Polyhedral and tropical geometry of flag positroids
Abstract.
A flag positroid of ranks on is a flag matroid that can be realized by a real matrix such that the minors of involving rows are nonnegative for all . In this paper we explore the polyhedral and tropical geometry of flag positroids, particularly when is a sequence of consecutive numbers. In this case we show that the nonnegative tropical flag variety equals the nonnegative flag Dressian , and that the points of give rise to coherent subdivisions of the flag positroid polytope into flag positroid polytopes. Our results have applications to Bruhat interval polytopes: for example, we show that a complete flag matroid polytope is a Bruhat interval polytope if and only if its -dimensional faces are Bruhat interval polytopes. Our results also have applications to realizability questions. We define a positively oriented flag matroid to be a sequence of positively oriented matroids which is also an oriented flag matroid. We then prove that every positively oriented flag matroid of ranks is realizable.
1. Introduction
In recent years there has been a great deal of interest in the tropical Grassmannian [SS04, HJJS08, HJS14, CEGM19, Bos21], and matroid polytopes and their subdivisions [Spe08, AFR10, Ear22], as well as “positive” [Pos, SW05, Oh08, ARW16, LF19, LPW20, SW21, AHLS20] and “flag” [TW15, BEZ21, BLMM17, JL22, JLLO, Bor22] versions of the above objects. The aim of this paper is to illustrate the beautiful relationships between the nonnegative tropical flag variety, the nonnegative flag Dressian, and flag positroid polytopes and their subdivisions, unifying and generalizing some of the existing results. We will particularly focus on the case of flag varieties (respectively, flag positroids) consisting of subspaces (respectively, matroids) of consecutive ranks. This case includes both Grassmannians and complete flag varieties.
For positive integers and with , we let denote the set and we let denote the collection of all -element subsets of . Given a subset we let denote the sum of standard basis vectors . For a collection , we let
The collection is said to define a matroid of rank on if every edge of the polytope is parallel to for some . In this case, we call the set of bases of , and define the matroid polytope of to be the polytope . When indexes the nonvanishing Plücker coordinates of an element of the Grassmannian , we say that realizes , and it is well-known that is the moment map image of the closure of the torus orbit of in the Grassmannian [GGMS87]. We assume familiarity with the fundamentals of matroid theory as in [Oxl11] and [BGW03].
The above definition of matroid in terms of its polytope is due to [GGMS87]. Flag matroids are natural generalizations of matroids that admit the following polytopal definition.
Definition 1.1.
[BGW03, Corollary 1.13.5 and Theorem 1.13.6] Let be a sequence of increasing integers in . A flag matroid of ranks on is a sequence of matroids of ranks on such that all vertices of the polytope
are equidistant from the origin. The polytope is called the flag matroid polytope of ; we sometimes say it is a flag matroid polytope of rank .
Flag matroids are exactly the type objects in the theory of Coxeter matroids [GS87, BGW03]. Just as a realization of a matroid is a point in a Grassmannian, a realization of a flag matroid is a point in a flag variety. More concretely, a realization of a flag matroid of ranks is an matrix over a field such that for each , the submatrix of formed by the first rows of is a realization of . For an equivalent definition of flag matroids in terms of Plücker relations on partial flag varieties, see [JL22, Proposition A].
There is a notion of moment map for any flag variety (indeed for any generalized partial flag variety ) [GS87, BGW03]. When a flag matroid can be realized by a point in the flag variety, then its matroid polytope is the moment map image of the closure of the torus orbit of in the flag variety [GS87], [BGW03, Corollary 1.13.5].
There are natural “positive” analogues of matroids, flag matroids, and their polytopes.
Definition 1.2.
Let be a sequence of increasing integers in . We say that a flag matroid of ranks on is a flag positroid if it has a realization by a real matrix such that the submatrix of formed by the first rows of has all nonnegative minors for each .
We refer to the flag matroid polytope of a flag positroid as a flag positroid polytope. It follows from our definition above that flag positroids are realizable.
Setting in 1.2 gives the well-studied notion of positroids and positroid polytopes [Pos, Oh08, ARW16]. Therefore each flag positroid is a sequence of positroids.
In recent years it has been gradually understood that the tropical geometry of the Grassmannian and flag variety, and in particular, the Dressian and flag Dressian, are intimately connected to (flag) matroid polytopes and their subdivisions [Spe08, HJJS08, BEZ21] (see also [MS15, §4]). A particularly attractive point of view, which sheds light on the above connections, is the theory of (flag) matroids over hyperfields [BB19, JL22]. In this framework, the Dressian and flag Dressian are the Grassmannian and flag variety over the tropical hyperfield, while matroids and flag matroids are the points of the Grassmannian and flag variety over the Krasner hyperfield.
The tropical geometry of the positive Grassmannian and flag variety are particularly nice: the positive tropical Grassmannian equals the positive Dressian, whose cones in turn parameterize subdivisions of the hypersimplex into positroid polytopes [SW05, SW21, LPW20, AHLS20]. And the positive tropical complete flag variety equals the positive complete flag Dressian, whose cones parameterize subdivisions of the permutohedron into Bruhat interval polytopes [Bor22, JLLO]. A below unifies and generalizes the above results.
Definition 1.3.
Let be the set underlying the tropical hyperfield, endowed with the topology such that is a homeomorphism. Given a point , we define the support of to be . When is the set of bases of a matroid, we identify with that matroid. Let be the tropical projective space of , which is defined as , where if for some .
Our main result is the following.
Theorem A.
Suppose is a sequence of consecutive integers for some . Then, for , the following statements are equivalent:
-
(a)
, the nonnegative tropicalization of the flag variety, i.e. the closure of the coordinate-wise valuation of points in .
-
(b)
, the nonnegative flag Dressian, i.e. the “solutions" to the positive-tropical Grassmann-Plücker and incidence-Plücker relations.
-
(c)
Every face in the coherent subdivision of the polytope induced by is a flag positroid polytope (of rank ).
-
(d)
Every face of dimension at most 2 in the subdivision of is a flag positroid polytope (of rank ).
-
(e)
The support of is a flag matroid, satisfies every three-term positive-tropical incidence relation when (respectively, every three-term positive-tropical Grassmann-Plücker relation when ), and either consists of uniform matroids or for at least one .
For the definitions of the objects in A, see 3.6 for (a), 3.3 for (b), 5.1 for (c), and 3.8 for (e).
We note that if is a single integer, A describes the relationship between the nonnegative tropical Grassmannian, the nonnegative Dressian, and subdivisions of positroid polytopes (e.g. the hypersimplex, if has no coordinates equal to ) into positroid polytopes. And when , A describes the relationship between the nonnegative tropical complete flag variety, the nonnegative complete flag Dressian, and subdivisions of Bruhat interval polytopes (e.g. the permutohedron, if has no coordinates equal to ) into Bruhat interval polytopes. We illustrate this relationship in the case where has no coordinates equal to in Figure 1.
We prove the equivalence (a)(b) in Section 3.2, the implications (b)(c)(d)(b) in Section 5.2, and the equivalence (b)(e) in Section 6.1.
A has applications to flag positroid polytopes.
Corollary 1.4.
For a flag matroid of consecutive ranks , its flag matroid polytope is a flag positroid polytope if and only if its -dimensional faces are flag positroid polytopes (of rank ).
Proof.
In the Grassmannian case, that is, the case that is a single integer, the flag positroid polytopes of rank are precisely the positroid polytopes, and in that case the above corollary appeared as [LPW20, Theorem 3.9].
Also in the Grassmannian case, the objects discussed in A are closely related to questions of realizability. Note that by definition, every positroid has a realization by a matrix whose Plücker coordinates are nonnegative, so it naturally defines a positively oriented matroid, that is, an oriented matroid defined by a chirotope whose values are all and . Conversely, every positively oriented matroid can be realized by a positroid: this was first proved in [ARW17] using positroid polytopes, and subsequently in [SW21], using the positive tropical Grassmannian. It is natural then to ask if there is an analogous realizability statement in the setting of flag matroids, and if one can characterize when a sequence of positroids forms a flag positroid; indeed, this was part of the motivation for [BCTJ22], which studied quotients of uniform positroids. Note however that questions of realizability for flag matroids are rather subtle: for example, a sequence of positroids that form a realizable flag matroid can still fail to be a flag positroid (see 4.4). By working with oriented flag matroids, we give an answer to this realizability question in 1.5, in the case of consecutive ranks.
Corollary 1.5.
Suppose is a sequence of positroids on of consecutive ranks . Then, when considered as a sequence of positively oriented matroids, is a flag positroid if and only if it is an oriented flag matroid.
We define a positively oriented flag matroid to be a sequence of positively oriented matroids which is also an oriented flag matroid. 1.5 then says that every positively oriented flag matroid of consecutive ranks is realizable.
See Section 4.1 for a review of oriented matroids and oriented flag matroids. Note that because a positroid by definition has a realization over with all nonnegative minors, it defines a positively oriented matroid. In Section 4.2, we deduce 1.5 from the equivalence of (a) and (b) in A. Another proof using ideas from discrete convex analysis is sketched in 4.7. In both proofs, the consecutive ranks condition is indispensable. We do not know whether the corollary holds if fails to satisfy the consecutive rank condition.
Question 1.6.
Suppose and are positroids on such that, when considered as positively oriented matroids, they form an oriented flag matroid . Is then a flag positroid?
One may attempt to answer the question by appealing to the fact [Kun86, Exercise 8.14] that for a flag matroid , one can always find a flag matroid of consecutive ranks such that and . However, the analogous statement fails for flag positroids: See 4.6 for an example of a flag positroid on of ranks such that there is no flag positroid with rank of equal to 2.
The consecutive rank condition has recently shown up in [BK22], which studied the relation between two notions of total positivity for partial flag varieties, “Lusztig positivity” and “Plücker positivity” (see Section 2.1). In particular, the Plücker positive subset of a partial flag variety agrees with the Lusztig positive subset of the partial flag variety precisely when the flag variety consists of linear subspaces of consecutive ranks [BK22, Theorem 1.1].
