, issn={0895-4801}, review={\MR {2448909}}, doi={10.1137/080716219}, }
Reference [Spe09]
David E. Speyer, A matroid invariant via the -theory of the Grassmannian, Adv. Math. 221 (2009), no. 3, 882–913, DOI 10.1016/j.aim.2009.01.010. MR2511042,
Show rawAMSref \bib{Ktheory}{article}{ label={Spe09}, author={Speyer, David E.}, title={A matroid invariant via the $K$-theory of the Grassmannian}, journal={Adv. Math.}, volume={221}, date={2009}, number={3}, pages={882--913}, issn={0001-8708}, review={\MR {2511042}}, doi={10.1016/j.aim.2009.01.010}, }
Reference [SS04]
David Speyer and Bernd Sturmfels, The tropical Grassmannian, Adv. Geom. 4 (2004), no. 3, 389–411, DOI 10.1515/advg.2004.023. MR2071813,
Show rawAMSref \bib{tropgrass}{article}{ label={SS04}, author={Speyer, David}, author={Sturmfels, Bernd}, title={The tropical Grassmannian}, journal={Adv. Geom.}, volume={4}, date={2004}, number={3}, pages={389--411}, issn={1615-715X}, review={\MR {2071813}}, doi={10.1515/advg.2004.023}, }
Reference [SW05]
David Speyer and Lauren Williams, The tropical totally positive Grassmannian, J. Algebraic Combin. 22 (2005), no. 2, 189–210, DOI 10.1007/s10801-005-2513-3. MR2164397,
Show rawAMSref \bib{troppos}{article}{ label={SW05}, author={Speyer, David}, author={Williams, Lauren}, title={The tropical totally positive Grassmannian}, journal={J. Algebraic Combin.}, volume={22}, date={2005}, number={2}, pages={189--210}, issn={0925-9899}, review={\MR {2164397}}, doi={10.1007/s10801-005-2513-3}, }
Reference [Tal08]
Kelli Talaska, A formula for Plücker coordinates associated with a planar network, Int. Math. Res. Not. IMRN, posted on 2008, Art. ID rnn 081, 19, DOI 10.1093/imrn/rnn081. MR2439562,
Show rawAMSref \bib{Talaska}{article}{ label={Tal08}, author={Talaska, Kelli}, title={A formula for Pl\"{u}cker coordinates associated with a planar network}, journal={Int. Math. Res. Not. IMRN}, date={2008}, pages={Art. ID rnn 081, 19}, issn={1073-7928}, review={\MR {2439562}}, doi={10.1093/imrn/rnn081}, }
Reference [Tal11]
Kelli Talaska, Combinatorial formulas for -coordinates in a totally nonnegative Grassmannian, J. Combin. Theory Ser. A 118 (2011), no. 1, 58–66, DOI 10.1016/j.jcta.2009.10.006. MR2737184,
Show rawAMSref \bib{Talaska2}{article}{ label={Tal11}, author={Talaska, Kelli}, title={Combinatorial formulas for $\Gamma $-coordinates in a totally nonnegative Grassmannian}, journal={J. Combin. Theory Ser. A}, volume={118}, date={2011}, number={1}, pages={58--66}, issn={0097-3165}, review={\MR {2737184}}, doi={10.1016/j.jcta.2009.10.006}, }
Reference [Whi86]
Neil White (ed.), Theory of matroids, Encyclopedia of Mathematics and its Applications, vol. 26, Cambridge University Press, Cambridge, 1986, DOI 10.1017/CBO9780511629563. MR849389,
Show rawAMSref \bib{White}{collection}{ label={Whi86}, title={Theory of matroids}, series={Encyclopedia of Mathematics and its Applications}, volume={26}, editor={White, Neil}, publisher={Cambridge University Press, Cambridge}, date={1986}, pages={xviii+316}, isbn={0-521-30937-9}, review={\MR {849389}}, doi={10.1017/CBO9780511629563}, }

Article Information

MSC 2020
Primary: 05E99 (None of the above, but in this section)
Secondary: 14M15 (Grassmannians, Schubert varieties, flag manifolds)
Author Information
David Speyer
Department of Mathematics, University of Michigan, 530 Church Street, Ann Arbor, Michigan 48109
speyer@umich.edu
MathSciNet
Lauren K. Williams
Department of Mathematics, Harvard University, 1 Oxford Street, Cambridge, Massachusetts 02138
williams@math.harvard.edu
MathSciNet
Additional Notes

The first author was partially supported by NSF grants DMS-1855135 and DMS-1854225. The second author was partially supported by NSF grants DMS-1854316 and DMS-1854512.

Journal Information
Transactions of the American Mathematical Society, Series B, Volume 8, Issue 11, ISSN 2330-0000, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2021 by the authors under Creative Commons Attribution-Noncommercial 3.0 License (CC BY NC 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/btran/67
  • MathSciNet Review: 4241765
  • Show rawAMSref \bib{4241765}{article}{ author={Speyer, David}, author={Williams, Lauren}, title={The positive Dressian equals the positive tropical Grassmannian}, journal={Trans. Amer. Math. Soc. Ser. B}, volume={8}, number={11}, date={2021}, pages={330-353}, issn={2330-0000}, review={4241765}, doi={10.1090/btran/67}, }

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.