This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Polyhedral and tropical geometry of flag positroids

Jonathan Boretsky, Christopher Eur, Lauren Williams Centre de Recherches Mathématiques in Montreal jonathan.boretsky@mail.mcgill.ca Carnegie Mellon University ceur@cmu.edu Harvard University williams@math.harvard.edu
Abstract.

A flag positroid of ranks 𝐫:=(r1<<rk)\mathbf{r}:=(r_{1}<\dots<r_{k}) on [n][n] is a flag matroid that can be realized by a real rk×nr_{k}\times n matrix AA such that the ri×rir_{i}\times r_{i} minors of AA involving rows 1,2,,ri1,2,\dots,r_{i} are nonnegative for all 1ik1\leq i\leq k. In this paper we explore the polyhedral and tropical geometry of flag positroids, particularly when 𝐫:=(a,a+1,,b)\mathbf{r}:=(a,a+1,\dots,b) is a sequence of consecutive numbers. In this case we show that the nonnegative tropical flag variety TrFl𝐫,n0\operatorname{TrFl}_{\mathbf{r},n}^{\geq 0} equals the nonnegative flag Dressian FlDr𝐫,n0\operatorname{FlDr}_{\mathbf{r},n}^{\geq 0}, and that the points 𝝁=(μa,,μb){\boldsymbol{\mu}}=(\mu_{a},\ldots,\mu_{b}) of TrFl𝐫,n0=FlDr𝐫,n0\operatorname{TrFl}_{\mathbf{r},n}^{\geq 0}=\operatorname{FlDr}_{\mathbf{r},n}^{\geq 0} give rise to coherent subdivisions of the flag positroid polytope P(𝝁¯)P(\underline{{\boldsymbol{\mu}}}) 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 (2)(\leq 2)-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 (χ1,,χk)(\chi_{1},\dots,\chi_{k}) which is also an oriented flag matroid. We then prove that every positively oriented flag matroid of ranks 𝐫=(a,a+1,,b)\mathbf{r}=(a,a+1,\dots,b) 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 nn and dd with d<nd<n, we let [n][n] denote the set {1,,n}\{1,\ldots,n\} and we let ([n]d){[n]\choose d} denote the collection of all dd-element subsets of [n][n]. Given a subset S[n]S\subseteq[n] we let 𝐞S\mathbf{e}_{S} denote the sum of standard basis vectors iS𝐞i\sum_{i\in S}\mathbf{e}_{i}. For a collection ([n]d)\mathcal{B}\subset{[n]\choose d}, we let

P()=the convex hull of {𝐞B:B} in n.P(\mathcal{B})=\text{the convex hull of $\{\mathbf{e}_{B}:B\in\mathcal{B}\}$ in $\mathbb{R}^{n}$}.

The collection \mathcal{B} is said to define a matroid MM of rank dd on [n][n] if every edge of the polytope P()P(\mathcal{B}) is parallel to 𝐞i𝐞j\mathbf{e}_{i}-\mathbf{e}_{j} for some ij[n]i\neq j\in[n]. In this case, we call \mathcal{B} the set of bases of MM, and define the matroid polytope P(M)P(M) of MM to be the polytope P()P(\mathcal{B}). When \mathcal{B} indexes the nonvanishing Plücker coordinates of an element AA of the Grassmannian Grd,n()\operatorname{Gr}_{d,n}(\mathbb{C}), we say that AA realizes MM, and it is well-known that P()P(\mathcal{B}) is the moment map image of the closure of the torus orbit of AA 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 𝐫=(r1,,rk)\mathbf{r}=(r_{1},\ldots,r_{k}) be a sequence of increasing integers in [n][n]. A flag matroid of ranks 𝐫\mathbf{r} on [n][n] is a sequence 𝑴=(M1,,Mk){\boldsymbol{M}}=(M_{1},\ldots,M_{k}) of matroids of ranks (r1,,rk)(r_{1},\ldots,r_{k}) on [n][n] such that all vertices of the polytope

P(𝑴)=P(M1)++P(Mk), the Minkowski sum of matroid polytopes,P({\boldsymbol{M}})=P(M_{1})+\cdots+P(M_{k}),\text{ the Minkowski sum of matroid polytopes},

are equidistant from the origin. The polytope P(𝑴)P({\boldsymbol{M}}) is called the flag matroid polytope of 𝑴{\boldsymbol{M}}; we sometimes say it is a flag matroid polytope of rank 𝐫\mathbf{r}.

Flag matroids are exactly the type AA 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 (r1,,rk)(r_{1},\dots,r_{k}) is an rk×nr_{k}\times n matrix AA over a field such that for each 1ik1\leq i\leq k, the ri×nr_{i}\times n submatrix of AA formed by the first rir_{i} rows of AA is a realization of MiM_{i}. 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 G/PG/P) [GS87, BGW03]. When a flag matroid 𝑴{\boldsymbol{M}} can be realized by a point AA in the flag variety, then its matroid polytope P(𝑴)P({\boldsymbol{M}}) is the moment map image of the closure of the torus orbit of AA 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 𝐫=(r1,,rk)\mathbf{r}=(r_{1},\cdots,r_{k}) be a sequence of increasing integers in [n][n]. We say that a flag matroid (M1,,Mk)(M_{1},\dots,M_{k}) of ranks 𝐫\mathbf{r} on [n][n] is a flag positroid if it has a realization by a real matrix AA such that the ri×nr_{i}\times n submatrix of AA formed by the first rir_{i} rows of AA has all nonnegative minors for each 1ik1\leq i\leq k.

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 k=1k=1 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 𝕋={}\mathbb{T}=\mathbb{R}\cup\{\infty\} be the set underlying the tropical hyperfield, endowed with the topology such that log:0𝕋-\log:\mathbb{R}_{\geq 0}\to\mathbb{T} is a homeomorphism. Given a point w𝕋([n]r)w\in\mathbb{T}^{\binom{[n]}{r}}, we define the support of ww to be w¯={S([n]r):wS}\underline{w}=\{S\in\binom{[n]}{r}:w_{S}\neq\infty\}. When w¯\underline{w} is the set of bases of a matroid, we identify w¯\underline{w} with that matroid. Let (𝕋([n]r))\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{r}}\right) be the tropical projective space of 𝕋([n]r)\mathbb{T}^{\binom{[n]}{r}}, which is defined as (𝕋([n]r){(,,)})/\big{(}\mathbb{T}^{\binom{[n]}{r}}\setminus\{(\infty,\ldots,\infty)\}\big{)}/\sim, where www\sim w^{\prime} if w=w+(c,,c)w=w^{\prime}+(c,\ldots,c) for some cc\in\mathbb{R}.

Our main result is the following.

Theorem A.

Suppose 𝐫\mathbf{r} is a sequence of consecutive integers (a,,b)(a,\ldots,b) for some 1abn1\leq a\leq b\leq n. Then, for 𝝁=(μa,,μb)i=ab(𝕋([n]i)){\boldsymbol{\mu}}=(\mu_{a},\ldots,\mu_{b})\in\prod_{i=a}^{b}\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{i}}\right), the following statements are equivalent:

  1. (a)

    𝝁TrFl𝐫,n0{\boldsymbol{\mu}}\in\operatorname{TrFl}_{\mathbf{r},n}^{\geq 0}, the nonnegative tropicalization of the flag variety, i.e. the closure of the coordinate-wise valuation of points in Fl𝐫,n(𝒞0)\operatorname{Fl}_{\mathbf{r},n}(\mathcal{C}_{\geq 0}).

  2. (b)

    𝝁FlDr𝐫,n0{\boldsymbol{\mu}}\in\operatorname{FlDr}_{\mathbf{r},n}^{\geq 0}, the nonnegative flag Dressian, i.e. the “solutions" to the positive-tropical Grassmann-Plücker and incidence-Plücker relations.

  3. (c)

    Every face in the coherent subdivision 𝒟𝝁\mathcal{D}_{\boldsymbol{\mu}} of the polytope P(𝝁¯)=P(μ1¯)++P(μk¯)P(\underline{{\boldsymbol{\mu}}})=P(\underline{\mu_{1}})+\cdots+P(\underline{\mu_{k}}) induced by 𝝁{\boldsymbol{\mu}} is a flag positroid polytope (of rank 𝐫\mathbf{r}).

  4. (d)

    Every face of dimension at most 2 in the subdivision 𝒟𝝁\mathcal{D}_{\boldsymbol{\mu}} of P(𝝁¯)P(\underline{\boldsymbol{\mu}}) is a flag positroid polytope (of rank 𝐫\mathbf{r}).

  5. (e)

    The support 𝝁¯\underline{{\boldsymbol{\mu}}} of 𝝁{\boldsymbol{\mu}} is a flag matroid, 𝝁{\boldsymbol{\mu}} satisfies every three-term positive-tropical incidence relation when a<ba<b (respectively, every three-term positive-tropical Grassmann-Plücker relation when a=ba=b), and either 𝝁¯\underline{{\boldsymbol{\mu}}} consists of uniform matroids or μiDri;n0\mu_{i}\in\operatorname{Dr}_{i;n}^{\geq 0} for at least one aiba\leq i\leq b.

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 𝐫=(d)\mathbf{r}=(d) 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 𝝁{\boldsymbol{\mu}} has no coordinates equal to \infty) into positroid polytopes. And when 𝐫=(1,2,,n)\mathbf{r}=(1,2,\dots,n), 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 𝝁{\boldsymbol{\mu}} has no coordinates equal to \infty) into Bruhat interval polytopes. We illustrate this relationship in the case where 𝝁{\boldsymbol{\mu}} has no coordinates equal to \infty in Figure 1.

\tkzDefPoint\tkzDefPoint\tkzDefPoint\tkzDefPoint\tkzDefPoint\tkzDefPoint\tkzDrawPolygon\tkzDrawPolygone24e_{24}e14e_{14}e23e_{23}e13e_{13}e12e_{12}e34e_{34}
\tkzDefPoint\tkzDefPoint\tkzDefPoint\tkzDefPoint\tkzDefPoint\tkzDefPoint\tkzDrawPolygon(2,3,1)(2,3,1)(1,3,2)(1,3,2)(3,2,1)(3,2,1)(3,1,2)(3,1,2)(2,1,3)(2,1,3)(1,2,3)(1,2,3)
Figure 1. At left: the coherent subdivision of the hypersimplex into positroid polytopes induced by a point 𝝁Dr2,4>0{\boldsymbol{\mu}}\in\operatorname{Dr}^{>0}_{2,4} such that μ13+μ24=μ23+μ14<μ12+μ34\mu_{13}+\mu_{24}=\mu_{23}+\mu_{14}<\mu_{12}+\mu_{34}. At right: the coherent subdivision of the permutohedron into Bruhat interval polytopes induced by a point 𝝁FlDr(1,2,3),3>0{\boldsymbol{\mu}}\in\operatorname{FlDr}^{>0}_{(1,2,3),3} such that μ2+μ13=μ1+μ23<μ3+μ12\mu_{2}+\mu_{13}=\mu_{1}+\mu_{23}<\mu_{3}+\mu_{12}.

We prove the equivalence (a)\iff(b) in Section 3.2, the implications (b)\implies(c)\implies(d)\implies(b) in Section 5.2, and the equivalence (b)\iff(e) in Section 6.1.

A has applications to flag positroid polytopes.

Corollary 1.4.

For a flag matroid 𝑴=(Ma,Ma+1,,Mb){\boldsymbol{M}}=(M_{a},M_{a+1},\dots,M_{b}) of consecutive ranks 𝐫=(a,a+1,,b)\mathbf{r}=(a,a+1,\dots,b), its flag matroid polytope P(𝑴)P({\boldsymbol{M}}) is a flag positroid polytope if and only if its (2)(\leq 2)-dimensional faces are flag positroid polytopes (of rank 𝐫\mathbf{r}).

Proof.

Let 𝝁=(μa,,μb){\boldsymbol{\mu}}=(\mu_{a},\dots,\mu_{b}), with μi{0,}([n]i)\mu_{i}\in\{0,\infty\}^{{[n]\choose i}}, where the coordinates of each μi\mu_{i} are either 0 or \infty based on whether we have a basis or nonbasis of MiM_{i}. This gives rise to the trivial subdivision of the corresponding flag matroid polytope P(𝝁¯)=P(𝑴)P(\underline{{\boldsymbol{\mu}}})=P({\boldsymbol{M}}). The result now follows from the equivalence of (c) and (d) in A. ∎

In the Grassmannian case, that is, the case that 𝐫=(d)\mathbf{r}=(d) is a single integer, the flag positroid polytopes of rank 𝐫\mathbf{r} 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 0 and 11. 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 (M1,,Mk)(M_{1},\ldots,M_{k}) is a sequence of positroids on [n][n] of consecutive ranks 𝐫=(r1,,rk)\mathbf{r}=(r_{1},\dots,r_{k}). Then, when considered as a sequence of positively oriented matroids, (M1,,Mk)(M_{1},\ldots,M_{k}) 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 (χ1,,χk)(\chi_{1},\dots,\chi_{k}) which is also an oriented flag matroid. 1.5 then says that every positively oriented flag matroid of consecutive ranks (r1,,rk)(r_{1},\dots,r_{k}) 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 \mathbb{R} 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 𝐫=(r1,,rk)\mathbf{r}=(r_{1},\dots,r_{k}) fails to satisfy the consecutive rank condition.

Question 1.6.

Suppose MM and MM^{\prime} are positroids on [n][n] such that, when considered as positively oriented matroids, they form an oriented flag matroid (M,M)(M,M^{\prime}). Is (M,M)(M,M^{\prime}) 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 (M,M)(M,M^{\prime}), one can always find a flag matroid (M1,,Mk)(M_{1},\ldots,M_{k}) of consecutive ranks such that M1=MM_{1}=M and Mk=MM_{k}=M^{\prime}. However, the analogous statement fails for flag positroids: See 4.6 for an example of a flag positroid (M,M)(M,M^{\prime}) on [4][4] of ranks (1,3)(1,3) such that there is no flag positroid (M,M2,M)(M,M_{2},M^{\prime}) with rank of M2M_{2} 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 AA in the nonnegative part (G/P)0(G/P)^{\geq 0} (in the sense of Lusztig) of a flag variety G/PG/P. When 𝐫\mathbf{r} is a sequence of consecutive integers, it then follows from [BK22] that generalized Bruhat interval polytopes for Fl𝐫;n0\operatorname{Fl}_{\mathbf{r};n}^{\geq 0} are precisely the flag positroid polytopes of ranks 𝐫\mathbf{r}. 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 (z(1),,z(n))(z(1),\dots,z(n)) for all permutations zz lying in some Bruhat interval [u,v][u,v].

We can now restate 1.4 as follows.

Corollary 1.7.

For a flag matroid on [n][n] of consecutive ranks 𝐫\mathbf{r}, its flag matroid polytope is a generalized Bruhat interval polytope if and only if its (2)(\leq 2)-dimensional faces are generalized Bruhat interval polytopes. In particular, for a complete flag matroid on [n][n], its flag matroid polytope is a Bruhat interval polytope if and only if its (2)(\leq 2)-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)\implies(c)\implies(d)\implies(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)\iff(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 TrFl𝐫;n>0\operatorname{TrFl}_{\mathbf{r};n}^{>0}; 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 TrFl4>0\operatorname{TrFl}_{4}^{>0}.

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 n+n\in\mathbb{Z}_{+} and let 𝐫={r1<<rk}[n].\mathbf{r}=\{r_{1}<\dots<r_{k}\}\subseteq[n]. For a field 𝕜\mathbbm{k}, let G=GLn(𝕜)G=\operatorname{GL}_{n}(\mathbbm{k}), and let P𝐫;n(𝕜)\operatorname{P}_{\mathbf{r};n}(\mathbbm{k}) denote the parabolic subgroup of GG of block upper-triangular matrices with diagonal blocks of sizes r1,r2r1,,rkrk1,nrkr_{1},r_{2}-r_{1},\dots,r_{k}-r_{k-1},n-r_{k}. We define the partial flag variety

Fl𝐫;n(𝕜):=GLn(𝕜)/P𝐫;n(𝕜).\operatorname{Fl}_{\mathbf{r};n}(\mathbbm{k}):=\operatorname{GL}_{n}(\mathbbm{k})/\operatorname{P}_{\mathbf{r};n}(\mathbbm{k}).

As usual, we identify Fl𝐫;n(𝕜)\operatorname{Fl}_{\mathbf{r};n}(\mathbbm{k}) with the variety of partial flags of subspaces in 𝕜n\mathbbm{k}^{n}:

Fl𝐫;n(𝕜)={(V1Vk):Vi a linear subspace of 𝕜n of dimension ri for i=1,,k}.\operatorname{Fl}_{\mathbf{r};n}(\mathbbm{k})=\{(V_{1}\subset\cdots\subset V_{k}):\text{$V_{i}$ a linear subspace of $\mathbbm{k}^{n}$ of dimension $r_{i}$ for $i=1,\ldots,k$}\}.

We write Fln(𝕜)\operatorname{Fl}_{n}(\mathbbm{k}) for the complete flag variety Fl1,2,,n;n(𝕜)\operatorname{Fl}_{1,2,\ldots,n;n}(\mathbbm{k}). Note that Fln(𝕜)\operatorname{Fl}_{n}(\mathbbm{k}) can be identified with GLn(𝕜)/B(𝕜)\operatorname{GL}_{n}(\mathbbm{k})/B(\mathbbm{k}), where B(𝕜)B(\mathbbm{k}) is the subgroup of upper-triangular matrices. There is a natural projection π\pi from Fln(𝕜)\operatorname{Fl}_{n}(\mathbbm{k}) to any partial flag variety by simply forgetting some of the subspaces.

If AA is an rk×nr_{k}\times n matrix such that VriV_{r_{i}} is the span of the first rir_{i} rows, we say that AA is a realization of V:=(V1Vk)Fl𝐫;nV:=(V_{1}\subset\dots\subset V_{k})\in\operatorname{Fl}_{\mathbf{r};n}. Given any realization AA of VV and any 1ik1\leq i\leq k, we have the Plücker coordinates or flag minors pI(A)p_{I}(A) where I([n]ri)I\in{[n]\choose r_{i}}; concretely, pI(A)p_{I}(A) is the determinant of the submatrix of AA occupying the first rir_{i} rows and columns II. This gives the Plücker embedding of Fl𝐫;n(𝕜)\operatorname{Fl}_{\mathbf{r};n}(\mathbbm{k}) into ([n]r1)1××([n]rk)1\mathbb{P}^{{[n]\choose r_{1}}-1}\times\dots\times\mathbb{P}^{{[n]\choose r_{k}}-1} taking VV to ((pI(A))I([n]r1),,(pI(A))I([n]rk))\left(\big{(}p_{I}(A)\big{)}_{I\in{[n]\choose r_{1}}},\dots,\big{(}p_{I}(A)\big{)}_{I\in{[n]\choose r_{k}}}\right).