A generalized Bruhat interval polytope [TW15, Definition 7.8 and Lemma 7.9] can be defined as the moment map image of the closure of the torus orbit of a point in the nonnegative part (in the sense of Lusztig) of a flag variety . When is a sequence of consecutive integers, it then follows from [BK22] that generalized Bruhat interval polytopes for are precisely the flag positroid polytopes of ranks . In the complete flag case, a generalized Bruhat interval polytope is just a Bruhat interval polytope [KW15], that is, the convex hull of the permutation vectors for all permutations lying in some Bruhat interval .
We can now restate 1.4 as follows.
Corollary 1.7.
For a flag matroid on of consecutive ranks , its flag matroid polytope is a generalized Bruhat interval polytope if and only if its -dimensional faces are generalized Bruhat interval polytopes. In particular, for a complete flag matroid on , its flag matroid polytope is a Bruhat interval polytope if and only if its -dimensional faces are Bruhat interval polytopes.
The structure of this paper is as follows. In Section 2, we give background on total positivity and Bruhat interval polytopes. In Section 3, we introduce the tropical flag variety, the flag Dressian, and nonnegative analogues of these objects; we also prove the equivalence of (a) and (b) in A. In Section 4 we discuss positively oriented flag matroids and prove 1.5. In Section 5 we explain the relation between the flag Dressian and subdivisions of flag matroid polytopes, then prove that (b)(c)(d)(b) in A. We prove some key results about three-term incidence and Grassmann-Plücker relations in Section 6, which allow us to prove (b)(e) in A. Section 7 concerns projections of positive Richardsons to positroids: we characterize the positroid constituents of complete flag positroids, and we characterize when two adjacent-rank positroids form an oriented matroid quotient, or equivalently, can appear as constituents of a complete flag positroid. In Section 8, we make some remarks about the various fan structures for ; we then discuss fan structures and coherent subdivisions in the case of the Grassmannian and complete flag variety, including a detailed look at the case of .
Acknowledgements
The first author is supported by the Natural Sciences and Engineering Research Council of Canada (NSERC). Le premier auteur a été financé par le Conseil de recherches en sciences naturelles et en génie du Canada (CRSNG), [Ref. no. 557353-2021]. The second author is partially supported by the US National Science Foundation (DMS-2001854). The third author is partially supported by the National Science Foundation (DMS-1854512 and DMS-2152991). We are grateful to Tony Bloch, Michael Joswig, Steven Karp, Georg Loho, Dante Luber, and Jorge Alberto Olarte for sharing their work with us, which partially inspired this project. We also thank Yue Ren, Vasu Tewari, and Jorge Olarte for useful comments. We are grateful to Lara Bossinger for several invaluable discussions about fan structures. Lastly, we thank Jidong Wang for pointing out an error in a previous version of this paper.
2. Background on total positivity and Bruhat interval polytopes
2.1. Background on total positivity
Let and let For a field , let , and let denote the parabolic subgroup of of block upper-triangular matrices with diagonal blocks of sizes . We define the partial flag variety
As usual, we identify with the variety of partial flags of subspaces in :
We write for the complete flag variety . Note that can be identified with , where is the subgroup of upper-triangular matrices. There is a natural projection from to any partial flag variety by simply forgetting some of the subspaces.
If is an matrix such that is the span of the first rows, we say that is a realization of . Given any realization of and any , we have the Plücker coordinates or flag minors where ; concretely, is the determinant of the submatrix of occupying the first rows and columns . This gives the Plücker embedding of into taking to .
We now let be the field of real numbers. With this understanding, we will often drop the from our notation.
Definition 2.1.
We say that a real matrix is totally positive if all of its minors are positive. We let denote the subset of of totally positive matrices.
There are two natural ways to define positivity for partial flag varieties. The first notion comes from work of Lusztig [Lus94]. The second notion uses Plücker coordinates, and was initiated in work of Postnikov [Pos].
Definition 2.2.
We define the (Lusztig) positive part of , denoted by , as the image of inside . We define the (Lusztig) nonnegative part of , denoted by , as the closure of in the Euclidean topology.
We define the Plücker positive part (respectively, Plücker nonnegative part) of to be the subset of where all Plücker coordinates are positive (respectively, nonnegative).111The reader who is concerned about the fact that we are working with projective coordinates can replace “all Plücker coordinates are positive” by “all Plücker coordinates are nonzero and have the same sign”.
It is well-known that the Lusztig positive part of is a subset of the Plücker positive part of , and that the two notions agree in the case of the Grassmannian [TW13, Corollary 1.2]. The two notions also agree in the case of the complete flag variety [Bor22, Theorem 5.21]. More generally, we have the following.
Theorem 2.3.
[BK22, Theorem 1.1] The Lusztig positive (respectively, Lusztig nonnegative) part of equals the Plücker positive (respectively, Plücker nonnegative) part of if and only if the set consists of consecutive integers.
See [BK22, Section 1.4] for more references and a nice discussion of the history. Since in this paper we will be mainly studying the case where consists of consecutive integers, we will use the two notions interchangeably when there is no ambiguity.
Let and be the opposite Borel subgroups consisting of upper-triangular and lower-triangular matrices. Let be the Weyl group of . Given , the Richardson variety is the intersection of opposite Bruhat cells
where and denote permutation matrices in representing and . It is well-known that is nonempty precisely when in Bruhat order, and in that case is irreducible of dimension .
For with , let be the positive part of the Richardson variety. Lusztig conjectured and Rietsch proved [Rie98] that
(1) |
is a cell decomposition of . Moreover, Rietsch showed that one obtains a cell decomposition of the nonnegative partial flag variety by projecting the cell decomposition of [Rie98], [Rie06, Section 6]. Specifically, if we let be the parabolic subgroup of generated by the simple reflections , then one obtains a cell decomposition by using the projections of the cells where and is a minimal-length coset representative of . (We note moreover that Rietsch’s results hold for a semisimple, simply connected linear algebraic group over split over ).
In the case of the Grassmannian, Postnikov studied the Plücker nonnegative part of the Grassmannian, and gave a decomposition of it into positroid cells by intersecting with the matroid strata [Pos]. Concretely, if is the collection of bases of an element of , then . This cell decomposition of agrees with Rietsh’s cell decomposition [TW13, Corollary 1.2].
2.2. Background on (generalized) Bruhat interval polytopes
Bruhat interval polytopes were defined in [KW15], motivated by the connections to the full Kostant-Toda hierarchy.
Definition 2.4 ([KW15]).
Given two permutations and in with in Bruhat order, the Bruhat interval polytope is defined as
(2) |
We also define the (twisted) Bruhat interval polytope by
(3) |
While the definition of Bruhat interval polytope in (2) is more natural from a combinatorial point of view, as we’ll see shortly, the definition in (3) is more natural from the point of view of the moment map. Note that the set of Bruhat interval polytopes is the same as the set of twisted Bruhat interval polytopes; it is just a difference in labeling.
Remark 2.5.
If we choose any point in the cell (thought of as an matrix), and let be the matroid represented by the first rows of , then is the Minkowski sum of the matroid polytopes [KW15, Corollary 6.11]. In particular, is the matroid polytope of the flag matroid .
Following [TW15], we can generalize the notion of Bruhat interval polytope as follows (see [TW15, Section 7.2] for notation).
Definition 2.6.
Choose a generalized partial flag variety , let be the associated parabolic subgroup of the Weyl group , and let with in Bruhat order and a minimal-length coset representative of . Let denote the projection from to , and let be an element of the cell of (Lusztig’s definition of) .
A generalized Bruhat interval polytope can be defined in any of the following equivalent ways [TW15, Definition 7.8, Lemma 7.9, Proposition 7.10, Remark 7.11] and [BGW03, Preface]:
-
•
the moment map image of the closure of the torus orbit of in (which is a Coxeter matroid polytope)
-
•
the moment map image of the closure of the cell
-
•
the moment map image of the closure of the projected Richardson variety
-
•
the convex hull
where is the sum of fundamental weights , and is the dual of the real part of the Lie algebra of the torus .
Remark 2.7.
When with fundamental weights , each generalized Bruhat interval polytope is the flag positroid polytope associated to a matrix representing a point of , with . In this case the generalized Bruhat interval polytope is precisely the Minkowski sum of the matroid polytopes , where is the matroid realized by the first rows of . In particular, the generalized Bruhat interval polytope is the Minkowski sum , where is the positroid realized by the first rows of any matrix representing a point of . We will discuss how to read off the matroids from in Section 7.2.
As mentioned in the introduction, when is a sequence of consecutive ranks, the generalized Bruhat interval polytopes for are precisely the flag positroid polytopes of ranks . When , we recover the notion of Bruhat interval polytope, and when is a single integer, we recover the notion of positroid polytope.
3. The nonnegative tropicalization
3.1. Background on tropical geometry
We define the main objects in (a) and (b) of A, and record some basic properties. For a more comprehensive treatment of tropicalizations and positive-tropicalizations, we refer to [MS15, Ch. 6] and [SW05], respectively.
For a point , we write for its image in the tropical projective space . For , write .
Definition 3.1.
For a real homogeneous polynomial
the extended tropical hypersurface and the nonnegative tropical hypersurface are subsets of the tropical projective space defined by
and | |||
We say that a point satisfies the tropical relation of if it is in , and that it satisfies the positive-tropical relation of if it is in .
When is a multihomogeneous real polynomial, we define and similarly as subsets of a product of tropical projective spaces. We will consider tropical hypersurfaces of polynomials that define the Plücker embedding of a partial flag variety.
Definition 3.2.
For integers , the (single-exchange) Plücker relations of type are polynomials in variables defined as
where . When , the elements of are called the Grassmann-Plücker relations (of type ), and when , the elements of are called the incidence-Plücker relations (of type ).
As in the introduction, let be a sequence of increasing integers in . We let and let be the ideal generated by the elements of . It is well-known that for any field the ideal set-theoretically carves out the partial flag variety embedded in via the standard Plücker embedding [Ful97, §9]. Similarly, the Plücker relations define the tropical analogues of partial flag varieties as follows.
Definition 3.3.
The tropicalization of , the nonnegative tropicalization of , the flag Dressian , and the nonnegative flag Dressian are subsets of defined as
When , i.e. when consists of one integer , one obtains the (nonnegative) tropicalization of the Grassmannian and the (nonnegative) Dressian studied in [SS04, SW05, SW21, AHLS20]. Like , we write only in the subscript when .