We now let 𝕜\mathbbm{k} be the field \mathbb{R} of real numbers. With this understanding, we will often drop the \mathbb{R} from our notation.

Definition 2.1.

We say that a real matrix is totally positive if all of its minors are positive. We let GLn>0\operatorname{GL}_{n}^{>0} denote the subset of GLn\operatorname{GL}_{n} 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 Fl𝐫;n\operatorname{Fl}_{\mathbf{r};n}, denoted by Fl𝐫;n>0\operatorname{Fl}_{\mathbf{r};n}^{>0}, as the image of GLn>0\operatorname{GL}_{n}^{>0} inside Fl𝐫;n=GLn/P𝐫;n\operatorname{Fl}_{\mathbf{r};n}=\operatorname{GL}_{n}/\operatorname{P}_{\mathbf{r};n}. We define the (Lusztig) nonnegative part of Fl𝐫;n\operatorname{Fl}_{\mathbf{r};n}, denoted by Fl𝐫;n0\operatorname{Fl}_{\mathbf{r};n}^{\geq 0}, as the closure of Fl𝐫;n>0\operatorname{Fl}_{\mathbf{r};n}^{>0} in the Euclidean topology.

We define the Plücker positive part (respectively, Plücker nonnegative part) of Fl𝐫;n\operatorname{Fl}_{\mathbf{r};n} to be the subset of Fl𝐫;n\operatorname{Fl}_{\mathbf{r};n} 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 Fl𝐫;n\operatorname{Fl}_{\mathbf{r};n} is a subset of the Plücker positive part of Fl𝐫;n\operatorname{Fl}_{\mathbf{r};n}, 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 Fl𝐫;n\operatorname{Fl}_{\mathbf{r};n} equals the Plücker positive (respectively, Plücker nonnegative) part of Fl𝐫;n\operatorname{Fl}_{\mathbf{r};n} if and only if the set 𝐫\mathbf{r} 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 𝐫\mathbf{r} consists of consecutive integers, we will use the two notions interchangeably when there is no ambiguity.

Let BB and BB^{-} be the opposite Borel subgroups consisting of upper-triangular and lower-triangular matrices. Let W=SnW=S_{n} be the Weyl group of GLn\operatorname{GL}_{n}. Given u,vWu,v\in W, the Richardson variety is the intersection of opposite Bruhat cells

u,v:=(Bv˙B/B)(Bu˙B/B),\mathcal{R}_{u,v}:=(B\dot{v}B/B)\cap(B^{-}\dot{u}B/B),

where v˙\dot{v} and u˙\dot{u} denote permutation matrices in GLn\operatorname{GL}_{n} representing vv and uu. It is well-known that u,v\mathcal{R}_{u,v} is nonempty precisely when uvu\leq v in Bruhat order, and in that case is irreducible of dimension (v)(u)\ell(v)-\ell(u).

For u,vWu,v\in W with uvu\leq v, let u,v>0:=u,vFln0\mathcal{R}_{u,v}^{>0}:=\mathcal{R}_{u,v}\cap\operatorname{Fl}_{n}^{\geq 0} be the positive part of the Richardson variety. Lusztig conjectured and Rietsch proved [Rie98] that

(1) Fln0=uvu,v>0\operatorname{Fl}_{n}^{\geq 0}=\bigsqcup_{u\leq v}\mathcal{R}_{u,v}^{>0}

is a cell decomposition of Fln0\operatorname{Fl}_{n}^{\geq 0}. Moreover, Rietsch showed that one obtains a cell decomposition of the nonnegative partial flag variety Fl𝐫;n0\operatorname{Fl}_{\mathbf{r};n}^{\geq 0} by projecting the cell decomposition of Fln0\operatorname{Fl}_{n}^{\geq 0} [Rie98], [Rie06, Section 6]. Specifically, if we let W𝐫W_{\mathbf{r}} be the parabolic subgroup of WW generated by the simple reflections {si| 1in1 and i{r1,,rk}}\{s_{i}\ |\ 1\leq i\leq n-1\text{ and }i\notin\{r_{1},\dots,r_{k}\}\}, then one obtains a cell decomposition by using the projections π(u,v>0)\pi(\mathcal{R}_{u,v}^{>0}) of the cells u,v>0\mathcal{R}_{u,v}^{>0} where uvu\leq v and vv is a minimal-length coset representative of W/W𝐫W/W_{\mathbf{r}}. (We note moreover that Rietsch’s results hold for GG a semisimple, simply connected linear algebraic group over \mathbb{C} split over \mathbb{R}).

In the case of the Grassmannian, Postnikov studied the Plücker nonnegative part Grd,n0\operatorname{Gr}_{d,n}^{\geq 0} of the Grassmannian, and gave a decomposition of it into positroid cells S>0S_{\mathcal{B}}^{>0} by intersecting Grd,n0\operatorname{Gr}_{d,n}^{\geq 0} with the matroid strata [Pos]. Concretely, if \mathcal{B} is the collection of bases of an element of Grd,n0\operatorname{Gr}_{d,n}^{\geq 0}, then S>0={AGrd,n0|pI(A)0 if and only if I}S_{\mathcal{B}}^{>0}=\{A\in\operatorname{Gr}_{d,n}^{\geq 0}\ |\ p_{I}(A)\neq 0\text{ if and only if }I\in\mathcal{B}\}. This cell decomposition of Grd,n0\operatorname{Gr}_{d,n}^{\geq 0} 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 uu and vv in SnS_{n} with uvu\leq v in Bruhat order, the Bruhat interval polytope Pu,vP_{u,v} is defined as

(2) Pu,v=Conv{(x(1),x(2),,x(n))|uxv}n.P_{u,v}=\operatorname{Conv}\{(x(1),x(2),\dots,x(n))\ |\ u\leq x\leq v\}\subset\mathbb{R}^{n}.

We also define the (twisted) Bruhat interval polytope P~u,v\tilde{P}_{u,v} by

(3) P~u,v=Conv{(n+1x1(1),n+1x1(2),,n+1x1(n))|uxv}n.\tilde{P}_{u,v}=\operatorname{Conv}\{(n+1-x^{-1}(1),n+1-x^{-1}(2),\dots,n+1-x^{-1}(n))\ |\ u\leq x\leq v\}\subset\mathbb{R}^{n}.

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 AA in the cell u,v>0Fln0\mathcal{R}_{u,v}^{>0}\subset\operatorname{Fl}_{n}^{\geq 0} (thought of as an n×nn\times n matrix), and let MiM_{i} be the matroid represented by the first ii rows of AA, then P~u,v\tilde{P}_{u,v} is the Minkowski sum of the matroid polytopes P(M1),,P(Mn)P(M_{1}),\dots,P(M_{n}) [KW15, Corollary 6.11]. In particular, P~u,v\tilde{P}_{u,v} is the matroid polytope of the flag matroid M1,,MnM_{1},\ldots,M_{n}.

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 G/P=G/PJG/P=G/P_{J}, let WJW_{J} be the associated parabolic subgroup of the Weyl group WW, and let u,vWu,v\in W with uvu\leq v in Bruhat order and vv a minimal-length coset representative of W/WJW/W_{J}. Let π\pi denote the projection from G/BG/B to G/PG/P, and let AA be an element of the cell π(u,v>0)\pi(\mathcal{R}_{u,v}^{>0}) of (Lusztig’s definition of) (G/P)0(G/P)^{\geq 0}.

A generalized Bruhat interval polytope P~u,vJ\tilde{P}_{u,v}^{J} 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 AA in G/PG/P (which is a Coxeter matroid polytope)

  • the moment map image of the closure of the cell π(u,v>0)\pi(\mathcal{R}_{u,v}^{>0})

  • the moment map image of the closure of the projected Richardson variety π(u,v)\pi(\mathcal{R}_{u,v})

  • the convex hull Conv{zρJ|uzv}𝔱,\operatorname{Conv}\{z\cdot\rho_{J}\ |\ u\leq z\leq v\}\subset\mathfrak{t}_{\mathbb{R}}^{*},
    where ρJ\rho_{J} is the sum of fundamental weights jJωj\sum_{j\in J}\omega_{j}, and 𝔱\mathfrak{t}_{\mathbb{R}}^{*} is the dual of the real part of the Lie algebra 𝔱\mathfrak{t} of the torus TGT\subset G.

Remark 2.7.

When G=GLnG=\operatorname{GL}_{n} with fundamental weights 𝐞1,𝐞1+𝐞2,,𝐞1++𝐞n1\mathbf{e}_{1},\mathbf{e}_{1}+\mathbf{e}_{2},\ldots,\mathbf{e}_{1}+\cdots+\mathbf{e}_{n-1}, each generalized Bruhat interval polytope P~u,vJ\tilde{P}_{u,v}^{J} is the flag positroid polytope associated to a matrix AA representing a point of Fl𝐫;n0\operatorname{Fl}_{\mathbf{r};n}^{\geq 0}, with 𝐫=(r1,,rk)\mathbf{r}=(r_{1},\dots,r_{k}). In this case the generalized Bruhat interval polytope is precisely the Minkowski sum P(M1)++P(Mk)P(M_{1})+\dots+P(M_{k}) of the matroid polytopes P(Mi)P(M_{i}), where MiM_{i} is the matroid realized by the first rir_{i} rows of AA. In particular, the generalized Bruhat interval polytope P~u,v=P~u,v\tilde{P}_{u,v}^{\emptyset}=\tilde{P}_{u,v} is the Minkowski sum P(M1)++P(Mn)P(M_{1})+\dots+P(M_{n}), where MiM_{i} is the positroid realized by the first ii rows of any matrix representing a point of Av,w>0A\in\mathcal{R}_{v,w}^{>0}. We will discuss how to read off the matroids MiM_{i} from (u,v)(u,v) in Section 7.2.

As mentioned in the introduction, when 𝐫\mathbf{r} is a sequence of consecutive ranks, the generalized Bruhat interval polytopes for Fl𝐫;n0\operatorname{Fl}_{\mathbf{r};n}^{\geq 0} are precisely the flag positroid polytopes of ranks 𝐫\mathbf{r}. When 𝐫=(1,2,,n)\mathbf{r}=(1,2,\dots,n), we recover the notion of Bruhat interval polytope, and when 𝐫\mathbf{r} 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 w=(w1,,wm)𝕋mw=(w_{1},\ldots,w_{m})\in\mathbb{T}^{m}, we write w¯\overline{w} for its image in the tropical projective space (𝕋m)\mathbb{P}(\mathbb{T}^{m}). For a=(a1,,an)ma=(a_{1},\ldots,a_{n})\in\mathbb{Z}^{m}, write aw=a1w1++amwma\bullet w=a_{1}w_{1}+\cdots+a_{m}w_{m}.

Definition 3.1.

For a real homogeneous polynomial

f=a𝒜caxa[x1,,xm],where 𝒜 is a finite subset of 0m and 0ca,f=\sum_{a\in\mathcal{A}}c_{a}x^{a}\in\mathbb{R}[x_{1},\ldots,x_{m}],\quad\text{where $\mathcal{A}$ is a finite subset of $\mathbb{Z}_{\geq 0}^{m}$ and $0\neq c_{a}\in\mathbb{R}$,}

the extended tropical hypersurface Vtrop(f)V_{\operatorname{trop}}(f) and the nonnegative tropical hypersurface Vtrop0(f)V_{\operatorname{trop}}^{\geq 0}(f) are subsets of the tropical projective space (𝕋m)\mathbb{P}(\mathbb{T}^{m}) defined by

Vtrop(f)\displaystyle V_{\operatorname{trop}}(f) ={w¯(𝕋m)|the minimum in mina𝒜(aw), if finite, is achieved at least twice},\displaystyle=\left\{\overline{w}\in\mathbb{P}(\mathbb{T}^{m})\ \middle|\ \text{the minimum in $\min_{a\in\mathcal{A}}(a\bullet w)$, if finite, is achieved at least twice}\right\},
and
Vtrop0(f)\displaystyle V_{\operatorname{trop}}^{\geq 0}(f) ={w¯(𝕋m)|the minimum in mina𝒜(aw), if finite, is achieved at least twice, including at some a,a𝒜 such that ca and ca have opposite signs}.\displaystyle=\left\{\overline{w}\in\mathbb{P}(\mathbb{T}^{m})\ \middle|\ \begin{matrix}\text{the minimum in $\displaystyle\min_{a\in\mathcal{A}}(a\bullet w)$, if finite, is achieved at least twice,}\\ \text{ including at some $a,a^{\prime}\in\mathcal{A}$ such that $c_{a}$ and $c_{a^{\prime}}$ have opposite signs}\end{matrix}\right\}.

We say that a point satisfies the tropical relation of ff if it is in Vtrop(f)V_{\operatorname{trop}}(f), and that it satisfies the positive-tropical relation of ff if it is in Vtrop0(f)V_{\operatorname{trop}}^{\geq 0}(f).

When ff is a multihomogeneous real polynomial, we define Vtrop(f)V_{\operatorname{trop}}(f) and Vtrop0(f)V_{\operatorname{trop}}^{\geq 0}(f) 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 0<rs<n0<r\leq s<n, the (single-exchange) Plücker relations of type (r,s;n)(r,s;n) are polynomials in variables {xI:I([n]r)([n]s)}\{x_{I}:I\in\binom{[n]}{r}\cup\binom{[n]}{s}\} defined as

𝒫r,s;n={jJIsign(j,I,J)xIjxJj|I([n]r1),J([n]s+1)},\mathscr{P}_{r,s;n}=\left\{\sum_{j\in J\setminus I}\operatorname{sign}(j,I,J)x_{I\cup j}x_{J\setminus j}\left|I\in{\binom{[n]}{r-1}},\>J\in{\binom{[n]}{s+1}}\right.\right\},

where sign(j,I,J)=(1)|{kJ|k<j}|+|{iI|j<i}|\operatorname{sign}(j,I,J)=(-1)^{|\{k\in J|k<j\}|+|\{i\in I|j<i\}|}. When r=sr=s, the elements of 𝒫r,r;n\mathscr{P}_{r,r;n} are called the Grassmann-Plücker relations (of type (r;n)(r;n)), and when r<sr<s, the elements of 𝒫r,s;n\mathscr{P}_{r,s;n} are called the incidence-Plücker relations (of type (r,s;n)(r,s;n)).

As in the introduction, let 𝐫=(r1<<rk)\mathbf{r}=(r_{1}<\cdots<r_{k}) be a sequence of increasing integers in [n][n]. We let 𝒫𝐫;n=rsr,s𝐫𝒫r,s;n,\mathscr{P}_{\mathbf{r};n}=\bigcup_{\begin{subarray}{c}r\leq s\\ r,s\in\mathbf{r}\end{subarray}}\mathscr{P}_{r,s;n}, and let 𝒫𝐫;n\langle\mathscr{P}_{\mathbf{r};n}\rangle be the ideal generated by the elements of 𝒫𝐫;n\mathscr{P}_{\mathbf{r};n}. It is well-known that for any field 𝕜\mathbbm{k} the ideal 𝒫𝐫;n\langle\mathscr{P}_{\mathbf{r};n}\rangle set-theoretically carves out the partial flag variety Fl𝐫;n(𝕜)\operatorname{Fl}_{\mathbf{r};n}(\mathbbm{k}) embedded in i=1k(𝕜([n]ri))\prod_{i=1}^{k}\mathbb{P}\left(\mathbbm{k}^{\binom{[n]}{r_{i}}}\right) 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 TrFl𝐫;n\operatorname{TrFl}_{\mathbf{r};n} of Fl𝐫;n\operatorname{Fl}_{\mathbf{r};n}, the nonnegative tropicalization TrFl𝐫;n0\operatorname{TrFl}_{\mathbf{r};n}^{\geq 0} of Fl𝐫;n\operatorname{Fl}_{\mathbf{r};n}, the flag Dressian FlDr𝐫;n\operatorname{FlDr}_{\mathbf{r};n}, and the nonnegative flag Dressian FlDr𝐫;n0\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0} are subsets of i=1k(𝕋([n]ri))\prod_{i=1}^{k}\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{r_{i}}}\right) defined as

TrFl𝐫;n\displaystyle\operatorname{TrFl}_{\mathbf{r};n} =f𝒫𝐫;nVtrop(f) and TrFl𝐫;n0=f𝒫𝐫;nVtrop0(f),\displaystyle=\bigcap_{f\in\langle\mathscr{P}_{\mathbf{r};n}\rangle}V_{\operatorname{trop}}(f)\quad\text{ and }\quad\operatorname{TrFl}_{\mathbf{r};n}^{\geq 0}=\bigcap_{f\in\langle\mathscr{P}_{\mathbf{r};n}\rangle}V_{\operatorname{trop}}^{\geq 0}(f),
FlDr𝐫;n\displaystyle\operatorname{FlDr}_{\mathbf{r};n} =f𝒫𝐫;nVtrop(f) and FlDr𝐫;n0=f𝒫𝐫;nVtrop0(f).\displaystyle=\bigcap_{f\in\mathscr{P}_{\mathbf{r};n}}V_{\operatorname{trop}}(f)\quad\text{ and }\quad\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0}=\bigcap_{f\in\mathscr{P}_{\mathbf{r};n}}V_{\operatorname{trop}}^{\geq 0}(f).

When k=1k=1, i.e. when 𝐫\mathbf{r} consists of one integer dd, one obtains the (nonnegative) tropicalization of the Grassmannian TrGrd;n(0)\operatorname{TrGr}_{d;n}^{(\geq 0)} and the (nonnegative) Dressian Drd;n(0)\operatorname{Dr}_{d;n}^{(\geq 0)} studied in [SS04, SW05, SW21, AHLS20]. Like Fln\operatorname{Fl}_{n}, we write only nn in the subscript when 𝐫=(1,2,,n)\mathbf{r}=(1,2,\ldots,n).

Remark 3.4.

In [JLLO, §6], the authors define the “positive flag Dressian” to consist of the elements 𝝁=(μ1,,μk)FlDr𝐫;n{\boldsymbol{\mu}}=(\mu_{1},\ldots,\mu_{k})\in\operatorname{FlDr}_{\mathbf{r};n} whose constituents μi\mu_{i} are each in the strictly positive Dressian. In our language, this is equal to considering the points of

fi=1k𝒫ri,ri;nVtrop0(f)fri<rj𝒫ri,rj;nVtrop(f)\bigcap_{f\in\bigcup_{i=1}^{k}\mathscr{P}_{r_{i},r_{i};n}}V_{\operatorname{trop}}^{\geq 0}(f)\cap\bigcap_{f\in\bigcup_{r_{i}<r_{j}}\mathscr{P}_{r_{i},r_{j};n}}V_{\operatorname{trop}}(f)

that have no \infty 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 𝒞={{t}}\mathcal{C}=\mathbb{C}\{\{t\}\} be the field of Puiseux series with coefficients in \mathbb{C}, with the usual valuation map val:𝒞𝕋\operatorname{val}:\mathcal{C}\to\mathbb{T}. Concretely, for f0f\neq 0, val(f)\operatorname{val}(f) is the exponent of the initial term of ff, and val(0)=\operatorname{val}(0)=\infty. Let

𝒞>0={f𝒞{0}:the initial coefficient of f is real and positive} and 𝒞0=𝒞>0{0}.\mathcal{C}_{>0}=\{f\in\mathcal{C}\setminus\{0\}:\text{the initial coefficient of $f$ is real and positive}\}\ \text{ and }\ \mathcal{C}_{\geq 0}=\mathcal{C}_{>0}\cup\{0\}.

For a point pFl𝐫;n(𝒞)i=1k(𝒞([n]ri))p\in\operatorname{Fl}_{\mathbf{r};n}(\mathcal{C})\subseteq\prod_{i=1}^{k}\mathbb{P}\left(\mathcal{C}^{\binom{[n]}{r_{i}}}\right), applying the valuation val:𝒞𝕋\operatorname{val}:\mathcal{C}\to\mathbb{T} coordinate-wise to the Plücker coordinates gives a point val(p)i=1k(𝕋([n]ri))\operatorname{val}(p)\in\prod_{i=1}^{k}\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{r_{i}}}\right). Noting that val(𝒞)={}𝕋\operatorname{val}(\mathcal{C})=\mathbb{Q}\cup\{\infty\}\subset\mathbb{T}, we say that a point in i=1k(𝕋([n]ri))\prod_{i=1}^{k}\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{r_{i}}}\right) has rational coordinates if it is a point in i=1k(({})([n]ri))\prod_{i=1}^{k}\mathbb{P}\left(\left(\mathbb{Q}\cup\{\infty\}\right)^{\binom{[n]}{r_{i}}}\right). Let Fl𝐫;n(𝒞0)\operatorname{Fl}_{\mathbf{r};n}(\mathcal{C}_{\geq 0}) be the subset of Fl𝐫;n(𝒞)\operatorname{Fl}_{\mathbf{r};n}(\mathcal{C}) consisting of points with all coordinates in 𝒞0\mathcal{C}_{\geq 0}, i.e. the points pFl𝐫;n(𝒞)i=1k(𝒞([n]ri))p\in\operatorname{Fl}_{\mathbf{r};n}(\mathcal{C})\subseteq\prod_{i=1}^{k}\mathbb{P}\left(\mathcal{C}^{\binom{[n]}{r_{i}}}\right) that have a representative in i=1k𝒞0([n]ri)\prod_{i=1}^{k}\mathcal{C}_{\geq 0}^{\binom{[n]}{r_{i}}}.

Proposition 3.6.

The set {val(p):pFl𝐫;n(𝒞)}\{\operatorname{val}(p):p\in\operatorname{Fl}_{\mathbf{r};n}(\mathcal{C})\} equals the set of points in TrFl𝐫;n\operatorname{TrFl}_{\mathbf{r};n} with rational coordinates. Likewise, the set {val(p):pFl𝐫;n(𝒞0)}\{\operatorname{val}(p):p\in\operatorname{Fl}_{\mathbf{r};n}(\mathcal{C}_{\geq 0})\} equals the set of points in TrFl𝐫;n0\operatorname{TrFl}_{\mathbf{r};n}^{\geq 0} with rational coordinates. Moreover, we have

TrFl𝐫;n\displaystyle\operatorname{TrFl}_{\mathbf{r};n} =the closure of {val(p):pFl𝐫;n(𝒞)} in i=1k(𝕋([n]ri))and\displaystyle=\text{the closure of }\{\operatorname{val}(p):p\in\operatorname{Fl}_{\mathbf{r};n}(\mathcal{C})\}\text{ in }\textstyle\prod_{i=1}^{k}\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{r_{i}}}\right)\quad\text{and}
TrFl𝐫;n0\displaystyle\operatorname{TrFl}_{\mathbf{r};n}^{\geq 0} =the closure of {val(p):pFl𝐫;n(𝒞0)} in i=1k(𝕋([n]ri)).\displaystyle=\text{the closure of }\{\operatorname{val}(p):p\in\operatorname{Fl}_{\mathbf{r};n}(\mathcal{C}_{\geq 0})\}\text{ in }\textstyle\prod_{i=1}^{k}\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{r_{i}}}\right).
Proof.

The first equality is known as the (extended) Fundamental Theorem of tropical geometry [MS15, Theorem 3.2.3 & Theorem 6.2.15]. The second equality is the analogue for nonnegative tropicalizations, established in [SW05, Proposition 2.2]. ∎

Remark 3.7.

The need to restrict to rational coordinates and the need to take the closure in 3.6 can be removed if we let 𝒞\mathcal{C} be the Mal’cev-Neumann ring (())\mathbb{C}((\mathbb{R})) (see [Poo93, §3]) which satisfies val(𝒞)=𝕋\operatorname{val}(\mathcal{C})=\mathbb{T}. See also [Mar10].

Let us also record an equivalent description of the (nonnegative) flag Dressian when 𝐫\mathbf{r} is a sequence of consecutive integers. We need the following definition. As is customary in matroid theory, we write SijSij for the union S{i,j}S\cup\{i,j\} of subsets SS and {i,j}\{i,j\} of [n][n].

Definition 3.8.

The set 𝒫r,r;n(3)\mathscr{P}_{r,r;n}^{(3)} of three-term Grassmann-Plücker relations (of type (r;n)(r;n)) is the subset of 𝒫r,r;n\mathscr{P}_{r,r;n} consisting of polynomials of the form

xSijxSkxSikxSj+xSixSjkx_{Sij}x_{Sk\ell}-x_{Sik}x_{Sj\ell}+x_{Si\ell}x_{Sjk}

for a subset S[n]S\subseteq[n] of cardinality r2r-2 and a subset {i<j<k<}[n]\{i<j<k<\ell\}\subseteq[n] disjoint from SS. Similarly, the set 𝒫r,r+1;n(3)\mathscr{P}_{r,r+1;n}^{(3)} of three-term incidence-Plücker relations (of type (r,r+1)(r,r+1)) is the subset of 𝒫r,r;n\mathscr{P}_{r,r;n} consisting of polynomials of the form

xSixSjkxSjxSik+xSkxSijx_{Si}x_{Sjk}-x_{Sj}x_{Sik}+x_{Sk}x_{Sij}

for a subset S[n]S\subseteq[n] of cardinality r1r-1 and a subset {i<j<k}[n]\{i<j<k\}\subseteq[n] disjoint from SS.

Let 𝒫𝐫;n(3)\mathscr{P}_{\mathbf{r};n}^{(3)} 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 𝐫=(r1<<rk)\mathbf{r}=(r_{1}<\dots<r_{k}) consists of consecutive integers. Then a point 𝝁=(μ1,,μk)i=1k(𝕋([n]ri)){\boldsymbol{\mu}}=(\mu_{1},\ldots,\mu_{k})\in\prod_{i=1}^{k}\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{r_{i}}}\right) is in the (nonnegative) flag Dressian if and only if its support 𝝁¯=(μ¯1,,μ¯k)\underline{{\boldsymbol{\mu}}}=(\underline{\mu}_{1},\ldots,\underline{\mu}_{k}) is a flag matroid and 𝝁{\boldsymbol{\mu}} satisfies the (nonnegative-)tropical three-term Plücker relations. More explicitly, we have

FlDr𝐫;n\displaystyle\operatorname{FlDr}_{\mathbf{r};n} ={𝝁i=1k(𝕋([n]ri))|𝝁¯ is a flag matroid and 𝝁f𝒫𝐫;n(3)Vtrop(f)}, and\displaystyle=\left\{{\boldsymbol{\mu}}\in\textstyle\prod_{i=1}^{k}\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{r_{i}}}\right)\ \ \middle|\ \ \underline{{\boldsymbol{\mu}}}\text{ is a flag matroid and }{\boldsymbol{\mu}}\in\bigcap_{f\in\mathscr{P}_{\mathbf{r};n}^{(3)}}V_{\operatorname{trop}}(f)\right\},\text{ and}
FlDr𝐫;n0\displaystyle\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0} ={𝝁i=1k(𝕋([n]ri))|𝝁¯ is a flag matroid and 𝝁f𝒫𝐫;n(3)Vtrop0(f)}.\displaystyle=\left\{{\boldsymbol{\mu}}\in\textstyle\prod_{i=1}^{k}\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{r_{i}}}\right)\ \ \middle|\ \ \underline{{\boldsymbol{\mu}}}\text{ is a flag matroid and }{\boldsymbol{\mu}}\in\bigcap_{f\in\mathscr{P}_{\mathbf{r};n}^{(3)}}V_{\operatorname{trop}}^{\geq 0}(f)\right\}.
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 𝕋\mathbb{T}\mathbb{R}, for which we note the following fact: The underlying set of 𝕋\mathbb{T}\mathbb{R} is (×{+,}){}(\mathbb{R}\times\{+,-\})\cup\{\infty\}, so given c𝕋c\in\mathbb{T}, one can identify it with the element (c,+)×{+,}(c,+)\in\mathbb{R}\times\{+,-\} of 𝕋\mathbb{T}\mathbb{R} if c<c<\infty and \infty otherwise.

In the language of hyperfields, for a homogeneous polynomial ff in mm variables and a hyperfield 𝔽\mathbb{F}, one has the notion of the “hypersurface of ff over 𝔽\mathbb{F},” which is a subset V𝔽(f)V_{\mathbb{F}}(f) of (𝔽m)\mathbb{P}(\mathbb{F}^{m}). When 𝔽\mathbb{F} is the tropical hyperfield 𝕋\mathbb{T}, this coincides with Vtrop(f)V_{\operatorname{trop}}(f) in 3.1. When 𝔽\mathbb{F} is the signed tropical hyperfield 𝕋\mathbb{T}\mathbb{R}, a point w𝕋mw\in\mathbb{T}^{m}, when considered as a point of 𝕋m\mathbb{T}\mathbb{R}^{m}, is in V𝕋(f)V_{\mathbb{T}\mathbb{R}}(f) if and only if it is in Vtrop0(f)V_{\operatorname{trop}}^{\geq 0}(f). Thus, in the language of flag matroids over hyperfields [JL22], the flag Dressian is the partial flag variety Fl𝐫;n(𝕋)\operatorname{Fl}_{\mathbf{r};n}(\mathbb{T}) over 𝕋\mathbb{T}, and the nonnegative flag Dressian is the subset of the partial flag variety Fl𝐫;n(𝕋)\operatorname{Fl}_{\mathbf{r};n}(\mathbb{T}\mathbb{R}) over 𝕋\mathbb{T}\mathbb{R} consisting of points that come from 𝕋\mathbb{T}.

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 𝐫\mathbf{r} consists of consecutive integers, for a perfect hyperfield 𝔽\mathbb{F}, a point pi=1k(𝔽([n]ri))p\in\prod_{i=1}^{k}\mathbb{P}\left(\mathbb{F}^{\binom{[n]}{r_{i}}}\right) is in the partial flag variety Fl𝐫;n(𝔽)\operatorname{Fl}_{\mathbf{r};n}(\mathbb{F}) over 𝔽\mathbb{F} if and only if the support of pp is a flag matroid and pp satisfies the three-term Plücker relations over 𝔽\mathbb{F}. ∎

For completeness, we include the proof of the following fact.

Lemma 3.10.

The signed tropical hyperfield 𝕋\mathbb{T}\mathbb{R} is doubly distributive. That is, for any x,y,z,w𝕋x,y,z,w\in\mathbb{T}\mathbb{R}, one has an equality of sets (xy)(zw)=xzxwyzyw(x\boxplus y)\cdot(z\boxplus w)=xz\boxplus xw\boxplus yz\boxplus yw.

Proof.

If any one of the four x,y,z,wx,y,z,w is \infty, then the desired equality is the usual distributivity of the signed tropical hyperfield. Thus, we now assume that all four elements are in ×{+,}\mathbb{R}\times\{+,-\}, and write x=(x,xSS)×{+,}x=(x_{\mathbb{R}},x_{\SS})\in\mathbb{R}\times\{+,-\} and similarly for y,z,wy,z,w. If x>yx_{\mathbb{R}}>y_{\mathbb{R}}, then xz>yzxz_{\mathbb{R}}>yz_{\mathbb{R}} and xw>ywxw_{\mathbb{R}}>yw_{\mathbb{R}}, so the equality follows again from the usual distributivity. So we now assume that all four elements have the same value in \mathbb{R}, and the equality then follows from the fact that the signed hyperfield SS\SS is doubly distributive. ∎

Remark 3.11.

Even when 𝐫\mathbf{r} 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 𝒫𝐫;n\mathscr{P}_{\mathbf{r};n} in the following way: Denoting by

𝒫𝐫;nadj=i=1k𝒫ri,ri;ni=1k1𝒫ri,ri+1;n,\mathscr{P}_{\mathbf{r};n}^{adj}=\bigcup_{i=1}^{k}\mathscr{P}_{r_{i},r_{i};n}\cup\bigcup_{i=1}^{k-1}\mathscr{P}_{r_{i},r_{i+1};n},

one has

FlDr𝐫;n=f𝒫𝐫;nadjVtrop(f)andFlDr𝐫;n0=f𝒫𝐫;nadjVtrop0(f).\operatorname{FlDr}_{\mathbf{r};n}=\bigcap_{f\in\mathscr{P}_{\mathbf{r};n}^{adj}}V_{\operatorname{trop}}(f)\quad\text{and}\quad\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0}=\bigcap_{f\in\mathscr{P}_{\mathbf{r};n}^{adj}}V_{\operatorname{trop}}^{\geq 0}(f).

This generalizes the fact that a sequence of matroids (M1,,Mk)(M_{1},\ldots,M_{k}) is a flag matroid if and only if (Mi,Mi+1)(M_{i},M_{i+1}) is a flag matroid for all i=1,,k1i=1,\ldots,k-1 [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 𝐫=(r1<<rk)\mathbf{r}=(r_{1}<\cdots<r_{k}) consists of consecutive integers, and suppose we have points 𝝁1,,𝝁ik𝕋([n]ri){\boldsymbol{\mu}}_{1},\ldots,{\boldsymbol{\mu}}_{\ell}\in\prod_{i}^{k}\mathbb{T}^{\binom{[n]}{r_{i}}} that are in FlDr𝐫;n0\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0}. Then, if a nonnegative linear combination c1𝝁1++c𝝁c_{1}{\boldsymbol{\mu}}_{1}+\cdots+c_{\ell}{\boldsymbol{\mu}}_{\ell} is in FlDr𝐫;n\operatorname{FlDr}_{\mathbf{r};n}, it is in FlDr𝐫;n0\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0}.

Proof.

We make the following general observation: Suppose f=cαxαcβxβ+cγxγf=c_{\alpha}x^{\alpha}-c_{\beta}x^{\beta}+c_{\gamma}x^{\gamma} is a three-term polynomial in [x1,,xm]\mathbb{R}[x_{1},\ldots,x_{m}] with cα,cβ,cγc_{\alpha},c_{\beta},c_{\gamma} positive. Then an element u𝕋mu\in\mathbb{T}^{m} satisfies the positive-tropical relation of ff if and only if βu=min{αu,γu}\beta\bullet u=\min\{\alpha\bullet u,\gamma\bullet u\}. Hence, if u1,,u𝕋mu_{1},\ldots,u_{\ell}\in\mathbb{T}^{m} each satisfy this relation, then a nonnegative linear combination of them can satisfy the tropical relation of ff only if the term at β\beta 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 𝐫\mathbf{r} be a sequence of consecutive integers (a,,b)(a,\ldots,b) for some 1abn1\leq a\leq b\leq n. We will show that TrFl𝐫;n0=FlDr𝐫;n0\operatorname{TrFl}_{\mathbf{r};n}^{\geq 0}=\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0}. The inclusion TrFl𝐫;n0FlDr𝐫;n0\operatorname{TrFl}_{\mathbf{r};n}^{\geq 0}\subseteq\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0} is immediate from 3.3. We will deduce TrFl𝐫;n0FlDr𝐫;n0\operatorname{TrFl}_{\mathbf{r};n}^{\geq 0}\supseteq\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0} by utilizing the two known cases of the equality TrFl𝐫;n0=FlDr𝐫;n0\operatorname{TrFl}_{\mathbf{r};n}^{\geq 0}=\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0} — when 𝐫=(r)\mathbf{r}=(r) and when 𝐫=(1,2,,n)\mathbf{r}=(1,2,\ldots,n).

We start by recalling that tropicalization behaves well on subtraction-free rational maps.

Definition 3.13.

Let f=a𝒜caxa[x1,,xm]f=\sum_{a\in\mathcal{A}}c_{a}x^{a}\in\mathbb{R}[x_{1},\ldots,x_{m}] be a real polynomial, where 𝒜\mathcal{A} is a finite subset of 0m\mathbb{Z}_{\geq 0}^{m} and 0ca0\neq c_{a}\in\mathbb{R}. We define the tropicalization Trop(f):m\operatorname{Trop}(f):\mathbb{R}^{m}\to\mathbb{R} to be the piecewise-linear map wmina𝒜(aw)w\mapsto\min_{a\in\mathcal{A}}(a\bullet w), where as before, aw=a1w1++amwma\bullet w=a_{1}w_{1}+\dots+a_{m}w_{m}.