Remark 3.4.
In [JLLO, §6], the authors define the “positive flag Dressian” to consist of the elements whose constituents are each in the strictly positive Dressian. In our language, this is equal to considering the points of
that have no coordinates. In a similar vein, we could consider defining the “nonnegative flag Dressian” to be the elements of the flag Dressian whose constituents are in the nonnegative Dressian. This gives a strictly larger set than our definition of the nonnegative flag Dressian, and has the shortcoming that the equivalence of (a) and (b) in A would no longer hold. See 4.4.
We record a useful equivalent description of the (nonnegative) tropicalization of a partial flag variety using Puiseux series. Recall the notion of the tropical semifield from 1.3.
Definition 3.5.
Let be the field of Puiseux series with coefficients in , with the usual valuation map . Concretely, for , is the exponent of the initial term of , and . Let
For a point , applying the valuation coordinate-wise to the Plücker coordinates gives a point . Noting that , we say that a point in has rational coordinates if it is a point in . Let be the subset of consisting of points with all coordinates in , i.e. the points that have a representative in .
Proposition 3.6.
The set equals the set of points in with rational coordinates. Likewise, the set equals the set of points in with rational coordinates. Moreover, we have
Proof.
Remark 3.7.
Let us also record an equivalent description of the (nonnegative) flag Dressian when is a sequence of consecutive integers. We need the following definition. As is customary in matroid theory, we write for the union of subsets and of .
Definition 3.8.
The set of three-term Grassmann-Plücker relations (of type ) is the subset of consisting of polynomials of the form
for a subset of cardinality and a subset disjoint from . Similarly, the set of three-term incidence-Plücker relations (of type ) is the subset of consisting of polynomials of the form
for a subset of cardinality and a subset disjoint from .
Let be the union of the three-term Grassmann-Plücker and three-term incidence-Plücker relations, which we refer to as the three-term Plücker relations.
Proposition 3.9.
Suppose consists of consecutive integers. Then a point is in the (nonnegative) flag Dressian if and only if its support is a flag matroid and satisfies the (nonnegative-)tropical three-term Plücker relations. More explicitly, we have
Proof.
We will use the language and results from the study of matroids over hyperfields. See [BB19] for hyperfields and relation to matroid theory, and see [Gun19, §2.3] for a description of the signed tropical hyperfield , for which we note the following fact: The underlying set of is , so given , one can identify it with the element of if and otherwise.
In the language of hyperfields, for a homogeneous polynomial in variables and a hyperfield , one has the notion of the “hypersurface of over ,” which is a subset of . When is the tropical hyperfield , this coincides with in 3.1. When is the signed tropical hyperfield , a point , when considered as a point of , is in if and only if it is in . Thus, in the language of flag matroids over hyperfields [JL22], the flag Dressian is the partial flag variety over , and the nonnegative flag Dressian is the subset of the partial flag variety over consisting of points that come from .
Now, both the tropical hyperfield and the signed tropical hyperfield are perfect hyperfields because they are doubly distributive [BB19, Corollary 3.45]. Our proposition then follows from [JL22, Theorem 2.16 & Corollary 2.24], which together state the following: When consists of consecutive integers, for a perfect hyperfield , a point is in the partial flag variety over if and only if the support of is a flag matroid and satisfies the three-term Plücker relations over . ∎
For completeness, we include the proof of the following fact.
Lemma 3.10.
The signed tropical hyperfield is doubly distributive. That is, for any , one has an equality of sets .
Proof.
If any one of the four is , then the desired equality is the usual distributivity of the signed tropical hyperfield. Thus, we now assume that all four elements are in , and write and similarly for . If , then and , so the equality follows again from the usual distributivity. So we now assume that all four elements have the same value in , and the equality then follows from the fact that the signed hyperfield is doubly distributive. ∎
Remark 3.11.
Even when does not consist of consecutive integers, [JL22, Theorem 2.16] implies that the flag Dressian and the nonnegative flag Dressian are carved out by fewer polynomials than in the following way: Denoting by
one has
This generalizes the fact that a sequence of matroids is a flag matroid if and only if is a flag matroid for all [BGW03, Theorem 1.7.1, Theorem 1.11.1].
The following corollary of 3.9 is often useful in computation. It states that the nonnegative tropical flag Dressian is in some sense “convex” inside the tropical flag Dressian.
Corollary 3.12.
Suppose consists of consecutive integers, and suppose we have points that are in . Then, if a nonnegative linear combination is in , it is in .
Proof.
We make the following general observation: Suppose is a three-term polynomial in with positive. Then an element satisfies the positive-tropical relation of if and only if . Hence, if each satisfy this relation, then a nonnegative linear combination of them can satisfy the tropical relation of only if the term at achieves the minimum, that is, only if the positive-tropical relation is satisfied. The corollary now follows from this general observation and 3.9. ∎
3.2. Equivalence of (a) and (b) in A
Let be a sequence of consecutive integers for some . We will show that . The inclusion is immediate from 3.3. We will deduce by utilizing the two known cases of the equality — when and when .
We start by recalling that tropicalization behaves well on subtraction-free rational maps.
Definition 3.13.
Let be a real polynomial, where is a finite subset of and . We define the tropicalization to be the piecewise-linear map , where as before, .
Note that Moreover, if and are two polynomials with positive coefficients, and , then These facts imply the following simple lemma, which appears as [RW19, Lemma 11.5]. See [SW05, Proposition 2.5] and [PS04] for closely related statements.
Lemma 3.14.
Let be a rational map defined by polynomials with positive coefficients (or more generally by subtraction-free rational expressions). Let , such that . Then
The next result states that we can extend points in the nonnegative Dressian to points in the nonnegative two-step flag Dressian.
Proposition 3.15.
Given with rational coordinates, there exists such that . Similarly, there exists such that .
The proof of 3.15 requires the following refined results about Rietsch’s cell decomposition of the nonnegative flag variety.
Theorem 3.16.
The nonnegative flag variety has a cell decomposition into positive Richardsons
where each cell can be parameterized using a map
Moreover, this parameterization can be expressed as an embedding into projective space (e.g. using the flag minors) using polynomials in the parameters with positive coefficients.
Proof.
Corollary 3.17.
Each -dimensional positroid cell in the nonnegative Grassmannian is the projection of some positive Richardson of dimension in , so we get a subtraction-free rational map
Proof.
That fact that each positroid cell is the projection of a positive Richardson was discussed in Section 2.1. The result now follows from 3.16. ∎
Proof of 3.15.
Using [AHLS20, Theorem 9.2], the fact that with rational coordinates implies that for some subspace , and hence lies in some positroid cell over Puiseux series.
By 3.17, is the projection of a point of , which in turn is the image of a point , and the Plücker coordinates of each are expressed as positive polynomials in the parameters .
In particular, we have subtraction-free maps
taking
The fact that the maps and are subtraction-free implies by 3.14 that we can tropicalize them, obtaining maps
taking
We now let and . By construction we have that all the three-term (incidence) Plücker relations hold for , and similarly for . Therefore and . ∎
The following consequence of 3.15 is very useful.
Corollary 3.18.
Let be positive integers, and let and be sequences of consecutive integers. Then any point with rational coordinates can be extended to a point .
Proof.
We start with . We take and repeatedly use 3.15 to construct , then . Similarly we take and use 3.15 to construct . Now by construction satisfies:
-
•
for ;
-
•
all three-term incidence-Plücker relations hold (because the three-term incidence-Plücker relations occur only in consecutive ranks).
Therefore by 3.9. ∎
Theorem 3.19.
Let be a sequence of consecutive integers, and let with rational coordinates. Then .
Proof.
Remark 3.20.
Remark 3.21.
Recall from 2.3 that if is a sequence of consecutive integers, the two notions of the positive/nonnegative part of the flag variety (see 2.2) coincide. The method used to prove the equivalence of (a) and (b) in A can be applied in a non-tropical context to prove 2.3 in an alternate way. We start by noting that the result holds when , which is to say, for the nonnegative Grassmannian [TW13, Corollary 1.2] and also when , which is to say, for the nonnegative complete flag variety [Bor22, Theorem 5.21]. To prove the result for , we start with a flag in ranks whose Plücker coordinates are all nonnegative, so that is Plücker nonnegative. As in 3.15, we can use the case to argue that the flag can be extended to lower ranks in such a way that all the Plücker coordinates are nonnegative. Dually, we can extend to higher ranks from the case. This yields a complete flag with all nonnegative Plücker coordinates. We can then apply the result in the complete flag case to conclude that lies in . Thus, is a projection of the nonnegative complete flag and itself lies in , which is to say, is Lusztig nonnegative.
The strictly positive tropicalization of a partial flag variety is the subset of consisting of points whose coordinates are never . Define similarly the strictly positive flag Dressian . The weaker version of 3.19 stating that was established in [JLLO, Lemma 19] as follows. One starts by noting that if , then the sequence of minors where is a point in . Then, the crucial step is a construction in discrete convex analysis [MS18, Proposition 2] that shows that every element of arises from an element of in this way. One then appeals to established in [SW21].
3.22 shows that the above argument does not work if one replaces “strictly positive” with “nonnegative.” In particular, the crucial step fails: that is, not every element of arises from an element of in such a way.
Example 3.22.
Let be matroids on whose sets of bases are . The matrix
shows that it is a flag positroid. However, we claim that there is no positroid of rank on such that , , and . Since all three cases involve deletion by 4, if we replace by , and decrease each of by , then we are claiming that there is no positroid of rank 3 on such that
(4) |
From and , we have that has bases , and similarly, we have has bases . Hence, the set of bases of contains , and does not contain . By considering the Plücker relation
we see that no positroid satisfies these properties.
4. Positively oriented flag matroids
In this section we explain the relationship between the nonnegative flag Dressian and positively oriented flag matroids, and we apply our previous results to flag matroids. In particular, we prove 1.5, which says that every positively oriented flag matroid of consecutive ranks is realizable. We also prove 4.8, which says that a positively oriented flag matroid of consecutive ranks can be extended to ranks (for ).
4.1. Oriented matroids and flag matroids
We give here a brief review of oriented matroids in terms of Plücker relations. Let be the hyperfield of signs. For a polynomial , we say that an element is in the null set of if the set is either or contains .