Note that Trop(f1f2)=Trop(f1)+Trop(f2).\operatorname{Trop}({f_{1}}{f_{2}})=\operatorname{Trop}({f_{1}})+\operatorname{Trop}({f_{2}}). Moreover, if f1f_{1} and f2f_{2} are two polynomials with positive coefficients, and a1,a2>0a_{1},a_{2}\in\mathbb{R}_{>0}, then Trop(a1f1+a2f2)=min(Trop(f1),Trop(f2)).\operatorname{Trop}(a_{1}{f_{1}}+a_{2}{f_{2}})=\min(\operatorname{Trop}({f_{1}}),\operatorname{Trop}({f_{2}})). 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 f=(f1,,fn):𝒞m𝒞nf=(f_{1},\dots,f_{n}):{\mathcal{C}}^{m}\to\mathcal{C}^{n} be a rational map defined by polynomials f1,,fnf_{1},\dots,f_{n} with positive coefficients (or more generally by subtraction-free rational expressions). Let (x1,,xm)(𝒞0)m(x_{1},\dots,x_{m})\in(\mathcal{C}_{\geq 0})^{m}, such that f(x1,,xm)=(y1,,yn)f(x_{1},\dots,x_{m})=(y_{1},\dots,y_{n}). Then

(Trop(f))(val(x1),,val(xm))=(val(y1),,val(yn)).(\operatorname{Trop}(f))(\operatorname{val}(x_{1}),\dots,\operatorname{val}(x_{m}))=(\operatorname{val}(y_{1}),\dots,\operatorname{val}(y_{n})).

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 μdDrd;n0\mu_{d}\in\operatorname{Dr}_{d;n}^{\geq 0} with rational coordinates, there exists μd+1Drd+1;n0\mu_{d+1}\in\operatorname{Dr}_{d+1;n}^{\geq 0} such that (μd,μd+1)FlDrd,d+1;n0(\mu_{d},\mu_{d+1})\in\operatorname{FlDr}_{d,d+1;n}^{\geq 0}. Similarly, there exists μd1Drd1;n0\mu_{d-1}\in\operatorname{Dr}_{d-1;n}^{\geq 0} such that (μd1,μd)FlDrd1,d;n0(\mu_{d-1},\mu_{d})\in\operatorname{FlDr}_{d-1,d;n}^{\geq 0}.

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

Fln(𝒞0)=vwv,w(𝒞>0)\operatorname{Fl}_{n}(\mathcal{C}_{\geq 0})=\bigsqcup_{v\leq w}\mathcal{R}_{v,w}(\mathcal{C}_{>0})

where each cell v,w(𝒞>0)\mathcal{R}_{v,w}(\mathcal{C}_{>0}) can be parameterized using a map

ϕv,w:(𝒞>0)(w)(v)v,w(𝒞>0).\phi_{v,w}:(\mathcal{C}_{>0})^{\ell(w)-\ell(v)}\to\mathcal{R}_{v,w}(\mathcal{C}_{>0}).

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.

The first statement comes from [MR04, Theorem 11.3]; Marsh and Rietsch were working over \mathbb{R} and >0\mathbb{R}_{>0} but the same proof holds over Puiseux series. The statement that the parameterization can be expressed as an embedding into projective space using positive polynomials comes from [RW08, Proposition 5.1]. ∎

Corollary 3.17.

Each mm-dimensional positroid cell S(𝒞>0)S_{\mathcal{B}}(\mathcal{C}_{>0}) in the nonnegative Grassmannian Grd,n(𝒞0)\operatorname{Gr}_{d,n}(\mathcal{C}_{\geq 0}) is the projection πd(v,w(𝒞>0))\pi_{d}(\mathcal{R}_{v,w}(\mathcal{C}_{>0})) of some positive Richardson of dimension m=(w)(v)m=\ell(w)-\ell(v) in Fln(𝒞0)\operatorname{Fl}_{n}(\mathcal{C}_{\geq 0}), so we get a subtraction-free rational map

πdϕv,w:(𝒞>0)mv,w(𝒞>0)S(𝒞>0).\pi_{d}\circ\phi_{v,w}:(\mathcal{C}_{>0})^{m}\to\mathcal{R}_{v,w}(\mathcal{C}_{>0})\to S_{\mathcal{B}}(\mathcal{C}_{>0}).
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 μdDrd;n0\mu_{d}\in\operatorname{Dr}_{d;n}^{\geq 0} with rational coordinates implies that μd=val({ΔI(Vd)})\mu_{d}=\operatorname{val}(\{\Delta_{I}(V_{d})\}) for some subspace VdGrd,n(𝒞0)V_{d}\in\operatorname{Gr}_{d,n}(\mathcal{C}_{\geq 0}), and hence VdV_{d} lies in some positroid cell S(𝒞>0)S_{\mathcal{B}}(\mathcal{C}_{>0}) over Puiseux series.

By 3.17, VdV_{d} is the projection of a point (V1,,Vn)(V_{1},\dots,V_{n}) of Fln(𝒞0)\operatorname{Fl}_{n}(\mathcal{C}_{\geq 0}), which in turn is the image of a point (x1,,xm)(𝒞>0)m(x_{1},\dots,x_{m})\in(\mathcal{C}_{>0})^{m}, and the Plücker coordinates ΔI(Vj)\Delta_{I}(V_{j}) of each VjV_{j} are expressed as positive polynomials ΔI(x1,,xm)\Delta_{I}(x_{1},\dots,x_{m}) in the parameters x1,,xmx_{1},\dots,x_{m}.

In particular, we have subtraction-free maps

πdϕv,w:(𝒞>0)mFln(𝒞0)Grd,n(𝒞0)\pi_{d}\circ\phi_{v,w}:(\mathcal{C}_{>0})^{m}\to\operatorname{Fl}_{n}(\mathcal{C}_{\geq 0})\to\operatorname{Gr}_{d,n}(\mathcal{C}_{\geq 0})

taking

(x1,,xm){ΔI(x1,,xm)|I[n]}{ΔI(x1,,xm)|I([n]d)}.(x_{1},\dots,x_{m})\mapsto\{\Delta_{I}(x_{1},\dots,x_{m})\ |\ I\subset[n]\}\mapsto\left\{\Delta_{I}(x_{1},\dots,x_{m})\ |\ I\in{[n]\choose d}\right\}.

The fact that the maps ϕv,w\phi_{v,w} and πd\pi_{d} are subtraction-free implies by 3.14 that we can tropicalize them, obtaining maps

Trop(πdϕv,w):mTrFln0TrGrd,n0\operatorname{Trop}(\pi_{d}\circ\phi_{v,w}):\mathbb{R}^{m}\to\operatorname{TrFl}_{n}^{\geq 0}\to\operatorname{TrGr}_{d,n}^{\geq 0}

taking

(val(x1),,val(xm)){val(ΔI(x1,,xm))|I[n]}{val(ΔI(x1,,xm))|I([n]d)}.(\operatorname{val}(x_{1}),\dots,\operatorname{val}(x_{m}))\mapsto\{\operatorname{val}(\Delta_{I}(x_{1},\dots,x_{m}))\ |\ I\subset[n]\}\mapsto\left\{\operatorname{val}(\Delta_{I}(x_{1},\dots,x_{m}))\ |\ I\in{[n]\choose d}\right\}.

We now let μd+1={val(ΔI(Vd+1))|I([n]d+1)}\mu_{d+1}=\{\operatorname{val}(\Delta_{I}(V_{d+1}))\ |\ I\in{[n]\choose d+1}\} and μd1={val(ΔI(Vd1))|I([n]d1)}\mu_{d-1}=\{\operatorname{val}(\Delta_{I}(V_{d-1}))\ |\ I\in{[n]\choose d-1}\}. By construction we have that all the three-term (incidence) Plücker relations hold for (μd,μd+1)(\mu_{d},\mu_{d+1}), and similarly for (μd1,μd)(\mu_{d-1},\mu_{d}). Therefore (μd,μd+1)FlDrd,d+1;n0(\mu_{d},\mu_{d+1})\in\operatorname{FlDr}_{d,d+1;n}^{\geq 0} and (μd1,μd)FlDrd1,d;n0(\mu_{d-1},\mu_{d})\in\operatorname{FlDr}_{d-1,d;n}^{\geq 0}. ∎

The following consequence of 3.15 is very useful.

Corollary 3.18.

Let aabba^{\prime}\leq a\leq b\leq b^{\prime} be positive integers, and let 𝐫=(a,a+1,,b)\mathbf{r}=(a,a+1,\dots,b) and 𝐫=(a,a+1,,b)\mathbf{r}^{\prime}=(a^{\prime},a^{\prime}+1,\dots,b^{\prime}) be sequences of consecutive integers. Then any point (μa,,μb)FlDr𝐫;n0(\mu_{a},\dots,\mu_{b})\in\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0} with rational coordinates can be extended to a point (μa,μa+1,,μa,,μb,,μb)FlDr𝐫;n0(\mu_{a^{\prime}},\mu_{a^{\prime}+1},\dots,\mu_{a},\dots,\mu_{b},\dots,\mu_{b^{\prime}})\in\operatorname{FlDr}_{\mathbf{r}^{\prime};n}^{\geq 0}.

Proof.

We start with 𝝁=(μa,μa+1,,μb)FlDr𝐫;n0{\boldsymbol{\mu}}=(\mu_{a},\mu_{a+1},\dots,\mu_{b})\in\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0}. We take μb\mu_{b} and repeatedly use 3.15 to construct μb+1\mu_{b+1}, then μb+2,,μb\mu_{b+2},\dots,\mu_{b^{\prime}}. Similarly we take μa\mu_{a} and use 3.15 to construct μa1,μa2,,μa\mu_{a-1},\mu_{a-2},\dots,\mu_{a^{\prime}}. Now by construction (μa,μa+1,,μa,,μb,,μb)(\mu_{a^{\prime}},\mu_{a^{\prime}+1},\dots,\mu_{a},\dots,\mu_{b},\dots,\mu_{b^{\prime}}) satisfies:

  • μiDri;n0\mu_{i}\in\operatorname{Dr}_{i;n}^{\geq 0} for i=a,a+1,,bi=a^{\prime},a^{\prime}+1,\dots,b^{\prime};

  • all three-term incidence-Plücker relations hold (because the three-term incidence-Plücker relations occur only in consecutive ranks).

Therefore (μa,μa+1,,μb)FlDr𝐫;n0(\mu_{a^{\prime}},\mu_{a^{\prime}+1},\dots,\mu_{b^{\prime}})\in\operatorname{FlDr}_{\mathbf{r}^{\prime};n}^{\geq 0} by 3.9. ∎

Theorem 3.19.

Let 𝐫=(a,a+1,,b)\mathbf{r}=(a,a+1,\dots,b) be a sequence of consecutive integers, and let 𝝁FlDr𝐫;n0{\boldsymbol{\mu}}\in\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0} with rational coordinates. Then 𝝁TrFl𝐫;n0{\boldsymbol{\mu}}\in\operatorname{TrFl}_{\mathbf{r};n}^{\geq 0}.

Proof.

We start with 𝝁=(μa,μa+1,,μb)FlDr𝐫;n0{\boldsymbol{\mu}}=(\mu_{a},\mu_{a+1},\dots,\mu_{b})\in\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0}, and use 3.18 to construct (μ1,,μn)FlDrn0(\mu_{1},\dots,\mu_{n})\in\operatorname{FlDr}_{n}^{\geq 0}. Now [Bor22, Theorem 5.21trop] states that FlDrn0=TrFln0\operatorname{FlDr}_{n}^{\geq 0}=\operatorname{TrFl}_{n}^{\geq 0}. Hence, we have (μ1,,μn)TrFln0(\mu_{1},\dots,\mu_{n})\in\operatorname{TrFl}_{n}^{\geq 0}, so (μa,μa+1,,μb)TrFl𝐫,n0(\mu_{a},\mu_{a+1},\dots,\mu_{b})\in\operatorname{TrFl}_{\mathbf{r},n}^{\geq 0}. ∎

Proof of (a)\iff(b) in A.

We only need show that (b)\implies(a), i.e. that TrFl𝐫;n0FlDr𝐫;n0\operatorname{TrFl}_{\mathbf{r};n}^{\geq 0}\supseteq\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0}, since the other direction is trivial. But this follows from 3.19 because the points in FlDr𝐫;n0\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0} with rational coordinates are dense in FlDr𝐫;n0\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0}, and TrFl𝐫;n0\operatorname{TrFl}_{\mathbf{r};n}^{\geq 0} is closed. ∎

Remark 3.20.

Note that our method of proof crucially used the fact that 𝐫\mathbf{r} is a sequence of consecutive integers: we used 3.15 to fill in the ranks from bb through nn and from aa down to 11. But if say we were considering 𝐫={a,b}\mathbf{r}=\{a,b\} with ba>1b-a>1 and 𝝁=(μa,μb){\boldsymbol{\mu}}=(\mu_{a},\mu_{b}), we could not guarantee using 3.15 that we could construct μb1,μb2,,μa+1\mu_{b-1},\mu_{b-2},\dots,\mu_{a+1} in a way that is consistent with μa\mu_{a}.

Remark 3.21.

Recall from 2.3 that if 𝐫\mathbf{r} 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 𝐫=(a)\mathbf{r}=(a), which is to say, for the nonnegative Grassmannian [TW13, Corollary 1.2] and also when 𝐫=(1,2,,n)\mathbf{r}=(1,2,\ldots,n), which is to say, for the nonnegative complete flag variety [Bor22, Theorem 5.21]. To prove the result for 𝐫=(a,a+1,,b)\mathbf{r}=(a,a+1,\ldots,b), we start with a flag V=(Va,,Vb)V_{\bullet}=(V_{a},\dots,V_{b}) in ranks 𝐫\mathbf{r} whose Plücker coordinates are all nonnegative, so that VV_{\bullet} is Plücker nonnegative. As in 3.15, we can use the 𝐫=(a)\mathbf{r}=(a) 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 𝐫=(b)\mathbf{r}=(b) case. This yields a complete flag (V1,,Vn)(V_{1},\dots,V_{n}) with all nonnegative Plücker coordinates. We can then apply the result in the complete flag case to conclude that (V1,,Vn)(V_{1},\dots,V_{n}) lies in Fln0\operatorname{Fl}_{n}^{\geq 0}. Thus, VV_{\bullet} is a projection of the nonnegative complete flag (V1,,Vn)(V_{1},\dots,V_{n}) and itself lies in Fl𝐫;n0\operatorname{Fl}_{\mathbf{r};n}^{\geq 0}, which is to say, VV_{\bullet} is Lusztig nonnegative.

The strictly positive tropicalization of a partial flag variety TrFl𝐫;n>0\operatorname{TrFl}_{\mathbf{r};n}^{>0} is the subset of TrFl𝐫;n0\operatorname{TrFl}_{\mathbf{r};n}^{\geq 0} consisting of points whose coordinates are never \infty. Define similarly the strictly positive flag Dressian FlDr𝐫;n>0\operatorname{FlDr}_{\mathbf{r};n}^{>0}. The weaker version of 3.19 stating that TrFln>0=FlDrn>0\operatorname{TrFl}_{n}^{>0}=\operatorname{FlDr}_{n}^{>0} was established in [JLLO, Lemma 19] as follows. One starts by noting that if μDrr+m;n+m0\mu\in\operatorname{Dr}_{r+m;n+m}^{\geq 0}, then the sequence of minors (μr,,μr+m)(\mu_{r},\ldots,\mu_{r+m}) where μr+i=μ{n+1,,n+i}/{n+i+1,,n+m}\mu_{r+i}=\mu\setminus\{n+1,\ldots,n+i\}/\{n+i+1,\ldots,n+m\} is a point in FlDrr,,r+m;n0\operatorname{FlDr}_{r,\ldots,r+m;n}^{\geq 0}. Then, the crucial step is a construction in discrete convex analysis [MS18, Proposition 2] that shows that every element of FlDrn>0\operatorname{FlDr}_{n}^{>0} arises from an element of Drn;2n>0\operatorname{Dr}_{n;2n}^{>0} in this way. One then appeals to Grr;n>0=Drr;n>0\operatorname{Gr}_{r;n}^{>0}=\operatorname{Dr}_{r;n}^{>0} 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 FlDrn0\operatorname{FlDr}_{n}^{\geq 0} arises from an element of Drn;2n0\operatorname{Dr}_{n;2n}^{\geq 0} in such a way.

Example 3.22.

Let (M1,M2,M3)(M_{1},M_{2},M_{3}) be matroids on [3][3] whose sets of bases are ({1,3},{13},{123})(\{1,3\},\{13\},\{123\}). The matrix

[101001010]\begin{bmatrix}1&0&1\\ 0&0&1\\ 0&-1&0\end{bmatrix}

shows that it is a flag positroid. However, we claim that there is no positroid MM of rank 33 on [6][6] such that M1=M4/56M_{1}=M\setminus 4/56, M2=M45/6M_{2}=M\setminus 45/6, and M3=M456M_{3}=M\setminus 456. Since all three cases involve deletion by 4, if we replace M4M\setminus 4 by MM^{\prime}, and decrease each of 5,65,6 by 11, then we are claiming that there is no positroid MM^{\prime} of rank 3 on [5][5] such that

(4) M1=M/45,M2=M4/5,and M3=M45.M_{1}=M^{\prime}/45,\ M_{2}=M^{\prime}\setminus 4/5,\ \text{and }M_{3}=M^{\prime}\setminus 45.

From M1=M/45M_{1}=M^{\prime}/45 and M2=M4/5M_{2}=M^{\prime}\setminus 4/5, we have that M/5M^{\prime}/5 has bases {14,34,13}\{14,34,13\}, and similarly, we have M4M^{\prime}\setminus 4 has bases {135,123}\{135,123\}. Hence, the set of bases of MM^{\prime} contains {123,135,145,345}\{123,135,145,345\}, and does not contain {125,235,245}\{125,235,245\}. By considering the Plücker relation

p134p235=p123p345+p135p234,p_{134}p_{235}=p_{123}p_{345}+p_{135}p_{234},

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 a,,ba,\dots,b can be extended to ranks a,,ba^{\prime},\dots,b^{\prime} (for aaaba^{\prime}\leq a\leq a\leq b).

4.1. Oriented matroids and flag matroids

We give here a brief review of oriented matroids in terms of Plücker relations. Let SS={1,0,1}\SS=\{-1,0,1\} be the hyperfield of signs. For a polynomial f=a𝒜caxa[x1,,xm]f=\sum_{a\in\mathcal{A}}c_{a}x^{a}\in\mathbb{R}[x_{1},\ldots,x_{m}], we say that an element χSSm\chi\in\SS^{m} is in the null set of ff if the set {sign(ca)χa}a𝒜\{\operatorname{sign}(c_{a})\chi^{a}\}_{a\in\mathcal{A}} is either {0}\{0\} or contains {1,1}\{-1,1\}.

Definition 4.1.

An oriented matroid of rank rr on [n][n] is a point χSS([n]r)\chi\in\SS^{\binom{[n]}{r}}, called a chirotope, such that χ\chi is in the null set of ff for every f𝒫r,r;nf\in\mathscr{P}_{r,r;n}. Similarly, an oriented flag matroid of ranks 𝐫\mathbf{r} is a point 𝝌=(χ1,,χk)i=1kSS([n]ri)\boldsymbol{\chi}=(\chi_{1},\ldots,\chi_{k})\in\prod_{i=1}^{k}\SS^{\binom{[n]}{r_{i}}} such that 𝝌\boldsymbol{\chi} is in the null set of ff for every f𝒫𝐫;nf\in\mathscr{P}_{\mathbf{r};n}.

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 χ\chi such that χ\chi only takes values 0 or 1. Similarly, we define a positively oriented flag matroid to be an oriented flag matroid 𝝌\boldsymbol{\chi} such that 𝝌\boldsymbol{\chi} only takes values 0 or 1.

A positroid MM defines a positively oriented matroid χ=χM\chi=\chi_{M} where χ\chi 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 χ\chi is realizable, i.e. χ\chi has the form χM\chi_{M} for some positroid MM.

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 (χ1,,χk)(\chi_{1},\dots,\chi_{k}) of consecutive ranks r1<<rkr_{1}<\dots<r_{k} 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 (M,M)(M,M^{\prime}) be matroids of ranks 1 and 2 on [3][3] whose sets of bases are {1,3}\{1,3\} and {12,13,23}\{12,13,23\}, respectively. Both are positroids. We can realize (M,M)(M,M^{\prime}) as a flag matroid using the matrix

[a0bcde],\begin{bmatrix}a&0&b\\ c&d&e\end{bmatrix},

where the nonvanishing minors a,b,ad,bd,aebca,b,ad,-bd,ae-bc are nonzero. In order to realize (M,M)(M,M^{\prime}) as a flag positroid, we need to choose real numbers a,b,c,d,ea,b,c,d,e such that all these minors are strictly positive. However, a>0a>0 and ad>0ad>0 implies d>0d>0, while b>0b>0 and bd>0-bd>0 implies d<0d<0.

This example is consistent with 1.5 because (M,M)(M,M^{\prime}), 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 𝐫\mathbf{r} can be identified with the set of points of the nonnegative flag Dressian FlDr𝐫;n0\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0} whose coordinates are all either 0 or \infty.

Proof.

Given a point χ=(χ1,,χm){0,1}mSSm\chi=(\chi_{1},\ldots,\chi_{m})\in\{0,1\}^{m}\subset\SS^{m},222Note that (χ1,,χm)(\chi_{1},\dots,\chi_{m}) is not a sequence of chirotopes in this proof, instead each χiSS.\chi_{i}\in\SS. we define t(χ)=(t1,,tm)𝕋mt(\chi)=(t_{1},\ldots,t_{m})\in\mathbb{T}^{m} by setting ti=0t_{i}=0 if χi=1\chi_{i}=1 and ti=t_{i}=\infty if χi=0\chi_{i}=0. Then, we observe that χ\chi is in the null set of a polynomial f[x1,,xm]f\in\mathbb{R}[x_{1},\ldots,x_{m}] if and only if the image of t(χ)t(\chi) in (𝕋m)\mathbb{P}(\mathbb{T}^{m}) is a point in Vtrop0(f)V_{\operatorname{trop}}^{\geq 0}(f). Therefore, each positively oriented flag matroid 𝝌\boldsymbol{\chi} can be identified with the element t(𝝌)t(\boldsymbol{\chi}) in the nonnegative flag Dressian FlDr𝐫;n0\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0}. ∎

We now prove that every positively oriented flag matroid 𝝌=(χ1,,χk)\boldsymbol{\chi}=(\chi_{1},\dots,\chi_{k}) of consecutive ranks r1<<rkr_{1}<\dots<r_{k} is realizable.

Proof of 1.5.

By the lemma, we may identify a positively oriented flag matroid 𝝌\boldsymbol{\chi} as an element t(𝝌)t(\boldsymbol{\chi}) of the nonnegative flag Dressian. Because the ranks 𝐫\mathbf{r} are consecutive integers, the equivalence (a)\iff(b) of A implies that t(𝝌)t(\boldsymbol{\chi}) is thus a point in TrFl𝐫;n0\operatorname{TrFl}_{\mathbf{r};n}^{\geq 0}. Because t(𝝌)t(\boldsymbol{\chi}) has rational coordinates (all non-\infty coordinates are 0), 3.6 implies that t(𝝌)=val(p)t(\boldsymbol{\chi})=\operatorname{val}(p) for some pi=1k(𝒞0([n]ri))p\in\prod_{i=1}^{k}\mathbb{P}\left(\mathcal{C}_{\geq 0}^{\binom{[n]}{r_{i}}}\right). Setting the parameter tt in each Puisseux series of pp to 0 now gives the realization of 𝝌\boldsymbol{\chi} as a flag positroid. ∎

As in 1.6, we do not know whether the corollary holds when 𝐫\mathbf{r} 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 (M,M)(M,M^{\prime}) on [4][4] of ranks (1,3)(1,3) such that there is no flag positroid (M,M2,M)(M,M_{2},M^{\prime}) with rank of M2M_{2} equal to 2. Let the sets of bases of MM and MM^{\prime} be {1,2,3,4}\{1,2,3,4\} and {123,234}\{123,234\}, respectively. The matrix

[111101000010]\begin{bmatrix}1&1&1&1\\ 0&1&0&0\\ 0&0&1&0\end{bmatrix}

for example shows that (M,M)(M,M^{\prime}) 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 (M,M)(M,M^{\prime}) as a flag positroid, after row-reducing by the first row, is of the form

[1abc0xy00zw0]\begin{bmatrix}1&a&b&c\\ 0&x&y&0\\ 0&z&w&0\end{bmatrix}

where a,b,c>0a,b,c>0 and xwyz>0xw-yz>0. The minors of the matrix formed by the first two rows include x,y,cx,cyx,y,-cx,-cy, which cannot be all nonnegative since c>0c>0 and not both of xx and yy are zero.

Remark 4.7.

Let us sketch an alternate proof of 1.5 that relies only on the weaker version of (a)\iff(b) in A that the strictly positive parts agree, i.e. that TrFl𝐫;n>0=FlDr𝐫;n>0\operatorname{TrFl}_{\mathbf{r};n}^{>0}=\operatorname{FlDr}_{\mathbf{r};n}^{>0}. For a matroid MM of rank dd, define ρM([n]d)\rho_{M}\in\mathbb{R}^{\binom{[n]}{d}} by ρM(S)=drkM(S)\rho_{M}(S)=d-\operatorname{rk}_{M}(S) for S([n]d)S\in\binom{[n]}{d}, where rkM\operatorname{rk}_{M} is the rank function of MM. If MM is a positively oriented matroid, then ρM\rho_{M} is a point in the positive Dressian Drd,n>0\operatorname{Dr}^{>0}_{d,n} [SW21, Proof of Theorem 5.1]. One can use this to show that if 𝑴=(M1,,Mk){\boldsymbol{M}}=(M_{1},\ldots,M_{k}) is a positively oriented flag matroid of consecutive ranks 𝐫\mathbf{r}, then the sequence 𝝆=(ρM1,,ρMk)\boldsymbol{\rho}=(\rho_{M_{1}},\ldots,\rho_{M_{k}}) is a point in FlDr𝐫;n>0\operatorname{FlDr}_{\mathbf{r};n}^{>0}. Since TrFl𝐫;n>0=FlDr𝐫;n>0\operatorname{TrFl}_{\mathbf{r};n}^{>0}=\operatorname{FlDr}_{\mathbf{r};n}^{>0} and 𝝆\boldsymbol{\rho} has rational coordinates, 3.6 implies that there is a point pFl𝐫;n(𝒞0)p\in\operatorname{Fl}_{\mathbf{r};n}(\mathcal{C}_{\geq 0}) with val(p)=𝝆\operatorname{val}(p)=\boldsymbol{\rho}. Consider the coordinate p(S)𝒞p(S)\in\mathcal{C} of pp at a subset S([n]ri)S\in\binom{[n]}{r_{i}}. By construction, the initial term of p(S)p(S) is ctqct^{q} for some positive real cc and a nonnegative integer qq, where qq is zero exactly when SS is a basis of MiM_{i}. Thus, setting the parameter tt to 0 in the Puisseux series of pp gives a realization of 𝑴{\boldsymbol{M}} as a flag positroid.

We now use 4.5 to give a matroidal analogue of 3.18.

Corollary 4.8.

Let aabba^{\prime}\leq a\leq b\leq b^{\prime} be positive integers, and let (Ma,Ma+1,,Mb)(M_{a},M_{a+1},\dots,M_{b}) be a positively oriented flag matroid on [n][n] of consecutive ranks a,a+1,,ba,a+1,\dots,b, that is, a sequence of positroids Ma,,MbM_{a},\dots,M_{b} which is also an oriented flag matroid. Then we can extend it to a positively oriented flag matroid (Ma,Ma+1,,Ma,,Mb,,Mb)(M_{a^{\prime}},M_{a^{\prime}+1},\dots,M_{a},\dots,M_{b},\dots,M_{b^{\prime}}) of consecutive ranks a,a+1,,ba^{\prime},a^{\prime}+1,\dots,b^{\prime}.

Proof.

As in 4.5, we view the positively oriented flag matroid (Ma,,Mb)(M_{a},\dots,M_{b}) as a point of the nonnegative flag Dressian (μa,,μb)FlDr𝐫;n0(\mu_{a},\dots,\mu_{b})\in\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0} whose coordinates are all either 0 or \infty. The desired statement almost follows from 3.15: we just need to check that we can extend (μa,,μb)(\mu_{a},\dots,\mu_{b}) in a way which preserves the fact that coordinates are all either 0 or \infty. This is true, and we prove it by following the proof of 3.15 and replacing all instances of the positive Puiseux series 𝒞>0\mathcal{C}_{>0} by the positive Puiseux series with constant coefficients, that is, by >0\mathbb{R}_{>0}. Alternatively, we can use our result that (Ma,,Mb)(M_{a},\dots,M_{b}) 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 𝝁=(μ1,,μk)i=1k(𝕋([n]ri)){\boldsymbol{\mu}}=(\mu_{1},\dots,\mu_{k})\in\prod_{i=1}^{k}\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{r_{i}}}\right) such that its support 𝝁¯\underline{{\boldsymbol{\mu}}} is a flag matroid. By construction, the vertices of the flag matroid polytope P(𝝁¯)P(\underline{{\boldsymbol{\mu}}}) have the form 𝐞B1++𝐞Bk\mathbf{e}_{B_{1}}+\cdots+\mathbf{e}_{B_{k}} where BiB_{i} is a basis of the matroid μ¯i\underline{\mu}_{i} for each i=1,,ki=1,\ldots,k.

Definition 5.1.

We define 𝒟𝝁\mathcal{D}_{\boldsymbol{\mu}} to be the coherent subdivision of P(𝝁¯)P(\underline{{\boldsymbol{\mu}}}) induced by assigning each vertex 𝐞B1++𝐞Bk\mathbf{e}_{B_{1}}+\cdots+\mathbf{e}_{B_{k}} of P(𝝁¯)P(\underline{{\boldsymbol{\mu}}}) the weight μ1(B1)++μk(Bk)\mu_{1}(B_{1})+\cdots+\mu_{k}(B_{k}). That is, the faces of 𝒟𝝁\mathcal{D}_{\boldsymbol{\mu}} correspond to the faces of the lower convex hull of the set of points

{(𝐞B1++𝐞Bk,μ1(B1)++μk(Bk))n×:𝐞B1++𝐞Bk a vertex of P(𝝁)}.\{(\mathbf{e}_{B_{1}}+\cdots+\mathbf{e}_{B_{k}},\mu_{1}(B_{1})+\cdots+\mu_{k}(B_{k}))\in\mathbb{R}^{n}\times\mathbb{R}:\mathbf{e}_{B_{1}}+\cdots+\mathbf{e}_{B_{k}}\text{ a vertex of $P({\boldsymbol{\mu}})$}\}.

The points of the flag Dressians are exactly the ones for which the subdivision 𝒟𝝁\mathcal{D}_{\boldsymbol{\mu}} consists of flag matroid polytopes.

Theorem 5.2.

[BEZ21, Theorem A.(a)&(c)] A point 𝝁i=1k(𝕋([n]ri)){\boldsymbol{\mu}}\in\prod_{i=1}^{k}\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{r_{i}}}\right) is in the flag Dressian FlDr𝐫;n\operatorname{FlDr}_{\mathbf{r};n} if and only if the all faces of the subdivision 𝒟𝝁\mathcal{D}_{\boldsymbol{\mu}} are flag matroid polytopes.

When 𝐫\mathbf{r} consists of consecutive integers (a,a+1,,b)(a,a+1,\ldots,b), the nonnegative analogue of this theorem is the equivalence of (b) and (c) in A, which states that a point 𝝁i=ab(𝕋([n]i)){\boldsymbol{\mu}}\in\prod_{i=a}^{b}\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{i}}\right) is in the nonnegative flag Dressian FlDr𝐫;n0\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0} if and only if all faces of the subdivision 𝒟𝝁\mathcal{D}_{\boldsymbol{\mu}} are flag positroid polytopes. A different nonnegative analogue of 5.2 that holds for 𝐫\mathbf{r} not necessarily consecutive, but loses the flag positroid property, can be found in 5.6.

5.2. The proof of (b)\implies(c)\implies(d)\implies(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 MM on [n][n] and a vector 𝐯n\mathbf{v}\in\mathbb{R}^{n}, let face(P(M),𝐯)\operatorname{face}(P(M),{\mathbf{v}}) be the face of the matroid polytope P(M)P(M) that maximizes the standard pairing with 𝐯\mathbf{v}.

Proposition 5.3.

Let MM be a matroid on [n][n] and let 𝒮=(S1S[n])\mathscr{S}=(\emptyset\subsetneq S_{1}\subsetneq\cdots\subsetneq S_{\ell}\subsetneq[n]) be a chain of nonempty proper subsets of [n][n]. For a vector 𝐯𝒮\mathbf{v}_{\mathscr{S}} in the relative interior of the cone 0{𝐞S1,,𝐞S}\mathbb{R}_{\geq 0}\{\mathbf{e}_{S_{1}},\ldots,\mathbf{e}_{S_{\ell}}\}, we have

face(P(M),𝐯𝒮)=P(M𝒮),\operatorname{face}(P(M),{\mathbf{v}_{\mathscr{S}}})=P(M^{\mathscr{S}}),

where M𝒮=M|S1M|S2/S1M|S3/S2M/SM^{\mathscr{S}}=M|S_{1}\oplus M|S_{2}/S_{1}\oplus M|S_{3}/S_{2}\oplus\cdots\oplus M/S_{\ell} is the direct sum of minors of MM.

For 𝑴=(M1,,Mk){\boldsymbol{M}}=(M_{1},\ldots,M_{k}) a flag matroid, since P(𝑴)P({\boldsymbol{M}}) is the Minkowski sum P(M1)++P(Mk)P(M_{1})+\cdots+P(M_{k}), we likewise have that face(P(𝑴),𝐯𝒮)=P(𝑴𝒮)=P(M1𝒮)++P(Mk𝒮)\operatorname{face}(P({\boldsymbol{M}}),\mathbf{v}_{\mathscr{S}})=P({\boldsymbol{M}}^{\mathscr{S}})=P(M_{1}^{\mathscr{S}})+\cdots+P(M_{k}^{\mathscr{S}}), where 𝑴𝒮=(M1𝒮,,Mk𝒮){\boldsymbol{M}}^{\mathscr{S}}=(M_{1}^{\mathscr{S}},\ldots,M_{k}^{\mathscr{S}}). 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 w𝕋([n]r)w\in\mathbb{T}^{\binom{[n]}{r}}, its support w¯\underline{w} is {S([n]r):wS}\{S\in\binom{[n]}{r}:w_{S}\neq\infty\}.

  • We consider a point w𝕋([n]r)w\in\mathbb{T}^{\binom{[n]}{r}} as a set of weights on the vertices {𝐞S:Sw¯}\{\mathbf{e}_{S}:S\in\underline{w}\} of P(w¯)nP(\underline{w})\subset\mathbb{R}^{n}. Given an affine-linear function φ:n\varphi:\mathbb{R}^{n}\to\mathbb{R} and an element w𝕋([n]r)w\in\mathbb{T}^{\binom{[n]}{r}}, we define

    φw𝕋([n]r)by(φw)(S)=φ(𝐞S)+w(S) for S([n]r).\varphi w\in\mathbb{T}^{\binom{[n]}{r}}\quad\text{by}\quad(\varphi w)(S)=\varphi(\mathbf{e}_{S})+w(S)\text{ for $S\in\binom{[n]}{r}$}.
  • For a point w𝕋([n]r)w\in\mathbb{T}^{\binom{[n]}{r}}, denote by win𝕋([n]r)w^{\mathrm{in}}\in\mathbb{T}^{\binom{[n]}{r}} its initial part, i.e.

    win(S)={0if w(S)=min{w(S):S([n]r)}otherwise.w^{\mathrm{in}}(S)=\begin{cases}0&\text{if $w(S)=\min\{w(S^{\prime}):S^{\prime}\in\binom{[n]}{r}$}\}\\ \infty&\text{otherwise}.\end{cases}
Proposition 5.4.

Let 𝐫=(r1,,rk)\mathbf{r}=(r_{1},\dots,r_{k}) be a sequence of increasing integers in [n][n]. Suppose 𝒘=(w1,,wk)FlDr𝐫;n0\boldsymbol{w}=(w_{1},\ldots,w_{k})\in\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0}. Then, the following hold.

  1. (1)

    The support 𝒘¯\underline{\boldsymbol{w}} is a positively oriented flag matroid. In particular, it is a flag positroid when 𝐫=(r1,,rk)\mathbf{r}=(r_{1},\ldots,r_{k}) consists of consecutive integers.

  2. (2)

    We have φ𝒘=(φw1,,φwk)FlDr𝐫;n0\varphi\boldsymbol{w}=(\varphi w_{1},\ldots,\varphi w_{k})\in\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0} for any affine-linear functional φ\varphi on n\mathbb{R}^{n}.

  3. (3)

    We have 𝒘in=(w1in,,wkin)FlDr𝐫;n0\boldsymbol{w}^{\mathrm{in}}=(w_{1}^{\mathrm{in}},\ldots,w_{k}^{\mathrm{in}})\in\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0}.