Definition 4.1.
An oriented matroid of rank on is a point , called a chirotope, such that is in the null set of for every . Similarly, an oriented flag matroid of ranks is a point such that is in the null set of for every .
While these definitions may seem different from those in the standard reference [BLVS+99] on oriented matroids, 4.1 is equivalent to [BLVS+99, Definition 3.5.3] by [BB19, Example 3.33]. The definition of oriented flag matroid here is equivalent to the definition of a sequence of oriented matroid quotients [BLVS+99, Definition 7.7.2] by [JL22, Example above Theorem D].
Definition 4.2.
A positively oriented matroid is an oriented matroid such that only takes values 0 or 1. Similarly, we define a positively oriented flag matroid to be an oriented flag matroid such that only takes values 0 or 1.
A positroid defines a positively oriented matroid where takes value 1 on its bases and 0 otherwise. In 1987, da Silva [dS87] conjectured that every positively oriented matroid arises in this way; this conjecture was subsequently proved in [ARW17] and then [SW21].
Theorem 4.3.
[ARW17] Every positively oriented matroid is realizable, i.e. has the form for some positroid .
By 4.3, each positively oriented flag matroid is a sequence of positroids which is also an oriented flag matroid.
In this section we will prove 1.5, which generalizes 4.3, and says that every positively oriented flag matroid of consecutive ranks can be realized by a flag positroid. But before we prove it, let us give an example that shows that imposing the oriented flag matroid condition is stronger than imposing that we have a realizable flag matroid whose consistent matroids are positroids.
Example 4.4.
We give an example of a realizable flag matroid that has positroids as its constituent matroids but is not a flag positroid. This example also appeared in [JLLO, Example 5] and [BK22, Example 6]. Let be matroids of ranks 1 and 2 on whose sets of bases are and , respectively. Both are positroids. We can realize as a flag matroid using the matrix
where the nonvanishing minors are nonzero. In order to realize as a flag positroid, we need to choose real numbers such that all these minors are strictly positive. However, and implies , while and implies .
This example is consistent with 1.5 because , when considered as a sequence of positively oriented matroids, is not an oriented flag matroid.
4.2. From the nonnegative flag Dressian to positively oriented flag matroids
We start with the following simple observation. While the proof is very simple, we label it a “theorem” to emphasize its importance.
Theorem 4.5.
The set of positively oriented flag matroids of ranks can be identified with the set of points of the nonnegative flag Dressian whose coordinates are all either 0 or .
Proof.
Given a point ,222Note that is not a sequence of chirotopes in this proof, instead each we define by setting if and if . Then, we observe that is in the null set of a polynomial if and only if the image of in is a point in . Therefore, each positively oriented flag matroid can be identified with the element in the nonnegative flag Dressian . ∎
We now prove that every positively oriented flag matroid of consecutive ranks is realizable.
Proof of 1.5.
By the lemma, we may identify a positively oriented flag matroid as an element of the nonnegative flag Dressian. Because the ranks are consecutive integers, the equivalence (a)(b) of A implies that is thus a point in . Because has rational coordinates (all non- coordinates are 0), 3.6 implies that for some . Setting the parameter in each Puisseux series of to 0 now gives the realization of as a flag positroid. ∎
As in 1.6, we do not know whether the corollary holds when does not consist of consecutive integers. The following example shows that one cannot reduce to the consecutive ranks case.
Example 4.6.
We give an example of a flag positroid on of ranks such that there is no flag positroid with rank of equal to 2. Let the sets of bases of and be and , respectively. The matrix
for example shows that is a flag positroid. However, this flag positroid cannot be extended to a flag positroid with consecutive ranks. To see this, note that any realization of as a flag positroid, after row-reducing by the first row, is of the form
where and . The minors of the matrix formed by the first two rows include , which cannot be all nonnegative since and not both of and are zero.
Remark 4.7.
Let us sketch an alternate proof of 1.5 that relies only on the weaker version of (a)(b) in A that the strictly positive parts agree, i.e. that . For a matroid of rank , define by for , where is the rank function of . If is a positively oriented matroid, then is a point in the positive Dressian [SW21, Proof of Theorem 5.1]. One can use this to show that if is a positively oriented flag matroid of consecutive ranks , then the sequence is a point in . Since and has rational coordinates, 3.6 implies that there is a point with . Consider the coordinate of at a subset . By construction, the initial term of is for some positive real and a nonnegative integer , where is zero exactly when is a basis of . Thus, setting the parameter to in the Puisseux series of gives a realization of as a flag positroid.
Corollary 4.8.
Let be positive integers, and let be a positively oriented flag matroid on of consecutive ranks , that is, a sequence of positroids which is also an oriented flag matroid. Then we can extend it to a positively oriented flag matroid of consecutive ranks .
Proof.
As in 4.5, we view the positively oriented flag matroid as a point of the nonnegative flag Dressian whose coordinates are all either 0 or . The desired statement almost follows from 3.15: we just need to check that we can extend in a way which preserves the fact that coordinates are all either or . This is true, and we prove it by following the proof of 3.15 and replacing all instances of the positive Puiseux series by the positive Puiseux series with constant coefficients, that is, by . Alternatively, we can use our result that is realizable by a flag positroid, and then argue as in 3.21. ∎
5. Subdivisions of flag matroid polytopes
5.1. Flag Dressian and flag matroidal subdivisions
Consider a point such that its support is a flag matroid. By construction, the vertices of the flag matroid polytope have the form where is a basis of the matroid for each .
Definition 5.1.
We define to be the coherent subdivision of induced by assigning each vertex of the weight . That is, the faces of correspond to the faces of the lower convex hull of the set of points
The points of the flag Dressians are exactly the ones for which the subdivision consists of flag matroid polytopes.
Theorem 5.2.
[BEZ21, Theorem A.(a)&(c)] A point is in the flag Dressian if and only if the all faces of the subdivision are flag matroid polytopes.
When consists of consecutive integers , the nonnegative analogue of this theorem is the equivalence of (b) and (c) in A, which states that a point is in the nonnegative flag Dressian if and only if all faces of the subdivision are flag positroid polytopes. A different nonnegative analogue of 5.2 that holds for not necessarily consecutive, but loses the flag positroid property, can be found in 5.6.
5.2. The proof of (b)(c)(d)(b) in A
We start by recording two observations. The first is a well-known consequence of the greedy algorithm for matroids; see for instance [AK06, Proposition 4.3]. For a matroid on and a vector , let be the face of the matroid polytope that maximizes the standard pairing with .
Proposition 5.3.
Let be a matroid on and let be a chain of nonempty proper subsets of . For a vector in the relative interior of the cone , we have
where is the direct sum of minors of .
For a flag matroid, since is the Minkowski sum , we likewise have that , where . In particular, the face of a flag matroid polytope is a flag matroid polytope.
The second observation concerns the following operations that we will show preserve the nonnegative flag Dressian. Recall that for , its support is .
-
•
We consider a point as a set of weights on the vertices of . Given an affine-linear function and an element , we define
-
•
For a point , denote by its initial part, i.e.
Proposition 5.4.
Let be a sequence of increasing integers in . Suppose . Then, the following hold.
-
(1)
The support is a positively oriented flag matroid. In particular, it is a flag positroid when consists of consecutive integers.
-
(2)
We have for any affine-linear functional on .
-
(3)
We have .
Proof.
We may consider as an element by assigning the value 0 to a subset if it is in the support of and otherwise. Then, we have because the terms in each of the tropical Plücker relations that achieve the minimum when evaluated at continue to do so when evaluated at . The statement (1) follows from 4.5 and 1.5
The support is unchanged by , so is a flag matroid. The statement (2) now follows because for each of the positive-tropical Plücker relations, the operation preserves the terms at which the minimum is achieved.
The support is a flag matroid by 5.2 and because is a face in the subdivision of . The statement (3) now follows because for each of the positive-tropical Plücker relations, the operation in either preserves the terms at which the minimum is achieved or changes all the terms involved to . ∎
Remark 5.5.
Proof of (b)(c).
Every face in the coherent subdivision is the initial one after an affine-linear transformation. Hence, the implication follows from 5.4. ∎
Remark 5.6.
One may modify the statement (c) to the following:
-
(c’)
Every face in the coherent subdivision of is the flag matroid polytope of a positively oriented flag matroid.
Similar argument as above shows that (b)(c’) even when doesn’t consist of consecutive integers. One can also verify the converse (c’)(b) in this more general case as follows:
Suppose for contradiction (c’) but not (b) for some . Then 5.2 implies that is in the flag Dressian, and thus the failure of (b) implies that there is a Plücker relation where the minimum occurs at least twice but at the terms whose coefficients have the same sign. 5.4 implies that, replacing by for some if necessary, we may conclude that the same is true for that Plücker relation evaluated at . But then , which arise as a face in the subdivision, is not a positively oriented flag matroid by 4.5, contradicting (c’).
There is no equivalence of (c’) and (e) since three-term incidence relations exist only for consecutive ranks.
Proof of (d)(b).
First, assumption (d) implies that every edge of the subdivision of is a flag matroid polytope, i.e. it is parallel to for some and its two vertices are equidistant from the origin. Hence the edges of have the same property, so is a flag matroid. By Proposition 3.9, to show (b) it now suffices to show that every positive-tropical three-term Plücker relation is satisfied.
We start with the case , where is just . We need check the validity of the three-term positive-tropical Grassmann-Plücker relations, say for an arbitrary choice of and . If is not independent in the matroid , then every term in the three-term relation involving and is , so we may assume is independent. Let be a maximal chain of subsets of with the property that and . Then, Proposition 5.3 implies that for a vector in the relative interior of the cone , we have
For the second identification, we have used that
-
(1)
the matroid polytope of a direct sum of matroids is the product of the matroid polytopes;
-
(2)
with the exception of , all other minors of the matroid corresponding to in the chain have their polytopes being a point because .
Since is assumed to be independent, the rank of the matroid minor is at most . If it is less than 2, then every term in the three-term relation involving and is , so let us now treat the case when the rank is exactly 2. For a basis of , let be the basis of such that the vertex of corresponds to the vertex of under the identification above. Identifying with , we may thus consider “restricting” to the face to obtain an element defined by
It is straightforward to check that for , the three-term positive-tropical Grassmann-Plücker relations are satisfied if and only if all 2-dimensional faces in the corresponding subdivision are positroid polytopes. Since the faces of the subdivision of are a subset of the faces of the subdivision , we have that satisfies the three-term tropical-positive Grassmann-Plücker relation involving and .