Proof.

We may consider 𝒘¯\underline{\boldsymbol{w}} as an element i=1k(𝕋([n]ri))\prod_{i=1}^{k}\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{r_{i}}}\right) by assigning the value 0 to a subset SS if it is in the support of 𝒘\boldsymbol{w} and \infty otherwise. Then, we have 𝒘¯FlDr𝐫;n0\underline{\boldsymbol{w}}\in\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0} because the terms in each of the tropical Plücker relations that achieve the minimum when evaluated at 𝒘\boldsymbol{w} continue to do so when evaluated at 𝒘¯\underline{\boldsymbol{w}}. The statement (1) follows from 4.5 and 1.5

The support is unchanged by φ\varphi, so φ𝒘¯\underline{\varphi\boldsymbol{w}} is a flag matroid. The statement (2) now follows because for each of the positive-tropical Plücker relations, the operation φ\varphi preserves the terms at which the minimum is achieved.

The support 𝒘in¯\underline{\boldsymbol{w}^{\mathrm{in}}} is a flag matroid by 5.2 and because P(𝒘in¯)P(\underline{\boldsymbol{w}^{\mathrm{in}}}) is a face in the subdivision 𝒟𝒘\mathcal{D}_{\boldsymbol{w}} of P(𝒘¯)P(\underline{\boldsymbol{w}}). 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 \infty. ∎

Remark 5.5.

While it’s not needed here, we note that 5.4 is the “positive” analogue of the following statement, which is proved similarly: If 𝒘FlDr𝐫;n\boldsymbol{w}\in\operatorname{FlDr}_{\mathbf{r};n}, then (1) 𝒘¯\underline{\boldsymbol{w}} is a flag matroid, (2) φ𝒘FlDr𝐫;n\varphi\boldsymbol{w}\in\operatorname{FlDr}_{\mathbf{r};n}, and (3) 𝒘inFlDr𝐫;n\boldsymbol{w}^{\rm{in}}\in\operatorname{FlDr}_{\mathbf{r};n}. See also [BEZ21, Corollary 4.3.2] for related statements.

Proof of (b)\implies(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 𝒟𝝁\mathcal{D}_{\boldsymbol{\mu}} of P(𝝁¯)P(\underline{{\boldsymbol{\mu}}}) is the flag matroid polytope of a positively oriented flag matroid.

Similar argument as above shows that (b)\implies(c’) even when 𝐫\mathbf{r} doesn’t consist of consecutive integers. One can also verify the converse (c’)\implies(b) in this more general case as follows:

Suppose for contradiction (c’) but not (b) for some 𝝁{\boldsymbol{\mu}}. Then 5.2 implies that 𝝁{\boldsymbol{\mu}} 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 𝝁{\boldsymbol{\mu}} by φ𝝁\varphi{\boldsymbol{\mu}} for some φ\varphi if necessary, we may conclude that the same is true for that Plücker relation evaluated at 𝝁in{\boldsymbol{\mu}}^{\rm{in}}. But then 𝝁in{\boldsymbol{\mu}}^{\rm{in}}, 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.

The implication (c)\implies(d) is immediate.

Proof of (d)\implies(b).

First, assumption (d) implies that every edge of the subdivision 𝒟𝝁\mathcal{D}_{{\boldsymbol{\mu}}} of P(𝝁¯)P(\underline{{\boldsymbol{\mu}}}) is a flag matroid polytope, i.e. it is parallel to 𝐞i𝐞j\mathbf{e}_{i}-\mathbf{e}_{j} for some ij[n]i\neq j\in[n] and its two vertices are equidistant from the origin. Hence the edges of P(𝝁¯)P(\underline{{\boldsymbol{\mu}}}) have the same property, so 𝝁¯\underline{{\boldsymbol{\mu}}} 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 a=ba=b, where 𝝁{\boldsymbol{\mu}} is just (μ)(\mu). We need check the validity of the three-term positive-tropical Grassmann-Plücker relations, say for an arbitrary choice of S([n]a2)S\in\binom{[n]}{a-2} and {i<j<k<}[n]S\{i<j<k<\ell\}\subseteq[n]\setminus S. If SS is not independent in the matroid μ¯\underline{\mu}, then every term in the three-term relation involving SS and ijkijk\ell is \infty, so we may assume SS is independent. Let 𝒮\mathscr{S} be a maximal chain S1SmS_{1}\subsetneq\cdots\subsetneq S_{m} of subsets of [n][n] with the property that Sa2=SS_{a-2}=S and Sa1=S{ijk}S_{a-1}=S\cup\{ijk\ell\}. Then, Proposition 5.3 implies that for a vector 𝐯𝒮\mathbf{v}_{\mathscr{S}} in the relative interior of the cone 0{𝐞S1,,𝐞Sm}\mathbb{R}_{\geq 0}\{\mathbf{e}_{S_{1}},\ldots,\mathbf{e}_{S_{m}}\}, we have

face(P(μ¯),𝐯𝒮)=P(μ¯𝒮)P(μ¯|Sijk/S).\operatorname{face}(P(\underline{\mu}),\mathbf{v}_{\mathscr{S}})=P(\underline{\mu}^{\mathscr{S}})\simeq P(\underline{\mu}|S\cup ijk\ell/S).

For the second identification, we have used that

  1. (1)

    the matroid polytope of a direct sum of matroids is the product of the matroid polytopes;

  2. (2)

    with the exception of (Sa2,Sa1)=(S,Sijk)(S_{a-2},S_{a-1})=(S,S\cup ijk\ell), all other minors of the matroid μ¯\underline{\mu} corresponding to (Sc,Sc+1)(S_{c},S_{c+1}) in the chain have their polytopes being a point because |Sc+1Sc|=1|S_{c+1}\setminus S_{c}|=1.

Since SS is assumed to be independent, the rank of the matroid minor μ¯|Sijk/S\underline{\mu}|S\cup ijk\ell/S is at most 22. If it is less than 2, then every term in the three-term relation involving SS and ijkijk\ell is \infty, so let us now treat the case when the rank is exactly 2. For a basis B^\widehat{B} of μ¯|Sijk/S\underline{\mu}|S\cup ijk\ell/S, let BB be the basis of μ¯\underline{\mu} such that the vertex 𝐞B\mathbf{e}_{B} of P(μ¯)P(\underline{\mu}) corresponds to the vertex 𝐞B^\mathbf{e}_{\widehat{B}} of P(μ¯|Sijk/S)P(\underline{\mu}|S\cup ijk\ell/S) under the identification above. Identifying [4]={1<2<3<4}[4]=\{1<2<3<4\} with {i<j<k<}\{i<j<k<\ell\}, we may thus consider “restricting” μ\mu to the face P(μ¯|Sijk/S)P(\underline{\mu}|S\cup ijk\ell/S) to obtain an element μ^=μ|Sijk/SDr2;4\widehat{\mu}=\mu|S\cup ijk\ell/S\in\operatorname{Dr}_{2;4} defined by

μ^(B^)={μ(B) if B^ a basis of μ¯|Sijk/Sotherwisefor B^([4]2).\widehat{\mu}(\widehat{B})=\begin{cases}\mu(B)&\text{ if $\widehat{B}$ a basis of $\underline{\mu}|S\cup ijk\ell/S$}\\ \infty&\text{otherwise}\end{cases}\qquad\text{for $\textstyle\widehat{B}\in\binom{[4]}{2}$}.

It is straightforward to check that for Dr2;4\operatorname{Dr}_{2;4}, 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 𝒟μ^\mathcal{D}_{\widehat{\mu}} of P(μ¯|Sijk/S)P(\underline{\mu}|S\cup ijk\ell/S) are a subset of the faces of the subdivision 𝒟μ\mathcal{D}_{\mu}, we have that μ\mu satisfies the three-term tropical-positive Grassmann-Plücker relation involving ijkijk\ell and SS.

Let us now treat the case a<ba<b. That the three-term Grassmann-Plücker relations are satisfied for every μi\mu_{i} where i=a,,bi=a,\dotsc,b follows from our previous case of a=ba=b once we show the following claim:

For a flag matroid 𝝁¯\underline{{\boldsymbol{\mu}}} with consecutive rank sequence (a,,b)(a,\dotsc,b), if every face of P(𝝁¯)P(\underline{{\boldsymbol{\mu}}}) of dimension at most 2 is a flag positroid polytope, then the same holds for every constituent matroid, i.e. for every c=a,,bc=a,\dotsc,b, every face of P(μ¯c)P(\underline{\mu}_{c}) of dimension at most 2 is a positroid polytope .

To prove the claim, suppose for some acba\leq c\leq b that a 2-dimensional face QQ of P(μ¯c)P(\underline{\mu}_{c}) is not a positroid polytope. Our goal is to use QQ to find a 22-dimensional face of P(𝝁¯)P(\underline{{\boldsymbol{\mu}}}) 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 𝐞Sij,𝐞Sk,𝐞Si,𝐞Sjk\mathbf{e}_{Sij},\mathbf{e}_{Sk\ell},\mathbf{e}_{Si\ell},\mathbf{e}_{Sjk}, where S[n]S\subset[n] with |S|=c2|S|=c-2 and {i<j<k<}[n]S\{i<j<k<\ell\}\subset[n]\setminus S; thus Q=conv(𝐞Sij,𝐞Sk,𝐞Si,𝐞Sjk)Q=\operatorname{conv}(\mathbf{e}_{Sij},\mathbf{e}_{Sk\ell},\mathbf{e}_{Si\ell},\mathbf{e}_{Sjk}) for such {S,i,j,k,l}\{S,i,j,k,l\}333One may also deduce this independently of [LPW20] by using the argument given in the first third of this proof of (d)\implies(b) concerning the a=ba=b case.. Note that this 2-face QQ is the Minkowski sum of 𝐞S\mathbf{e}_{S} with the product conv(𝐞i,𝐞k)×conv(𝐞j,𝐞)\operatorname{conv}(\mathbf{e}_{i},\mathbf{e}_{k})\times\operatorname{conv}(\mathbf{e}_{j},\mathbf{e}_{\ell}).

Let 𝒮\mathscr{S} be a maximal chain S1SmS_{1}\subsetneq\dotsb\subsetneq S_{m} of subsets of [n][n] with the property that Sc1=SS_{c-1}=S, Sc=SikS_{c}=S\cup ik, and Sc+1=SijkS_{c+1}=S\cup ijk\ell. Then, Proposition 5.3 implies that for a vector v𝒮v_{\mathscr{S}} in the relative interior of the cone 0{𝐞S1,,𝐞Sm}\mathbb{R}_{\geq 0}\{\mathbf{e}_{S_{1}},\ldots,\mathbf{e}_{S_{m}}\}, we have

face(P(𝝁¯),𝐯𝒮)=P(𝝁¯𝒮)P(𝝁¯|Sc+1/Sc)×P(𝝁¯|Sc/Sc1).\operatorname{face}(P(\underline{{\boldsymbol{\mu}}}),\mathbf{v}_{\mathscr{S}})=P(\underline{{\boldsymbol{\mu}}}^{\mathscr{S}})\simeq P(\underline{{\boldsymbol{\mu}}}|S_{c+1}/S_{c})\times P(\underline{{\boldsymbol{\mu}}}|S_{c}/S_{c-1}).

For the second identification, we have used that

  1. (1)

    the matroid polytope of a direct sum of matroids is the product of the matroid polytopes;

  2. (2)

    with the exception of (Sc1,Sc)=(S,Sik)(S_{c-1},S_{c})=(S,S\cup ik) and (Sc,Sc+1)=(Sik,Sijkl)(S_{c},S_{c+1})=(S\cup ik,S\cup ijkl), all other minors of the constituent matroids of 𝝁¯\underline{{\boldsymbol{\mu}}} corresponding to (Sd,Sd+1)(S_{d},S_{d+1}) in the chain have their polytopes being a point because |Sd+1Sd|=1|S_{d+1}\setminus S_{d}|=1.

Note that the polytope P(𝝁¯|Sc+1/Sc)×P(𝝁¯|Sc/Sc1)P(\underline{{\boldsymbol{\mu}}}|S_{c+1}/S_{c})\times P(\underline{{\boldsymbol{\mu}}}|S_{c}/S_{c-1}) is at most 2-dimensional since 𝝁¯|Sc+1/Sc\underline{{\boldsymbol{\mu}}}|S_{c+1}/S_{c} and 𝝁¯|Sc/Sc1\underline{{\boldsymbol{\mu}}}|S_{c}/S_{c-1} are flag matroids on ground sets {j,}\{j,\ell\} and {i,k}\{i,k\}, respectively. The polytope has QQ 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 S[n]S\subset[n] with a1|S|b2a-1\leq|S|\leq b-2 and {i<j<k}[n]S\{i<j<k\}\subseteq[n]\setminus S. We may assume that SS has rank |S||S| in the matroid μ|S|+1\mu_{|S|+1}, since otherwise every term in the three-term positive-tropical incidence relation is \infty, so that the relation is vacuously satisfied. Let 𝒮\mathscr{S} be a maximal chain S1SmS_{1}\subsetneq\cdots\subsetneq S_{m} of subsets of [n][n] with the property that Sc=SS_{c}=S and Sc+1=SijkS_{c+1}=S\cup ijk for c=|S|c=|S|. Then, Proposition 5.3 implies that for a vector 𝐯𝒮\mathbf{v}_{\mathscr{S}} in the relative interior of the cone 0{𝐞S1,,𝐞Sm}\mathbb{R}_{\geq 0}\{\mathbf{e}_{S_{1}},\ldots,\mathbf{e}_{S_{m}}\}, we have

face(P(𝝁¯),𝐯𝒮)=P(𝝁¯𝒮)P(𝝁¯|Sijk/S).\operatorname{face}(P(\underline{{\boldsymbol{\mu}}}),\mathbf{v}_{\mathscr{S}})=P(\underline{{\boldsymbol{\mu}}}^{\mathscr{S}})\simeq P(\underline{{\boldsymbol{\mu}}}|S\cup ijk/S).

For the second identification, we have used that

  1. (1)

    the matroid polytope of a direct sum of matroids is the product of the matroid polytopes;

  2. (2)

    with the exception of (Sc,Sc+1)=(S,Sijk)(S_{c},S_{c+1})=(S,S\cup ijk), all other minors of the constituent matroids of 𝝁¯\underline{{\boldsymbol{\mu}}} corresponding to (Sd,Sd+1)(S_{d},S_{d+1}) in the chain have their polytopes being a point because |Sd+1Sd|=1|S_{d+1}\setminus S_{d}|=1.

Note that the polytope P(𝝁¯|Sijk/S)P(\underline{{\boldsymbol{\mu}}}|S\cup ijk/S) is at most 2-dimensional since it is a flag matroid polytope on 3 elements. Similarly to the a=ba=b case, we may “restrict” 𝝁¯\underline{{\boldsymbol{\mu}}} to the face P(𝝁¯|Sijk/S)P(\underline{{\boldsymbol{\mu}}}|S\cup ijk/S) to obtain an element 𝝁^=𝝁|Sijk/SFlDr𝒓^;3\widehat{\boldsymbol{\mu}}={\boldsymbol{\mu}}|S\cup ijk/S\in\operatorname{FlDr}_{\widehat{\boldsymbol{r}};3}. We may assume that 𝒓^=(1,2)\widehat{\boldsymbol{r}}=(1,2) since otherwise every term in the three-term incidence relation of the pair (S,ijk)(S,ijk) is \infty. For FlDr3\operatorname{FlDr}_{3}, it is straightforward to verify that the unique three-term positive-tropical incidence relation involving SS and ijkijk is satisfied if and only if the subdivision 𝒟𝝁^\mathcal{D}_{\widehat{\boldsymbol{\mu}}} consists only of flag positroid polytopes. Since the faces of the subdivision 𝒟𝝁^\mathcal{D}_{\widehat{\boldsymbol{\mu}}} are a subset of the faces of the subdivision 𝒟𝝁\mathcal{D}_{{\boldsymbol{\mu}}}, we have that 𝝁{\boldsymbol{\mu}} satisfies the three-term incidence relation involving SS and {i,j,k}\{i,j,k\}. ∎

6. Three-term incidence relations

6.1. The proof of (e)\iff(b) in A

In the case that a=ba=b in A, the equivalence (e)\iff(b) is the content of 3.9.

To prove the implication when a<ba<b, we will show the following key theorem.

Theorem 6.1.

Suppose 𝝁=(μ1,μ2)(𝕋([n]r))×(𝕋([n]r+1))\boldsymbol{\mu}=(\mu_{1},\mu_{2})\in\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{r}}\right)\times\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{r+1}}\right) satisfies every three-term positive-tropical incidence relation, and suppose that the support 𝝁¯\underline{\boldsymbol{\mu}} is a flag matroid. Then, we have 𝝁FlDrr,r+1;n0\boldsymbol{\mu}\in\operatorname{FlDr}^{\geq 0}_{r,r+1;n} if either of the following (incomparable) conditions hold:

  1. (i)

    The support 𝝁¯\underline{{\boldsymbol{\mu}}} consists of uniform matroids.

  2. (ii)

    Either μ1Drr;n0\mu_{1}\in\operatorname{Dr}^{\geq 0}_{r;n} or μ2Drr+1;n0\mu_{2}\in\operatorname{Dr}^{\geq 0}_{r+1;n}.

Proof of (b)\iff(e).

By Proposition 3.9, the implication (b)\implies(e) is immediate. For the converse, since 𝐫\mathbf{r} consists of consecutive integers, if 𝝁¯\underline{{\boldsymbol{\mu}}} is a flag matroid and 𝝁{\boldsymbol{\mu}} satisfies every three-term positive-tropical incidence relation, then 𝝁{\boldsymbol{\mu}} 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 𝝁{\boldsymbol{\mu}} is an element of FlDr𝐫;n0\operatorname{FlDr}_{\mathbf{r};n}^{\geq 0} by 3.9. ∎

The proof of 6.1 relies on the following technical lemma.

Lemma 6.2.