Let us now treat the case . That the three-term Grassmann-Plücker relations are satisfied for every where follows from our previous case of once we show the following claim:
For a flag matroid with consecutive rank sequence , if every face of of dimension at most 2 is a flag positroid polytope, then the same holds for every constituent matroid, i.e. for every , every face of of dimension at most 2 is a positroid polytope .
To prove the claim, suppose for some that a 2-dimensional face of is not a positroid polytope. Our goal is to use to find a -dimensional face of that is not a flag positroid polytope. By [LPW20, Theorem 3.9], a 2-dimensional matroid polytope which is not a positroid polytope has vertices of the form , where with and ; thus for such 333One may also deduce this independently of [LPW20] by using the argument given in the first third of this proof of (d)(b) concerning the case.. Note that this 2-face is the Minkowski sum of with the product .
Let be a maximal chain of subsets of with the property that , , and . Then, Proposition 5.3 implies that for a vector in the relative interior of the cone , we have
For the second identification, we have used that
-
(1)
the matroid polytope of a direct sum of matroids is the product of the matroid polytopes;
-
(2)
with the exception of and , all other minors of the constituent matroids of corresponding to in the chain have their polytopes being a point because .
Note that the polytope is at most 2-dimensional since and are flag matroids on ground sets and , respectively. The polytope has as a Minkowski summand, and thus in particular is not a flag positroid polytope.
Lastly, we check the validity of the three-term positive-tropical incidence-Plücker relations, say for an arbitrary choice of with and . We may assume that has rank in the matroid , since otherwise every term in the three-term positive-tropical incidence relation is , so that the relation is vacuously satisfied. Let be a maximal chain of subsets of with the property that and for . Then, Proposition 5.3 implies that for a vector in the relative interior of the cone , we have
For the second identification, we have used that
-
(1)
the matroid polytope of a direct sum of matroids is the product of the matroid polytopes;
-
(2)
with the exception of , all other minors of the constituent matroids of corresponding to in the chain have their polytopes being a point because .
Note that the polytope is at most 2-dimensional since it is a flag matroid polytope on 3 elements. Similarly to the case, we may “restrict” to the face to obtain an element . We may assume that since otherwise every term in the three-term incidence relation of the pair is . For , it is straightforward to verify that the unique three-term positive-tropical incidence relation involving and is satisfied if and only if the subdivision consists only of flag positroid polytopes. Since the faces of the subdivision are a subset of the faces of the subdivision , we have that satisfies the three-term incidence relation involving and . ∎
6. Three-term incidence relations
6.1. The proof of (e)(b) in A
To prove the implication when , we will show the following key theorem.
Theorem 6.1.
Suppose satisfies every three-term positive-tropical incidence relation, and suppose that the support is a flag matroid. Then, we have if either of the following (incomparable) conditions hold:
-
(i)
The support consists of uniform matroids.
-
(ii)
Either or .
Proof of (b)(e).
By Proposition 3.9, the implication (b)(e) is immediate. For the converse, since consists of consecutive integers, if is a flag matroid and satisfies every three-term positive-tropical incidence relation, then also satisfies every three-term positive-tropical Grassmann-Plücker relation if either of the conditions (i) or (ii) of Theorem 6.1 is satisfied. The hypothesis of (e) satisfies this, so is an element of by 3.9. ∎
The proof of 6.1 relies on the following technical lemma.
Lemma 6.2.
Suppose satisfies all three-term positive-tropical Grassmann-Plücker relations involving the element 5. Suppose moreover that for some . Then , i.e. also satisfies the three-term positive-tropical Grassmann-Plücker relation not involving 5.
Proof.
The idea of the proof of 6.2 is that in the usual Grassmannian , if we can invert certain Plücker coordinates, then we can write the three-term Grassmann-Plücker relation not involving as a linear combination of three of the other three-term Grassmann-Plücker relations. In particular, we have the following identity, which is easy to verify.
Lemma 6.3.
If (respectively, ) then can be written in the following ways.
We next note that we can interpret the first (respectively, second) expression in 6.3 tropically as long as (respectively, ).
Case 1: Then we can make sense of the terms on the right hand side of the first expression of 6.3 tropically. Since the three-term positive tropical Plücker relations involving hold, and , we have
We now simplify these expressions and underline terms that agree, obtaining:
(5) | ||||
(6) | ||||
(7) |
There are now eight cases to consider, based on whether the minimum is achieved by the first or second term in each of (5), (6), (7). All cases are straightforward. If the minimum is achieved by the first term in (5) and the second term in (7), then we find that . In the other six cases, we find that . Therefore the positive tropical Plücker relation involving is satisfied.
Case 2: The argument for Case 2 is the same as for Case 1, except we use the tropicalization of the second identity in 6.3.
Case 3: In this case, since is not a loop, either or Suppose that . Then the positive tropical Plücker relations
-
•
-
•
-
•
imply that , and hence the positive tropical Plücker relation involving is satisfied. The case where is similar. ∎
For , define its dual by . It is straightforward to verify that is an element of (resp. ) if and only if is an element of (resp. ). This matroid duality gives the following dual formulation of 6.2.
Corollary 6.4.
Suppose satisfies all three-term positive-tropical Grassmann-Plücker relations that contain a variable indexed by with . If is a matroid such that 5 is not a coloop, then , i.e. also satisfies the three-term positive-tropical Grassmann-Plücker relation whose every variable contains 5 in its indexing subset.
We are now ready to prove 6.1. We expect that the proof of 6.1 here adapts well to give an analogous statement for arbitrary perfect hyperfields.
Proof of 6.1.
Given such , define by
Because is a flag matroid, we have that is a matroid, with the element that is neither a loop nor a coloop. We observe that if and only if because the validity of the three-term positive-tropical Grassmann-Plücker relations for is equivalent to the validity of both the three-term positive-tropical incidence relations and the three-term positive-tropical Grassmann-Plücker relations for .
We need to check that satisfies every three-term positive-tropical Grassmann-Plücker relation of type . Consider the three-term relation associated to the subset of cardinality and disjoint from . We have three cases:
-
•
. In this case, erasing the index in the expression for the corresponding three-term Grassmann-Plücker relation yields a three-term incidence relation of type , which is satisfied by our assumption on .
-
•
. In this case, if is not a coloop in the minor , then applying 6.4 to implies that the three-term Grassmann-Plücker relation is satisfied.
-
•
. In this case, if is not a loop in the minor , then applying 6.2 to implies that the three-term Grassmann-Plücker relation is satisfied.
Under condition (i) of Theorem 6.1, i.e. when the support consists of uniform matroids, the element is not a coloop in the minor , and is not a loop in the minor . Hence, both Corollary 6.4 and Lemma 6.2 apply respectively, and we conclude that in every case the three-term positive-tropical Grassmann-Plücker relation is satisfied.
Now suppose condition (ii) of Theorem 6.1 holds. We verify that in the cases where Corollary 6.4 or Lemma 6.2 do not apply, the relevant positive-tropical Grassmann-Plücker relation is satisfied. Let us consider the third bullet point, and suppose that is a loop in the minor , i.e. where Lemma 6.2 does not apply; the argument for the second bullet point is similar by matroid duality. In this case, since is not a loop in the matroid , belongs to the closure (also called span) in of . Since is also independent, there is an element such that is independent and has the same closure as in . Let . For any , by our choice of , we have that is a basis of if and only if is a basis of . Moreover, for any such that the values involved below are finite, we claim
Note that using the definition of , the above claim can be equivalently written as
From the claim, we conclude as follows. Let be the projection of to the coordinates labelled by where , and let be the projection of to the coordinates labelled by where . Then, as elements of , the two tropical vectors and are equal. Hence, the claim implies that if one of or satisfies the three-term Grassmann-Plücker relations on these coordinates, then so does the other.
The claim follows from the validity of three-term tropical incidence relations, which is implied by the validity of three-term positive-tropical incidence relations. Namely, we have that the minimum is achieved at least twice in
from which the claim follows because is not a basis of , forcing . ∎
7. Projections of positive Richardsons to positroids
One recurrent theme in our paper has been the utility of projecting a complete flag positroid (equivalently, a positive Richardson) to a positroid (or a positroid cell). This has come up in Rietsch’s cell decomposition of a nonnegative (partial) flag variety, in our proofs in Section 3.2, and in the expression of a Bruhat interval polytope as a Minkowski sum of positroid polytopes in 2.7. Positive Richardsons can be indexed by pairs of permutations with . Meanwhile, by work of Postnikov [Pos], positroid cells of can be indexed by Grassmann necklaces. In this section we will give several concrete combinatorial recipes for constructing the positroids obtained by projecting a (complete) flag positroid. We will also discuss the problem of determining when a collection of positroids can be identified with a (complete) flag positroid.
7.1. Indexing sets for cells of
As discussed in Section 2.1, there are two equivalent ways of thinking about the positroid cell decomposition of :
In the union on the right, is the projection from to , and range over all permutations in , such that is a minimal-length coset representative of , and We write for the set of minimal-length coset representatives of . Recall that a descent of a permutation is a position such that . We have that is the subset of permutations in which have at most one descent, and if it exists, that descent must be in position .
Even if , the projection of to is still a positroid, which we will characterize below. We start by defining Grassmann necklaces [Pos].
Definition 7.1.
Let be a sequence of subsets of . We say is a Grassmann necklace of type if the following holds:
-
•
If , then for some .
-
•
If , then .
In order to define the bijection between these Grassmann necklaces and positroids, we need to define the -Gale order on .
Definition 7.2.
We write for the following shifted linear order on .
We also define the -Gale order on -element subsets by setting
if and only if for all .
Given a positroid , we define a sequence of subsets of by letting be the minimal basis of in the -Gale order. The following result is from [Pos, Theorem 17.1].
Proposition 7.3.
For any positroid , is a Grassmann necklace. The map gives a bijection between positroids of rank on and Grassmann necklaces of type .
7.2. Projecting positive Richardsons to positroids
In this section we will give several descriptions of the constituent positroids appearing in a complete flag positroid (that is, a flag matroid represented by a positive Richardson). We start by reviewing a cryptomorphic definition of flag matroid, based on [BGW03, Sections 1.7-1.11].
A flag on is an increasing sequence of finite subsets of . A flag matroid is a collection of flags satisfying the Maximality Property. Recall that denotes the indicator vector in associated to a subset . For a flag we let . In this language, the flag matroid polytope of is , whose vertices are precisely the points for .
In the complete flag case, each point is a permutation vector for some . Note that we can read off from by setting , where is the unique element of .
Given in Bruhat order, we define the Bruhat interval flag matroid to be the complete flag matroid whose flags are precisely
where denotes and denotes . Then by the above discussion, the (twisted) Bruhat interval polytope
is the flag matroid polytope of the Bruhat interval flag matroid .
This observation leads naturally to the following definition.
Definition 7.4.
Consider a complete flag matroid on , which we identify with a collection of permutations on . By the Maximality Property [BGW03, Section 1.7.2] and its relation to the tableau criterion for Bruhat order [BGW03, Theorem 5.17.3], contains a unique permutation (respectively, ) which is minimal (respectively, maximal) in Bruhat order among all elements of . We say that is the Bruhat interval envelope of .
It follows from 7.4 that the Bruhat interval envelope of a complete flag matroid contains ; however, in general this inclusion is strict. It is an equality precisely when is a Bruhat interval flag matroid.
Recall that if and are flags, we say that is less than or equal to in the Gale order (and write ) if and only if for all . (We also talk about the “usual” Gale order with respect to the total order .) The Maximality Property for flag matroids implies that for any flag matroid , there is always a unique element which is maximal (and a unique element which is minimal) with respect to .
We now give a Grassmann necklace characterization of the positroid constituents of a complete flag positroid, which follows from the previous discussion plus 7.3.
Proposition 7.5.
Consider a complete flag positroid on , that is, the flag positroid associated to any point of , for some . For each , let be the Gale-minimal permutation with respect to in the interval . Then the Grassmann necklace of the positroid is .
Example 7.6.
Consider the flag positroid associated to a point of , where and (which we abbreviate as 1243 and 4213). The interval consists of
We now use 7.5, and find that the Gale-minimal permutations of with respect to are . Therefore the Grassmann necklaces for the constituent positroids and are , , , and .
Alternatively, we can read off the flags in the flag positroid from the permutations in , obtaining the flags
(Note that for brevity, we have omitted the subset 1234 from the end of each flag above.) We can now read off the bases of from the flags, obtaining , , , and . We can then directly calculate the Grassmann necklaces from these sets of bases, getting the same answer as above.
If we compute the Minkowski sum of the positroids above, we obtain the twisted Bruhat interval polytope , whose vertices are
as noted in 2.7.
The following result gives an alternative description of the constituent positroids of a complete flag matroid, this time in terms of bases.
Lemma 7.7 ([KW15, Lemma 3.11] and [BW22, Theorem 1.4]).
Consider a complete flag positroid, that is, a flag matroid represented by a point of a positive Richardson , where and in Bruhat order. Choose . Let denote the projection from to . Then the bases of the rank positroid represented by are
Finally, we remark that [BK22, Remark 5.24] gives yet another description of the constituent positroids of a complete flag positroid, this time in terms of pairs of permutations.
7.3. Characterizing when two adjacent-rank positroids form an oriented matroid quotient
We have discussed how to compute the projection of a complete flag positroid to a positroid. Moreover, it is well-known that every positroid is the projection of a complete flag positroid. In this section we will give a criterion for determining when two positroids and on of ranks and can be obtained as the projection of a complete flag positroid (see 7.14).
We recall the definition of oriented matroid quotient in the setting at hand.
Definition 7.8.
We say that two positroids and on of ranks and form an oriented matroid quotient if is an oriented flag matroid.
The following statement is a direct consequence of 4.8.
Proposition 7.9.
Let and be positroids on of ranks and . Then there is a complete flag positroid with and as constituents if and only if form an oriented matroid quotient.
Proposition 7.10.
Suppose that is a sequence of positroids of ranks on , such that each pair and forms an oriented matroid quotient. Then is a complete flag positroid. Moreover, it is realized by a point of the positive Richardson , where we can explicitly construct and as follows:
-
•
Let (respectively, ) be the bases of which are minimal (maximal) with respect to the usual Gale ordering. Then are defined by
Proof.
As in 4.5, we identify each positroid with the image of its chirotope ; we have that lies in . The fact that each pair forms an oriented matroid quotient means that satisfies all three-term incidence-Plücker relations, and hence . Since , we have proved that is a complete flag positroid.
As we’ve seen in 4.4 it is a subtle question to determine whether a pair of positroids and of ranks and form an oriented matroid quotient. One way is to construct an by matrix such that the minor in rows and columns is non-zero if and only if is a basis of while the maximal minor in rows and columns is non-zero if and only if is a basis of . Another way is to check the three-term relations over the signed tropical hyperfield, as in 3.9. We do not have an efficient way to do either of these things. Instead, in 7.14, we will give an algorithmic, combinatorial way to verify whether and form an oriented matroid quotient.
Construction # 1. Given two positroids and on the ground set of ranks and , respectively, which form a positively oriented matroid quotient, we construct a positroid of rank on the ground set where is neither a loop nor a coloop. The bases of are precisely
Construction #2. Conversely, given a rank positroid on ground set , where is neither a loop nor coloop, we construct two positroids and which form a positively oriented matroid quotient, as follows. Let be a matrix realizing ; therefore its Plücker coordinates are nonnegative. We apply row operations to rewrite in the form
Let denote the matroid on realized by and let denote the matroid on realized by together with the row of ’s below it. Then and are both positroids (since the Plücker coordinates of and are all nonnegative), and they form a positively oriented quotient. Moreover, it is clear that and .
The idea of our algorithm is to translate Constructions and into operations on Grassmann necklaces, so that Construction is well-defined even if and fail to form a positively oriented quotient. Clearly if we start with positroids and forming a positively oriented matroid quotient, then Construction followed by is the identity map. Conversely, if Construction followed by is the identity map, then since Construction always outputs a positively oriented matroid quotient, we must have started with positroids forming a positively oriented matroid quotient.
We let denote the minimum of the sets in the order.
Proposition 7.11.
Let and be positroids of consecutive ranks which form a positively oriented quotient. Let be the Grassmann necklace of for . Define
Then is the Grassmann necklace of the positroid on whose bases are precisely
Proof.
It suffices to show that each basis of is -Gale greater than for all . One also need to check that the are in fact bases of but this is clear by definition.
Note that the minimal flag of a flag matroid consists of the minimal bases of each of its constituent matroids [BGW03, Corollary 7.2.1]. Thus, for each .
First, let be a basis of . For , we have . Since neither nor contain , . By our earlier observation, for some . Thus, . We conclude that for all .
Next, consider for a basis of . For , we have . Since neither nor contain , we have and . Since , we have . We conclude that for all . ∎
If and form a positively oriented quotient, we should obtain them from the positroid , constructed as in 7.11, by deleting and contracting . The following result explains how these operations affect Grassmann necklaces.
Proposition 7.12.
[Oh08, Proposition 7 and Lemma 9] Let be a positroid on such that is neither a loop nor a coloop, with Grassmann necklace . Then the Grassmann necklaces and of and , are as follows:
Taken together, the last two results yield a recipe for verifying whether two positroids, given in terms of their Grassmann necklaces, form a positively oriented quotient. First apply the construction of 7.11. If that yields a Grassmann necklace, apply 7.12 and see if that yields the original Grassmann necklaces. If so, the two Grassmann necklaces form a positively oriented quotient.
Our next goal is to streamline this recipe. Let and be Grassmann necklaces of positroids of ranks and , respectively. Note that a necessary condition for the positroids corresponding to and forming a positively oriented quotient is that for all . Now, we define a subset as follows: For each , if , let . Since for some , this is as simple as checking whether . If the positroids corresponding to and form a positively oriented quotient, applying 7.11 and then 7.12 should leave them unchanged. It is straightforward to see that if and only if in 7.12. In particular, since is a Grassmann necklace, must either be an interval of the form , or empty.
Next we claim that, once we verify that is an interval of the form or is empty, then it follows automatically that , as constructed in 7.11, is a Grassmann necklace.
Lemma 7.13.
Let and be Grassmann necklaces of types and , respectively. Construct as in 7.11. Let . If for some or , then is a Grassmann necklace.
Proof.
It is clear from the definition that satisfies the Grassmann necklace condition for each pair of consecutive sets and except for when , and (where we label sets cyclically so that ).
If , then . This makes it clear that the Grassmann necklace condition holds for and . Also, sing the fact that for all , it is not hard to verify the Grassmann necklace condition for and .
This leaves us to check the condition for and . In this case, and . Our goal is to show that for some . It is immediately obvious that we necessarily have . Thus, we are left to show that , or that .
Let be defined by , let be defined by and let be defined by . We observe that . Also, . Comparing these two equalities, we conclude that either and , or and . The first case is what we want to prove, so let us show by contradiction that the second case cannot occur.
Assume and . By assumption, and . Thus, and . Since , this means they are both equal to . However, if , then has as a coloop. it follows that , which means as well. Thus, in this case, , as desired.
Finally, if , we can check that the Grassmann necklace condition holds for and as before. The we are just left to verify this condition for and . We can apply the same logic but with replaced by and replaced by . Specifically, we find . We then must show that it is impossible for and . However, . Moreover, it is always true that . Using , we then find and which means that and we can conclude as in the previous paragraph. ∎
Theorem 7.14.
Fix positroids and on of ranks and , respectively. Let and be their Grassmann necklaces. We now set , where denotes the Gale order on . Define and . Then and form a positively oriented quotient if and only if the following conditions hold:
-
(1)
For , .
-
(2)
is an interval of the form or .
-
(3)
For , .
-
(4)
For , .
Proof.
First, suppose that we have a positively oriented quotient. As explained earlier, the first two conditions always hold for positively oriented quotients. We know that applying the constructions of 7.11 and 7.12 in sequence should preserve our positively oriented quotient. Observing what conditions this imposes on the constituent Grassmann necklaces yields conditions and .