Suppose w𝕋([5]2)w\in\mathbb{T}^{\binom{[5]}{2}} satisfies all three-term positive-tropical Grassmann-Plücker relations involving the element 5. Suppose moreover that wi5<w_{i5}<\infty for some i=1,2,3,4i=1,2,3,4. Then wDr2;50w\in\operatorname{Dr}_{2;5}^{\geq 0}, i.e. ww 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 Gr2,5\operatorname{Gr}_{2,5}, if we can invert certain Plücker coordinates, then we can write the three-term Grassmann-Plücker relation not involving 55 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 p250p_{25}\neq 0 (respectively, p350p_{35}\neq 0) then p13p24p12p34p14p23p_{13}p_{24}-p_{12}p_{34}-p_{14}p_{23} can be written in the following ways.

p13p24p12p34p14p23\displaystyle p_{13}p_{24}-p_{12}p_{34}-p_{14}p_{23}
=(p13p25p12p35p15p23)p24p25(p14p25p12p45p15p24)p23p25+(p24p35p23p45p25p34)p12p25\displaystyle=(p_{13}p_{25}-p_{12}p_{35}-p_{15}p_{23})\frac{p_{24}}{p_{25}}-(p_{14}p_{25}-p_{12}p_{45}-p_{15}p_{24})\frac{p_{23}}{p_{25}}+(p_{24}p_{35}-p_{23}p_{45}-p_{25}p_{34})\frac{p_{12}}{p_{25}}
=(p13p25p12p35p15p23)p34p35(p14p35p13p45p15p34)p23p35+(p24p35p23p45p25p34)p13p35.\displaystyle=(p_{13}p_{25}-p_{12}p_{35}-p_{15}p_{23})\frac{p_{34}}{p_{35}}-(p_{14}p_{35}-p_{13}p_{45}-p_{15}p_{34})\frac{p_{23}}{p_{35}}+(p_{24}p_{35}-p_{23}p_{45}-p_{25}p_{34})\frac{p_{13}}{p_{35}}.

We next note that we can interpret the first (respectively, second) expression in 6.3 tropically as long as w25<w_{25}<\infty (respectively, w35<w_{35}<\infty).

Case 1: w25<.w_{25}<\infty. 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 55 hold, and w25<w_{25}<\infty, we have

w13+w25+w24w25\displaystyle w_{13}+w_{25}+w_{24}-w_{25} =min(w12+w35+w24w25,w15+w23+w24w25)\displaystyle=\min(w_{12}+w_{35}+w_{24}-w_{25},w_{15}+w_{23}+w_{24}-w_{25})
w14+w25+w23w25\displaystyle w_{14}+w_{25}+w_{23}-w_{25} =min(w12+w45+w23w25,w15+w24+w23w25)\displaystyle=\min(w_{12}+w_{45}+w_{23}-w_{25},w_{15}+w_{24}+w_{23}-w_{25})
w24+w35+w12w25\displaystyle w_{24}+w_{35}+w_{12}-w_{25} =min(w23+w45+w12w25,w25+w34+w12w25).\displaystyle=\min(w_{23}+w_{45}+w_{12}-w_{25},w_{25}+w_{34}+w_{12}-w_{25}).

We now simplify these expressions and underline terms that agree, obtaining:

(5) w13+w24\displaystyle w_{13}+w_{24} =min(w12+w35+w24w25¯,w15+w23+w24w25¯¯)\displaystyle=\min(\underline{w_{12}+w_{35}+w_{24}-w_{25}},\underline{\underline{w_{15}+w_{23}+w_{24}-w_{25}}})
(6) w14+w23\displaystyle w_{14}+w_{23} =min(w12+w45+w23w25,w15+w24+w23w25¯¯)\displaystyle=\min(\uwave{w_{12}+w_{45}+w_{23}-w_{25}},\underline{\underline{w_{15}+w_{24}+w_{23}-w_{25}}})
(7) w24+w35+w12w25¯\displaystyle\underline{w_{24}+w_{35}+w_{12}-w_{25}} =min(w23+w45+w12w25,w34+w12).\displaystyle=\min(\uwave{w_{23}+w_{45}+w_{12}-w_{25}},w_{34}+w_{12}).

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 w13+w24=w12+w34w14+w23w_{13}+w_{24}=w_{12}+w_{34}\leq w_{14}+w_{23}. In the other six cases, we find that w13+w24=w14+w23w12+w34w_{13}+w_{24}=w_{14}+w_{23}\leq w_{12}+w_{34}. Therefore the positive tropical Plücker relation involving 1,2,3,41,2,3,4 is satisfied.

Case 2: w35<.w_{35}<\infty. 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: w25=w35=.w_{25}=w_{35}=\infty. In this case, since 55 is not a loop, either w15<w_{15}<\infty or w45<.w_{45}<\infty. Suppose that w15<w_{15}<\infty. Then the positive tropical Plücker relations

  • w13+w25=min(w12+w35,w15+w23)w_{13}+w_{25}=\min(w_{12}+w_{35},w_{15}+w_{23})

  • w14+w25=min(w12+w45,w15+w24)w_{14}+w_{25}=\min(w_{12}+w_{45},w_{15}+w_{24})

  • w14+w35=min(w13+w45,w15+w34)w_{14}+w_{35}=\min(w_{13}+w_{45},w_{15}+w_{34})

imply that w23=w24=w34=w_{23}=w_{24}=w_{34}=\infty, and hence the positive tropical Plücker relation involving 1,2,3,41,2,3,4 is satisfied. The case where w45<w_{45}<\infty is similar. ∎

For w𝕋([n]r)w\in\mathbb{T}^{\binom{[n]}{r}}, define its dual w𝕋([n]nr)w^{\perp}\in\mathbb{T}^{\binom{[n]}{n-r}} by w(I)=w([n]I)w^{\perp}(I)=w([n]\setminus I). It is straightforward to verify that ww is an element of Drr;n\operatorname{Dr}_{r;n} (resp. Drr;n0\operatorname{Dr}_{r;n}^{\geq 0}) if and only if ww^{\perp} is an element of Drnr;n\operatorname{Dr}_{n-r;n} (resp. Drnr;n0\operatorname{Dr}_{n-r;n}^{\geq 0}). This matroid duality gives the following dual formulation of 6.2.

Corollary 6.4.

Suppose w𝕋([5]3)w\in\mathbb{T}^{\binom{[5]}{3}} satisfies all three-term positive-tropical Grassmann-Plücker relations that contain a variable indexed by S([5]3)S\in\binom{[5]}{3} with 5S5\notin S. If w¯\underline{w} is a matroid such that 5 is not a coloop, then wDr3;50w\in\operatorname{Dr}_{3;5}^{\geq 0}, i.e. ww 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 𝝁=(μ1,μ2)(𝕋([n]r))×(𝕋([n]r+1))\boldsymbol{\mu}=(\mu_{1},\mu_{2})\in\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{r}}\right)\times\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{r+1}}\right), define 𝝁~(𝕋([n+1]r+1))\widetilde{\boldsymbol{\mu}}\in\mathbb{P}\left(\mathbb{T}^{\binom{[n+1]}{r+1}}\right) by

𝝁~(S)={μ1(S(n+1))if (n+1)Sμ2(S)otherwise.\widetilde{\boldsymbol{\mu}}(S)=\begin{cases}\mu_{1}(S\setminus(n+1))&\text{if $(n+1)\in S$}\\ \mu_{2}(S)&\text{otherwise}.\end{cases}

Because 𝝁¯\underline{\boldsymbol{\mu}} is a flag matroid, we have that 𝝁¯~\underline{\widetilde{\boldsymbol{\mu}}} is a matroid, with the element (n+1)(n+1) that is neither a loop nor a coloop. We observe that 𝝁~Drr+1;n+10\widetilde{\boldsymbol{\mu}}\in\operatorname{Dr}_{r+1;n+1}^{\geq 0} if and only if 𝝁FlDrr,r+1;n0\boldsymbol{\mu}\in\operatorname{FlDr}^{\geq 0}_{r,r+1;n} because the validity of the three-term positive-tropical Grassmann-Plücker relations for 𝝁~\widetilde{\boldsymbol{\mu}} is equivalent to the validity of both the three-term positive-tropical incidence relations and the three-term positive-tropical Grassmann-Plücker relations for 𝝁\boldsymbol{\mu}.

We need to check that 𝝁~\widetilde{\boldsymbol{\mu}} satisfies every three-term positive-tropical Grassmann-Plücker relation of type (r+1;n+1)(r+1;n+1). Consider the three-term relation associated to the subset S[n+1]S\subseteq[n+1] of cardinality r1r-1 and {i<j<k<}[n+1]\{i<j<k<\ell\}\subseteq[n+1] disjoint from SS. We have three cases:

  • =n+1\ell=n+1. In this case, erasing the index n+1n+1 in the expression for the corresponding three-term Grassmann-Plücker relation yields a three-term incidence relation of type (r,r+1;n)(r,r+1;n), which is satisfied by our assumption on 𝝁{\boldsymbol{\mu}}.

  • (n+1)S(n+1)\in S. In this case, if (n+1)(n+1) is not a coloop in the minor 𝝁~|Sijk/(S(n+1))\widetilde{\boldsymbol{\mu}}|S\cup ijk\ell/(S\setminus(n+1)), then applying 6.4 to 𝝁~|Sijk/(S(n+1))\widetilde{\boldsymbol{\mu}}|S\cup ijk\ell/(S\setminus(n+1)) implies that the three-term Grassmann-Plücker relation is satisfied.

  • (n+1)Sijk(n+1)\notin S\cup ijk\ell. In this case, if (n+1)(n+1) is not a loop in the minor 𝝁~|Sijk(n+1)/S\widetilde{\boldsymbol{\mu}}|S\cup ijk\ell(n+1)/S, then applying 6.2 to 𝝁~|Sijk(n+1)/S\widetilde{\boldsymbol{\mu}}|S\cup ijk\ell(n+1)/S implies that the three-term Grassmann-Plücker relation is satisfied.

Under condition (i) of Theorem 6.1, i.e. when the support 𝝁¯\underline{{\boldsymbol{\mu}}} consists of uniform matroids, the element (n+1)(n+1) is not a coloop in the minor 𝝁~|Sijk/(S(n+1))\widetilde{\boldsymbol{\mu}}|S\cup ijk\ell/(S\setminus(n+1)), and is not a loop in the minor 𝝁~|Sijk(n+1)/S\widetilde{\boldsymbol{\mu}}|S\cup ijk\ell(n+1)/S. 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 (n+1)(n+1) is a loop in the minor 𝝁~|Sijk(n+1)/S\widetilde{\boldsymbol{\mu}}|S\cup ijk\ell(n+1)/S, 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 (n+1)(n+1) is not a loop in the matroid 𝝁¯~\underline{\widetilde{\boldsymbol{\mu}}}, (n+1)(n+1) belongs to the closure (also called span) in 𝝁¯~\underline{\widetilde{\boldsymbol{\mu}}} of SS. Since SS is also independent, there is an element sSs\in S such that (Ss)(n+1)(S\setminus s)\cup(n+1) is independent and has the same closure as SS in 𝝁¯~\underline{\widetilde{\boldsymbol{\mu}}}. Let S=SsS^{\prime}=S\setminus s. For any a,b{i,j,k,}a,b\in\{i,j,k,\ell\}, by our choice of sSs\in S, we have that SabSab is a basis of 𝝁¯~\underline{\widetilde{\boldsymbol{\mu}}} if and only if Sab(n+1)S^{\prime}ab(n+1) is a basis of 𝝁¯~\underline{\widetilde{\boldsymbol{\mu}}}. Moreover, for any a,b,c{i,j,k,}a,b,c\in\{i,j,k,\ell\} such that the values involved below are finite, we claim

𝝁~(Sab)𝝁~(Sac)=𝝁~(Sab(n+1))𝝁~(Sac(n+1)).\widetilde{\boldsymbol{\mu}}(Sab)-\widetilde{\boldsymbol{\mu}}(Sac)=\widetilde{\boldsymbol{\mu}}(S^{\prime}ab(n+1))-\widetilde{\boldsymbol{\mu}}(S^{\prime}ac(n+1)).

Note that using the definition of 𝝁~\widetilde{\boldsymbol{\mu}}, the above claim can be equivalently written as

μ2(Sab)μ2(Sac)=μ1(Sab)μ1(Sac).\mu_{2}(Sab)-\mu_{2}(Sac)=\mu_{1}(S^{\prime}ab)-\mu_{1}(S^{\prime}ac).

From the claim, we conclude as follows. Let μ¯1\overline{\mu}_{1} be the projection of μ1\mu_{1} to the coordinates labelled by SxyS^{\prime}xy where xy{i,j,k,}x\neq y\in\{i,j,k,\ell\}, and let μ¯2\overline{\mu}_{2} be the projection of μ2\mu_{2} to the coordinates labelled by SxySxy where xy{i,j,k,}x\neq y\in\{i,j,k,\ell\}. Then, as elements of (𝕋({i,j,k,}2))\mathbb{P}(\mathbb{T}^{\binom{\{i,j,k,\ell\}}{2}}), the two tropical vectors μ¯1\overline{\mu}_{1} and μ¯2\overline{\mu}_{2} are equal. Hence, the claim implies that if one of μ1\mu_{1} or μ2\mu_{2} 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

{μ1(Sab)+μ2(Sasc),μ1(Sas)+μ2(Sabc),μ1(Sac)+μ2(Sasb)},\{\mu_{1}(S^{\prime}ab)+\mu_{2}(S^{\prime}asc),\mu_{1}(S^{\prime}as)+\mu_{2}(S^{\prime}abc),\mu_{1}(S^{\prime}ac)+\mu_{2}(S^{\prime}asb)\},

from which the claim follows because Sa(n+1)Sa(n+1) is not a basis of 𝝁¯~\underline{\widetilde{\boldsymbol{\mu}}}, forcing μ1(Sas)=\mu_{1}(S^{\prime}as)=\infty. ∎

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 (u,v)(u,v) of permutations with uvu\leq v. Meanwhile, by work of Postnikov [Pos], positroid cells of Grd,n0\operatorname{Gr}_{d,n}^{\geq 0} 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 Grd,n0\operatorname{Gr}_{d,n}^{\geq 0}

As discussed in Section 2.1, there are two equivalent ways of thinking about the positroid cell decomposition of Grd,n0\operatorname{Gr}_{d,n}^{\geq 0}:

Grd,n0=S>0=u,vπ(u,v>0).\operatorname{Gr}_{d,n}^{\geq 0}=\bigsqcup S_{\mathcal{B}}^{>0}=\bigsqcup_{u,v}\pi(\mathcal{R}_{u,v}^{>0}).

In the union on the right, π\pi is the projection from Fln\operatorname{Fl}_{n} to Grd,n\operatorname{Gr}_{d,n}, and u,vu,v range over all permutations uvu\leq v in SnS_{n}, such that vv is a minimal-length coset representative of W/WdW/W_{d}, and Wd=s1,,sd1,s^d,sd+1,,sn1.W_{d}=\langle s_{1},\dots,s_{d-1},\hat{s}_{d},s_{d+1},\dots,s_{n-1}\rangle. We write WdW^{d} for the set of minimal-length coset representatives of W/WdW/W_{d}. Recall that a descent of a permutation zz is a position jj such that z(j)>z(j+1)z(j)>z(j+1). We have that WdW^{d} is the subset of permutations in SnS_{n} which have at most one descent, and if it exists, that descent must be in position dd.

Even if vWdv\notin W^{d}, the projection of u,v>0\mathcal{R}_{u,v}^{>0} to Grd,n0\operatorname{Gr}_{d,n}^{\geq 0} is still a positroid, which we will characterize below. We start by defining Grassmann necklaces [Pos].

Definition 7.1.

Let =(I1,,In)\mathcal{I}=(I_{1},\ldots,I_{n}) be a sequence of subsets of ([n]d){[n]\choose d}. We say \mathcal{I} is a Grassmann necklace of type (d,n)(d,n) if the following holds:

  • If iIii\in I_{i}, then Ii+1=(Iii)jI_{i+1}=(I_{i}\setminus i)\cup j for some j[n]j\in[n].

  • If iIii\notin I_{i}, then Ii+1=IiI_{i+1}=I_{i}.

In order to define the bijection between these Grassmann necklaces and positroids, we need to define the ii-Gale order on ([n]d)\binom{[n]}{d}.

Definition 7.2.

We write <i<_{i} for the following shifted linear order on [n][n].

i<ii+1<i<in<i1<i<ii1.i<_{i}i+1<_{i}\ldots<_{i}n<_{i}1<_{i}\ldots<_{i}i-1.

We also define the ii-Gale order on dd-element subsets by setting

{a1<i<iad}i{b1<i<ibd}\{a_{1}<_{i}\dots<_{i}a_{d}\}\leq_{i}\{b_{1}<_{i}\dots<_{i}b_{d}\}

if and only if aiba_{\ell}\leq_{i}b_{\ell} for all 1d1\leq\ell\leq d.

Given a positroid MM, we define a sequence M=(I1,,In)\mathcal{I}_{M}=(I_{1},\dots,I_{n}) of subsets of [n][n] by letting Ii{I}_{i} be the minimal basis of MM in the ii-Gale order. The following result is from [Pos, Theorem 17.1].

Proposition 7.3.

For any positroid MM, M\mathcal{I}_{M} is a Grassmann necklace. The map MMM\mapsto\mathcal{I}_{M} gives a bijection between positroids of rank dd on [n][n] and Grassmann necklaces of type (d,n)(d,n).

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 F=F1F2FkF=F_{1}\subset F_{2}\subset\dots\subset F_{k} on [n][n] is an increasing sequence of finite subsets of [n][n]. A flag matroid is a collection \mathcal{F} of flags satisfying the Maximality Property. Recall that eSe_{S} denotes the 0101 indicator vector in n\mathbb{R}^{n} associated to a subset S[n]S\subset[n]. For a flag F=F1F2FkF=F_{1}\subset F_{2}\subset\dots\subset F_{k} we let eF=eF1++eFke_{F}=e_{F_{1}}+\dots+e_{F_{k}}. In this language, the flag matroid polytope of \mathcal{F} is P=Conv{eF|F}P_{\mathcal{F}}=\operatorname{Conv}\{e_{F}\ |\ F\in\mathcal{F}\}, whose vertices are precisely the points eFe_{F} for FF\in\mathcal{F}.

In the complete flag case, each point eFe_{F} is a permutation vector (z(1),,z(n))(z(1),\dots,z(n)) for some zSnz\in S_{n}. Note that we can read off z:=z(F)z:=z(F) from FF by setting z(i)=jz(i)=j, where jj is the unique element of FiFi1F_{i}\setminus F_{i-1}.