Conversely, if the conditions in the theorem statement hold, then by 7.13, applying the construction of 7.11 to and yields another Grassmann necklace on such that is neither a loop nor a coloop of the positroid corresponding to . Then, conditions and guarantee that applying the construction of 7.12 to will recover and . The result of applying 7.12 to the Grassmann necklace of a positroid with neither a loop nor a coloop is the pair of Grassmann necklaces corresponding to and , which form a positively oriented quotient. ∎
Example 7.15.
Let and . Then is an interval with upper endpoint . Note that , and , while , and . The positroids with these Grassmann necklaces do not form a positively oriented quotient since it is false that .
However, if we start with Grassmann necklaces and , then the values of the and are unchanged. It is straightforward to verify that the conditions of 7.14 hold and so the positroids corresponding to and do in fact form a positively oriented quotient.
We now have a tool that allows us to recognize flag positroids in consecutive ranks without finding a realization or certifying the incidence relations over the signed hyperfield.
Corollary 7.16.
Suppose is a sequence of positroids of ranks . Then is a flag positroid if and only if for , the pair of positroids satisfy the conditions of 7.14.
8. Fan structures for and coherent subdivisions from and
In this section we make some brief remarks about the various fan structures for and coherent subdivisions from points of . Codes written for computations here are available at https://github.com/chrisweur/PosTropFlagVar. We take a detailed look at the Grassmannian and complete flag variety, in particular the case of .
8.1. Fan structures
There are multiple possibly different natural fan structures for :
-
(i)
The Plücker fan (induced by the three-term tropical Plücker relations).
-
(ii)
The secondary fan (induced according to the coherent subdivision as in 8.3).
-
(iii)
The Gröbner fan (induced according to the initial ideal of the ideal ).
-
(iv)
The simultaneous refinement of the fans dual to the Newton polytopes of the Plücker coordinates, when the Plücker coordinates are expressed in terms of a “positive parameterization” of , such as an -cluster chart.
-
(v)
(If the cluster algebra associated to has finitely many cluster variables) the same fan as above but with (the larger set of) cluster variables replacing Plücker coordinates.
Note that by definition, fan (v) is always a refinement of (iv).
In the case of the positive tropical Grassmannian, the fan structures in (iv) and (v) were studied in [SW05, Definition 4.2 & Section 8], where the authors observed that for , fan (iv) (which coincides with (v)) is isomorphic to the cluster complex444See [FZ03] for background on the cluster complex. of type ; for and , fan (iv) is isomorphic to a coarsening of the corresponding cluster complex, while fan (v) is isomorphic to the cluster complex (of types and , respectively). [SW05, Conjecture 8.1] says that fan (v) (associated to the positive tropicalization of a full rank cluster variety of finite type) should be isomorphic to the corresponding cluster complex. This conjecture was essentially resolved in [JLS21, AHHL21] by working with -polynomials.
[OPS19, Theorem 14] states that the Plücker fan and the secondary fan structures for Dressians coincide, and hence implies that (i) and (ii) coincide because the positive Dressian and the positive tropical Grassmannian are the same [SW21]. For , the results of [SS04, §4] imply that (i), (ii), and (iii) agree, and combining this with [SW05, §5] implies that all five fan structures agree for . For , we computed that (iii) and (v) strictly refine (i), but the two fan structures are not comparable.
We can consider the same fan structures in the case of the positive tropical complete flag variety. When , the fan modulo its lineality space is a one-dimensional fan, and all fan structures coincide. For (before taking the positive part), one can find computations of the fan (iii) for and in [BLMM17, §3], the fan (i) and its relation to (iii) for in [BEZ21, Example 5.2.3], and the fan (ii) and its relation to (iii) for in [JLLO, §5]. Returning to the positive tropicalization, [Bos22, Section 5.1] computed the fan structure (iii) for , and found it was dual to the three-dimensional associahedron; in particular, there are maximal cones and the -vector is . Using the positive parameterization of [Bor22] (a graphical version of the parameterizations of [MR04]) for , we computed the polyhedral complex underlying (iv) for in Macaulay2 by computing the normal fan of the Minkowski sum of the Newton polytopes of the Plücker coordinates expressed in the chosen parametrization; we obtained the -vector . We also computed (v) after incorporating the additional non-Plücker cluster variable . Combining these, we find that for , (i)=(iv) and (ii)=(iii). We also find that both (ii) and (v) strictly refine (i)=(iv) and are both isomorphic to the normal fan of the three-dimensional associahedron, but are not comparable fan structures.
The fact that the fan structure (v) of is dual to the three-dimensional associahedron is consistent with [SW05, Conjecture 8.1] and the fact that has a cluster algebra structure of finite type [GLS08, Table 1], whose cluster complex is dual to the associahedron.
We now give a graphical way to think about the fan structure on , building on the ideas of [SW05] and [BEZ21, Example 5.2.3].
Example 8.1.
A planar tree on is an unrooted tree drawn in the plane with leaves labeled by (in counterclockwise order). By [SW05], parameterizes metric planar trees, and its cones correspond to the various combinatorial types of planar trees. In particular, if we assign real-valued lengths to the edges of a planar tree, then the negative of the distance between leaf and encodes the positive tropical Plücker coordinate of a point in the corresponding cone. In particular, it is easy to see that the negative distances associated to such a planar tree satisfy the positive tropical Plücker relations.
Now as in [BEZ21, Example 5.2.3], we note that for a valuated matroid whose underlying matroid is the uniform matroid , the tropical linear spaces and associated to and its dual are translates of each other. This allows us to identify points of with planar trees on the vertices such that the vertices and separately the vertices appear in counterclockwise order. To see this, note that (using the same idea as Construction #1 from Section 7) we can identify , with Plücker coordinates , with an element of : we simply set for . Similarly, we identify , where has Plücker coordinates , with an element of : we simply set , where .
This gives us the Plücker fan structure (i)=(iv) with thirteen maximal cones, as shown in Figure 2. To get the Gröbner fan structure (iii) we subdivide one of the cones into two, along the squiggly line shown in Figure 2. This squiggly line occurs when , where and are the two black trivalent nodes in the tree on . To obtain the fan structure (v), instead of the squiggly line, the square face is subdivided along the other diagonal.
Using the computation of in [BLMM17], available at https://github.com/Saralamboglia/Toric-Degenerations/blob/master/Flag5.rtf and 3.12, we further computed that with (iii) has 938 maximal cones (906 of which are simplicial) and that (iv) has 406 maximal cones. According to [SW05, Conjecture 8.1], the (v) fan structure for has 672 maximal cones.
8.2. Coherent subdivisions
We next discuss coherent subdivisions coming from the positive tropical Grassmannian and positive tropical complete flag variety. When is the Grassmannian and the support is the uniform matroid, A gives rise to the following corollary (which was first proved in [LPW20] and [AHLS20]).
Corollary 8.2.
Let , and suppose it has no coordinates. Then the following statements are equivalent.
-
•
, that is, lies in the strictly positive tropical Grassmannian.
-
•
Every face in the coherent subdivision of the hypersimplex induced by is a positroid polytope.
The coherent subdivisions above (called positroidal subdivisions) were further studied in [SW21], where the finest positroidal subdivisions were characterized in terms of series-parallel matroids. Furthermore, all finest positroidal subdivisions of achieve equality in Speyer’s -vector theorem; in particular, they all consist of facets [SW21, Corollary 6.7].
When is the complete flag variety , and the support is the uniform flag matroid, A gives rise to the following corollary, which appeared in [JLLO, Theorem 20].
Corollary 8.3.
Let , and suppose it has no coordinates. Then the following statements are equivalent.
-
•
, that is, lies in the strictly positive tropical flag variety.
-
•
Every face in the coherent subdivision of the permutohedron induced by is a Bruhat interval polytope.
In light of the results of [SW21], it is natural to ask if one can characterize the finest coherent subdivisions of the permutohedron into Bruhat interval polytopes. Furthermore, do they all have the same -vector?
Explicit computations for show that the answer to the second question is no. We find that with the fan structure (iii) (which agrees with (ii) by [JLLO, §5]) has maximal cones. We choose a point in the relative interior of each of the cones to use as a height function (thinking of points in as weights on the vertices of as in Item (c) of A), then use Sage to compute the corresponding coherent subdivision of . As expected, precisely of the cones induce subdivisions of into Bruhat interval polytopes, see Table 1.
|
|
|
|||||
---|---|---|---|---|---|---|---|
|
|
||||||
|
|
||||||
|
|
||||||
|
|
||||||
|
|
||||||
|
|
||||||
|
|
||||||
|
|
||||||
|
|
||||||
|
|
||||||
|
|
||||||
|
|
||||||
|
|
||||||
|
|
Of the coherent subdivisions coming from maximal cones of , of them contain facets, while the other contain facets. Table 1 lists the facets and -vectors of each of these subdivisions. Note that each Bruhat interval polytope which appears as a facet satisfies . Thus, any Bruhat interval polytope properly contained inside would have the property that , and hence . Since is -dimensional, all of these subdivisions are finest subdivisions.
We note that the finest subdivisions whose -vector is are subdivisions of the permutohedron into cubes. Subdivisions of the permutohedron into Bruhat interval polytopes which are cubes have been previously studied in [HHMP19, Sections 5 and 6] [LMP21], and in [NT22, Section 6]. In particular, there is a subdivision of into Bruhat interval polytopes
The first subdivision in Table 1 has this form.
We can further study the -vectors of subdivisions of which are coarsest (without being trivial), rather than finest. In this case, we observe three different -vectors, each of which occurs in multiple subdivisions. The detailed results of our explicit computations on coarsest subdivisions can be found in Table 2.
|
|
|
|||||
---|---|---|---|---|---|---|---|
|
|
||||||
|
|
||||||
|
|
||||||
|
|
||||||
|
|
||||||
|
|
||||||
|
|
||||||
|
|
||||||
|
|
References
- [AFR10] Federico Ardila, Alex Fink, and Felipe Rincón. Valuations for matroid polytope subdivisions. Canad. J. Math., 62(6):1228–1245, 2010.
- [AHHL21] Nima Arkani-Hamed, Song He, and Thomas Lam. Cluster configuration spaces of finite type. SIGMA Symmetry Integrability Geom. Methods Appl., 17:Paper No. 092, 41, 2021.
- [AHLS20] Nima Arkani-Hamed, Thomas Lam, and Marcus Spradlin. Positive configuration space. 2020. Preprint, arXiv:2003.03904v3.
- [AK06] Federico Ardila and Caroline J. Klivans. The Bergman complex of a matroid and phylogenetic trees. J. Combin. Theory Ser. B, 96(1):38–49, 2006.
- [ARW16] Federico Ardila, Felipe Rincón, and Lauren Williams. Positroids and non-crossing partitions. Trans. Amer. Math. Soc., 368(1):337–363, 2016.
- [ARW17] Federico Ardila, Felipe Rincón, and Lauren Williams. Positively oriented matroids are realizable. J. Eur. Math. Soc. (JEMS), 19(3):815–833, 2017.
- [BB19] Matthew Baker and Nathan Bowler. Matroids over partial hyperstructures. Adv. Math., 343:821–863, 2019.
- [BCTJ22] Carolina Benedetti, Anastasia Chavez, and Daniel Tamayo Jiménez. Quotients of uniform positroids. Electron. J. Combin., 29(1):Paper No. 1.13, 20, 2022.
- [BEZ21] Madeline Brandt, Christopher Eur, and Leon Zhang. Tropical flag varieties. Adv. Math., 384:Paper No. 107695, 41, 2021.
- [BGW03] Alexandre V. Borovik, I. M. Gelfand, and Neil White. Coxeter matroids, volume 216 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, 2003.
- [BK22] Carolina Benedetti and Kolja Knauer. Lattice path matroids and quotients. 2022. Preprint, arXiv:2202.11634.
- [BLMM17] Lara Bossinger, Sara Lamboglia, Kalina Mincheva, and Fatemeh Mohammadi. Computing toric degenerations of flag varieties. In Combinatorial algebraic geometry, volume 80 of Fields Inst. Commun., pages 247–281. Fields Inst. Res. Math. Sci., Toronto, ON, 2017.
- [BLVS+99] Anders Björner, Michel Las Vergnas, Bernd Sturmfels, Neil White, and Günter M. Ziegler. Oriented matroids, volume 46 of Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge, second edition, 1999.
- [Bor22] Jonathan Boretsky. Totally nonnegative tropical flags and the totally nonnegative flag Dressian. 2022. Preprint, arXiv:2208.09128v2.
- [Bos21] Lara Bossinger. Birational sequences and the tropical Grassmannian. J. Algebra, 585:784–803, 2021.
- [Bos22] Lara Bossinger. Tropical totally positive cluster varieties, 2022. Preprint, arXiv:2208.01723.
- [BW22] Sara Billey and Jordan Weaver. Criteria for smoothness of positroid varieties via pattern avoidance, johnson graphs, and spirographs, 2022. Preprint, arXiv:2207.06508.
- [CEGM19] Freddy Cachazo, Nick Early, Alfredo Guevara, and Sebastian Mizera. Scattering equations: from projective spaces to tropical Grassmannians. J. High Energy Phys., (6):039, 32, 2019.
- [dS87] Ilda P.F. da Silva. Quelques propriétés des matroides orientés. Ph.D. Dissertation, Université Paris VI, 1987.
- [Ear22] Nick Early. From weakly separated collections to matroid subdivisions. Comb. Theory, 2(2):Paper No. 2, 35, 2022.
- [Ful97] William Fulton. Young tableaux, volume 35 of London Mathematical Society Student Texts. Cambridge University Press, Cambridge, 1997. With applications to representation theory and geometry.
- [FZ03] Sergey Fomin and Andrei Zelevinsky. Cluster algebras. II. Finite type classification. Invent. Math., 154(1):63–121, 2003.
- [GGMS87] I. M. Gelfand, R. M. Goresky, R. D. MacPherson, and V. V. Serganova. Combinatorial geometries, convex polyhedra, and Schubert cells. Adv. in Math., 63(3):301–316, 1987.
- [GLS08] Christof Geiss, Bernard Leclerc, and Jan Schröer. Preprojective algebras and cluster algebras. In Trends in representation theory of algebras and related topics, EMS Ser. Congr. Rep., pages 253–283. Eur. Math. Soc., Zürich, 2008.
- [GS87] I. M. Gelfand and V. V. Serganova. Combinatorial geometries and the strata of a torus on homogeneous compact manifolds. Uspekhi Mat. Nauk, 42(2(254)):107–134, 287, 1987.
- [Gun19] Trevor Gunn. A Newton polygon rule for formally-real valued fields and multiplicities over the signed tropical hyperfield, 2019. Preprint, arXiv:1911.12274.
- [HHMP19] M. Harada, T. Horiguchi, M. Masuda, and Sondzhon Park. The volume polynomial of regular semisimple Hessenberg varieties and the Gelfand-Tsetlin polytope. Tr. Mat. Inst. Steklova, 305(Algebraicheskaya Topologiya Kombinatorika i Matematicheskaya Fizika):344–373, 2019.
- [HJJS08] Sven Herrmann, Anders Nedergaard Jensen, Michael Joswig, and Bernd Sturmfels. How to draw tropical planes. Electr. J. Comb., 16, 2008.
- [HJS14] Sven Herrmann, Michael Joswig, and David E. Speyer. Dressians, tropical Grassmannians, and their rays. Forum Math., 26(6):1853–1881, 2014.
- [JL22] Manoel Jarra and Oliver Lorscheid. Flag matroids with coefficients, 2022. Preprint, arXiv:2204.04658.
- [JLLO] Michael Joswig, Georg Loho, Dante Luber, and Jorge Alberto Olarte. Generalized permutahedra and positive flag Dressians. Int. Math. Res. Not. (to appear). arXiv:2111.13676.
- [JLS21] Dennis Jahn, Robert Löwe, and Christian Stump. Minkowski decompositions for generalized associahedra of acyclic type. Algebr. Comb., 4(5):757–775, 2021.
- [Kun86] Joseph P. S. Kung. Strong maps. pages 224–253, 1986.
- [KW15] Yuji Kodama and Lauren Williams. The full Kostant-Toda hierarchy on the positive flag variety. Comm. Math. Phys., 335(1):247–283, 2015.
- [LF19] Ian Le and Chris Fraser. Tropicalization of positive Grassmannians. Selecta Math. (N.S.), 25(5):Paper No. 75, 55, 2019.
- [LMP21] Eunjeong Lee, Mikiya Masuda, and Seonjeong Park. Toric Bruhat interval polytopes. J. Combin. Theory Ser. A, 179:Paper No. 105387, 41, 2021.
- [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.
- [Lus94] G. Lusztig. Total positivity in reductive groups. In Lie theory and geometry, volume 123 of Progr. Math., pages 531–568. Birkhäuser Boston, Boston, MA, 1994.
- [Mar10] T. Markwig. A field of generalised Puiseux series for tropical geometry. Rend. Semin. Mat. Univ. Politec. Torino, 68(1):79–92, 2010.
- [MR04] B. R. Marsh and K. Rietsch. Parametrizations of flag varieties. Represent. Theory, 8:212–242 (electronic), 2004.
- [MS15] Diane Maclagan and Bernd Sturmfels. Introduction to tropical geometry, volume 161 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2015.
- [MS18] Kazuo Murota and Akiyoshi Shioura. On equivalence of -concavity of a set function and submodularity of its conjugate. J. Oper. Res. Soc. Japan, 61(2):163–171, 2018.
- [NT22] Philippe Nadeau and Vasu Tewari. Remixed Eulerian numbers, 2022. Preprint, arXiv:2208.04128.
- [Oh08] Suho Oh. Contraction and restriction of positroids in terms of decorated permutations, 2008.
- [OPS19] Jorge Alberto Olarte, Marta Panizzut, and Benjamin Schröter. On local Dressians of matroids. In Algebraic and geometric combinatorics on lattice polytopes, pages 309–329. World Sci. Publ., Hackensack, NJ, 2019.
- [Oxl11] James Oxley. Matroid theory, volume 21 of Oxford Graduate Texts in Mathematics. Oxford University Press, Oxford, 2 edition, 2011.
- [Poo93] Bjorn Poonen. Maximally complete fields. Enseign. Math. (2), 39(1-2):87–106, 1993.
- [Pos] Alexander Postnikov. Total positivity, Grassmannians, and networks. Preprint, http://math.mit.edu/~apost/papers/tpgrass.pdf.
- [PS04] Lior Pachter and Bernd Sturmfels. Tropical geometry of statistical models. Proc. Natl. Acad. Sci. USA, 101(46):16132–16137, 2004.
- [Rie98] Konstanze Christina Rietsch. Total Positivity and Real Flag Varieties. Ph.D. thesis, Massachusetts Institute of Technology, 1998.
- [Rie06] K. Rietsch. Closure relations for totally nonnegative cells in . Math. Res. Lett., 13(5-6):775–786, 2006.
- [RW08] Konstanze Rietsch and Lauren Williams. The totally nonnegative part of is a CW complex. Transform. Groups, 13(3-4):839–853, 2008.
- [RW19] K. Rietsch and L. Williams. Newton-Okounkov bodies, cluster duality, and mirror symmetry for Grassmannians. Duke Math. J., 168(18):3437–3527, 2019.
- [Spe08] David E. Speyer. Tropical linear spaces. SIAM J. Discrete Math., 22(4):1527–1558, 2008.
- [SS04] David Speyer and Bernd Sturmfels. The tropical Grassmannian. Adv. Geom., 4(3):389–411, 2004.
- [SW05] David Speyer and Lauren Williams. The tropical totally positive Grassmannian. J. Algebraic Combin., 22(2):189–210, 2005.
- [SW21] David Speyer and Lauren K. Williams. The positive Dressian equals the positive tropical Grassmannian. Trans. Amer. Math. Soc. Ser. B, 8:330–353, 2021.
- [TW13] Kelli Talaska and Lauren Williams. Network parametrizations for the Grassmannian. Algebra Number Theory, 7(9):2275–2311, 2013.
- [TW15] E. Tsukerman and L. Williams. Bruhat interval polytopes. Adv. Math., 285:766–810, 2015.