Given uvu\leq v in Bruhat order, we define the Bruhat interval flag matroid u,v\mathcal{F}_{u,v} to be the complete flag matroid whose flags are precisely

{z([1])z([2])z([n])} for uzv,\{z([1])\subset z([2])\subset\dots\subset z([n])\}\text{ for }u\leq z\leq v,

where [i][i] denotes {1,2,,i}\{1,2,\dots,i\} and z([i])z([i]) denotes {z(1),,z(i)}\{z(1),\dots,z(i)\}. Then by the above discussion, the (twisted) Bruhat interval polytope

P~u,v=Conv{(n+1z1(1),n+1z1(2),,n+1z1(n))|uzv}\tilde{P}_{u,v}=\operatorname{Conv}\{(n+1-z^{-1}(1),n+1-z^{-1}(2),\dots,n+1-z^{-1}(n))\ |\ u\leq z\leq v\}

is the flag matroid polytope of the Bruhat interval flag matroid u,v\mathcal{F}_{u,v}.

This observation leads naturally to the following definition.

Definition 7.4.

Consider a complete flag matroid \mathcal{F} on [n][n], which we identify with a collection 𝒮\mathcal{S} of permutations on [n][n]. By the Maximality Property [BGW03, Section 1.7.2] and its relation to the tableau criterion for Bruhat order [BGW03, Theorem 5.17.3], 𝒮\mathcal{S} contains a unique permutation uu (respectively, vv) which is minimal (respectively, maximal) in Bruhat order among all elements of 𝒮\mathcal{S}. We say that u,v\mathcal{F}_{u,v} is the Bruhat interval envelope of \mathcal{F}.

It follows from 7.4 that the Bruhat interval envelope of a complete flag matroid \mathcal{F} contains \mathcal{F}; however, in general this inclusion is strict. It is an equality precisely when \mathcal{F} is a Bruhat interval flag matroid.

Recall that if F=(F1,,Fn)F=(F_{1},\dots,F_{n}) and G=(G1,,Gn)G=(G_{1},\dots,G_{n}) are flags, we say that FF is less than or equal to GG in the j\leq_{j} Gale order (and write FjGF\leq_{j}G) if and only if FijGiF_{i}\leq_{j}G_{i} for all 1in1\leq i\leq n. (We also talk about the “usual” Gale order with respect to the total order 1<2<<n1<2<\dots<n.) The Maximality Property for flag matroids implies that for any flag matroid \mathcal{F}, there is always a unique element which is maximal (and a unique element which is minimal) with respect to j\leq_{j}.

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 𝑴=(M1,,Mn){\boldsymbol{M}}=(M_{1},\ldots,M_{n}) on [n][n], that is, the flag positroid associated to any point of u,v>0\mathcal{R}_{u,v}^{>0}, for some uvu\leq v. For each 1jn1\leq j\leq n, let z(j)z^{(j)} be the Gale-minimal permutation with respect to j\leq_{j} in the interval [u,v][u,v]. Then the Grassmann necklace of the positroid MjM_{j} is (z(1)([j]),z(2)([j]),,z(n)([j]))(z^{(1)}([j]),z^{(2)}([j]),\dots,z^{(n)}([j])).

Example 7.6.

Consider the flag positroid associated to a point of u,v>0\mathcal{R}_{u,v}^{>0}, where u=(1,2,4,3)u=(1,2,4,3) and v=(4,2,1,3)v=(4,2,1,3) (which we abbreviate as 1243 and 4213). The interval [u,v][u,v] consists of

[u,v]={1243,1423,2143,2413,4123,4213}.[u,v]=\{1243,1423,2143,2413,4123,4213\}.

We now use 7.5, and find that the Gale-minimal permutations of [u,v][u,v] with respect to 1,2,3,4\leq_{1},\leq_{2},\leq_{3},\leq_{4} are 1243,2413,4123,41231243,2413,4123,4123. Therefore the Grassmann necklaces for the constituent positroids M1,M2,M3M_{1},M_{2},M_{3} and M4M_{4} are (1,2,4,4)(1,2,4,4), (12,24,14,14)(12,24,14,14), (124,124,124,124)(124,124,124,124), and (1234,1234,1234,1234)(1234,1234,1234,1234).

Alternatively, we can read off the flags in the flag positroid from the permutations in [u,v][u,v], obtaining the flags

{112124,114124,212124,224124,414124,424124}.\{1\subset 12\subset 124,1\subset 14\subset 124,2\subset 12\subset 124,2\subset 24\subset 124,4\subset 14\subset 124,4\subset 24\subset 124\}.

(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 M1,M2,M3,M4M_{1},M_{2},M_{3},M_{4} from the flags, obtaining {1,2,4}\{1,2,4\}, {12,14,24}\{12,14,24\}, {124}\{124\}, and {1234}\{1234\}. 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 M1,M2,M3,M4M_{1},M_{2},M_{3},M_{4} above, we obtain the twisted Bruhat interval polytope P~1243,4213=P2314,4312\tilde{P}_{1243,4213}=P_{2314,4312}, whose vertices are

{(4,3,1,2),(4,2,1,3),(3,4,1,2),(3,2,1,4),(2,4,1,3),(2,3,1,4)},\{(4,3,1,2),(4,2,1,3),(3,4,1,2),(3,2,1,4),(2,4,1,3),(2,3,1,4)\},

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 u,v>0\mathcal{R}_{u,v}^{>0}, where u,vSnu,v\in S_{n} and uvu\leq v in Bruhat order. Choose 1dn1\leq d\leq n. Let π\pi denote the projection from Fln\operatorname{Fl}_{n} to Grd,n\operatorname{Gr}_{d,n}. Then the bases of the rank dd positroid represented by π(u,v>0)\pi(\mathcal{R}_{u,v}^{>0}) are {z([d])|uzv}.\{z([d])\ |\ u\leq z\leq v\}.

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 MiM_{i} and Mi+1M_{i+1} on [n][n] of ranks ii and i+1i+1 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 MiM_{i} and Mi+1M_{i+1} on [n][n] of ranks ii and i+1i+1 form an oriented matroid quotient if (Mi,Mi+1)(M_{i},M_{i+1}) is an oriented flag matroid.

The following statement is a direct consequence of 4.8.

Proposition 7.9.

Let MiM_{i} and Mi+1M_{i+1} be positroids on [n][n] of ranks ii and i+1i+1. Then there is a complete flag positroid with MiM_{i} and Mi+1M_{i+1} as constituents if and only if (Mi,Mi+1)(M_{i},M_{i+1}) form an oriented matroid quotient.

Proposition 7.10.

Suppose that (M1,,Mn)(M_{1},\dots,M_{n}) is a sequence of positroids of ranks 1,2,,n1,2,\dots,n on [n][n], such that each pair MiM_{i} and Mi+1M_{i+1} forms an oriented matroid quotient. Then (M1,,Mn)(M_{1},\dots,M_{n}) is a complete flag positroid. Moreover, it is realized by a point of the positive Richardson u,v>0\mathcal{R}_{u,v}^{>0}, where we can explicitly construct uu and vv as follows:

  • Let B1min,,BnminB_{1}^{\min},\dots,B_{n}^{\min} (respectively, B1max,,BnmaxB_{1}^{\max},\dots,B_{n}^{\max}) be the bases of M1,,MnM_{1},\dots,M_{n} which are minimal (maximal) with respect to the usual Gale ordering. Then u,vSnu,v\in S_{n} are defined by

    u(i)=BiminBi1min and v(i)=BimaxBi1max.u(i)=B_{i}^{\min}\setminus B_{i-1}^{\min}\ \text{ and }\ v(i)=B_{i}^{\max}\setminus B_{i-1}^{\max}.
Proof.

As in 4.5, we identify each positroid MiM_{i} with the image t(χi)t(\chi_{i}) of its chirotope χi\chi_{i}; we have that t(χi)t(\chi_{i}) lies in Dri;n0\operatorname{Dr}_{i;n}^{\geq 0}. The fact that each pair Mi,Mi+1M_{i},M_{i+1} forms an oriented matroid quotient means that (t(χ1),,t(χn))(t(\chi_{1}),\dots,t(\chi_{n})) satisfies all three-term incidence-Plücker relations, and hence (t(χ1),,t(χn))FlDrn0(t(\chi_{1}),\dots,t(\chi_{n}))\in\operatorname{FlDr}_{n}^{\geq 0}. Since FlDrn0=TrFln0\operatorname{FlDr}_{n}^{\geq 0}=\operatorname{TrFl}_{n}^{\geq 0}, we have proved that (M1,,Mn)(M_{1},\dots,M_{n}) is a complete flag positroid.

To prove the characterization of uu and vv, we use 7.7. In particular, it follows from 7.7 and the Tableaux Criterion for Bruhat order that the Gale-minimal and Gale-maximal bases of the rank dd positroid π(u,v>0)\pi(\mathcal{R}_{u,v}^{>0}) are u([d])u([d]) and v([d])v([d]). The result now follows. ∎

As we’ve seen in 4.4 it is a subtle question to determine whether a pair of positroids M1M_{1} and M2M_{2} of ranks rr and r+1r+1 form an oriented matroid quotient. One way is to construct an nn by r+1r+1 matrix such that the minor in rows 1,,r1,\ldots,r and columns II is non-zero if and only if II is a basis of M1M_{1} while the maximal minor in rows 1,,r+11,\ldots,r+1 and columns JJ is non-zero if and only if JJ is a basis of M2M_{2}. 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 M1M_{1} and M2M_{2} form an oriented matroid quotient.

Construction # 1. Given two positroids M1M_{1} and M2M_{2} on the ground set [n][n] of ranks rr and r+1r+1, respectively, which form a positively oriented matroid quotient, we construct a positroid M:=M(M1,M2)M:=M(M_{1},M_{2}) of rank r+1r+1 on the ground set [n+1][n+1] where n+1n+1 is neither a loop nor a coloop. The bases of MM are precisely

(M)=(M2){B{n+1}B(M1)}.\mathcal{B}(M)=\mathcal{B}(M_{2})\cup\{B\cup\{n+1\}\mid B\in\mathcal{B}(M_{1})\}.

Construction #2. Conversely, given a rank rr positroid MM on ground set [n+1][n+1], where (n+1)(n+1) is neither a loop nor coloop, we construct two positroids M1:=M1(M)M_{1}:=M_{1}(M) and M2:=M2(M)M_{2}:=M_{2}(M) which form a positively oriented matroid quotient, as follows. Let A~\tilde{A} be a matrix realizing MM; therefore its Plücker coordinates are nonnegative. We apply row operations to rewrite A~\tilde{A} in the form

A=[0A001].A=\begin{bmatrix}\hskip 28.45274pt&\vline&0\\ \hskip 28.45274ptA^{\prime}\ \ \ \ \hskip 56.9055pt&\vline&0\\ \hskip 28.45274pt&\vline&0\\ \hline\cr*\;\;\;\;*\;\;\;\;*\;\;\;\;\;\;\cdots\;\;\;\;\;\;*&\vline&1\end{bmatrix}.

Let M1M_{1} denote the matroid on [n][n] realized by AA^{\prime} and let M2M_{2} denote the matroid on [n][n] realized by AA^{\prime} together with the row of *’s below it. Then M1M_{1} and M2M_{2} are both positroids (since the Plücker coordinates of AA^{\prime} and AA are all nonnegative), and they form a positively oriented quotient. Moreover, it is clear that M1=M(n+1)M_{1}=M\setminus(n+1) and M2=M/(n+1)M_{2}=M/(n+1).

The idea of our algorithm is to translate Constructions #1\#1 and #2\#2 into operations on Grassmann necklaces, so that Construction #1\#1 is well-defined even if M1M_{1} and M2M_{2} fail to form a positively oriented quotient. Clearly if we start with positroids M1M_{1} and M2M_{2} forming a positively oriented matroid quotient, then Construction #1\#1 followed by #2\#2 is the identity map. Conversely, if Construction #1\#1 followed by #2\#2 is the identity map, then since Construction #2\#2 always outputs a positively oriented matroid quotient, we must have started with positroids forming a positively oriented matroid quotient.

We let mini{S1,,Sk}\min_{i}\{S_{1},\cdots,S_{k}\} denote the minimum of the sets S1,,SkS_{1},\cdots,S_{k} in the i\leq_{i} order.

Proposition 7.11.

Let M1M_{1} and M2M_{2} be positroids of consecutive ranks which form a positively oriented quotient. Let Mj=(I1(j),,In(j))\mathcal{I}_{M_{j}}=(I_{1}^{(j)},\ldots,I_{n}^{(j)}) be the Grassmann necklace of MjM_{j} for j=1,2j=1,2. Define

Ji={I1(2), for i=1mini{Ii(1){n+1},Ii(2)}, for 2inI1(1){n+1}, for i=n+1..J_{i}=\begin{cases}I_{1}^{(2)},&\text{ for }i=1\\ \min_{i}\{I_{i}^{(1)}\cup\{n+1\},I_{i}^{(2)}\},&\text{ for }2\leq i\leq n\\ I_{1}^{(1)}\cup\{n+1\},&\text{ for }i=n+1.\end{cases}.

Then 𝒥=(J1,,Jn+1)\mathcal{J}=(J_{1},\dots,J_{n+1}) is the Grassmann necklace of the positroid M=M(M1,M2)M=M(M_{1},M_{2}) on [n+1][n+1] whose bases are precisely

(M)=(M2){B{n+1}B(M1)}.\mathcal{B}(M)=\mathcal{B}(M_{2})\cup\{B\cup\{n+1\}\mid B\in\mathcal{B}(M_{1})\}.
Proof.

It suffices to show that each basis of MM is ii-Gale greater than J(i)J^{(i)} for all i[n+1]i\in[n+1]. One also need to check that the J(i)J^{(i)} are in fact bases of MM but this is clear by definition.

Note that the i\leq_{i} minimal flag of a flag matroid consists of the i\leq_{i} minimal bases of each of its constituent matroids [BGW03, Corollary 7.2.1]. Thus, It(1)It(2)I_{t}^{(1)}\subset I_{t}^{(2)} for each t[n]t\in[n].

First, let S[n]S\subset[n] be a basis of M2M_{2}. For i[n]i\in[n], we have SiIi(2)iJiS\geq_{i}I_{i}^{(2)}\geq_{i}J_{i}. Since neither SS nor Ii(2)I_{i}^{(2)} contain n+1n+1, Sn+1I1(2)S\geq_{n+1}I_{1}^{(2)}. By our earlier observation, I1(2)=I1(1){a}I_{1}^{(2)}=I_{1}^{(1)}\cup\{a\} for some a[n]a\in[n]. Thus, I1(2)n+1I1(1){n+1}I_{1}^{(2)}\geq_{n+1}I_{1}^{(1)}\cup\{n+1\}. We conclude that SiJiS\geq_{i}J_{i} for all i[n+1]i\in[n+1].

Next, consider S{n+1}S\cup\{n+1\} for SS a basis of M1M_{1}. For 2in2\leq i\leq n, we have S{n+1}iIi(1){n+1}iJiS\cup\{n+1\}\geq_{i}I_{i}^{(1)}\cup\{n+1\}\geq_{i}J_{i}. Since neither SS nor Ii(1)I_{i}^{(1)} contain n+1n+1, we have S1I1(1)S\geq_{1}I_{1}^{(1)} and S{n+1}n+1I1(1){n+1}=Jn+1S\cup\{n+1\}\geq_{n+1}I_{1}^{(1)}\cup\{n+1\}=J_{n+1}. Since I1(2)=I1(1){a}I_{1}^{(2)}=I_{1}^{(1)}\cup\{a\}, we have I1(1){n+1}1I1(2)=J1I_{1}^{(1)}\cup\{n+1\}\geq_{1}I_{1}^{(2)}=J_{1}. We conclude that S{n+1}iJiS\cup\{n+1\}\geq_{i}J_{i} for all i[n+1]i\in[n+1]. ∎

If M1M_{1} and M2M_{2} form a positively oriented quotient, we should obtain them from the positroid M=M(M1,M2)M=M(M_{1},M_{2}), constructed as in 7.11, by deleting and contracting n+1n+1. The following result explains how these operations affect Grassmann necklaces.

Proposition 7.12.

[Oh08, Proposition 7 and Lemma 9] Let MM be a positroid on [n+1][n+1] such that n+1n+1 is neither a loop nor a coloop, with Grassmann necklace (Ji)i=1n+1(J_{i})_{i=1}^{n+1}. Then the Grassmann necklaces (K1(1),,Kn(1))\left(K_{1}^{(1)},\cdots,K_{n}^{(1)}\right) and (K1(2),,Kn(2))\left(K_{1}^{(2)},\cdots,K_{n}^{(2)}\right) of M1=M/(n+1)M_{1}=M/(n+1) and M2=M(n+1)M_{2}=M\setminus(n+1), are as follows:

Ki(1)=\displaystyle K_{i}^{(1)}= {Ji{n+1},n+1JiJi{maxi(JiJn+1)},n+1Ji\displaystyle\begin{cases}J_{i}\setminus\{n+1\},&n+1\in J_{i}\\ J_{i}\setminus\{\max_{i}(J_{i}\setminus J_{n+1})\},&n+1\notin J_{i}\end{cases}
Ki(2)=\displaystyle K_{i}^{(2)}= {(Ji{n+1}){mini(Jn+1Ji)},n+1JiJi,n+1Ji.\displaystyle\begin{cases}(J_{i}\setminus\{n+1\})\cup\{\min_{i}(J_{n+1}\setminus J_{i})\},&n+1\in J_{i}\\ J_{i},&n+1\notin J_{i}\end{cases}.

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 (1)=(I1(1),,In(1))\mathcal{I}^{(1)}=\left(I^{(1)}_{1},\ldots,I^{(1)}_{n}\right) and (2)=(I1(2),,In(2))\mathcal{I}^{(2)}=\left(I^{(2)}_{1},\ldots,I^{(2)}_{n}\right) be Grassmann necklaces of positroids of ranks rr and r+1r+1, respectively. Note that a necessary condition for the positroids corresponding to (1)\mathcal{I}^{(1)} and (2)\mathcal{I}^{(2)} forming a positively oriented quotient is that Ii(1)Ii(2)I^{(1)}_{i}\subset I^{(2)}_{i} for all i[n]i\in[n]. Now, we define a subset SS as follows: For each ii, if Ii(1){n+1}<iIi(2)I^{(1)}_{i}\cup\{n+1\}<_{i}I^{(2)}_{i}, let iSi\in S. Since Ii(2)=Ii(1)aI^{(2)}_{i}=I^{(1)}_{i}\cup a for some a[n]a\in[n], this is as simple as checking whether a<in+1a<_{i}n+1. If the positroids corresponding to (1)\mathcal{I}^{(1)} and (2)\mathcal{I}^{(2)} form a positively oriented quotient, applying 7.11 and then 7.12 should leave them unchanged. It is straightforward to see that iSi\in S if and only if n+1Jin+1\in J_{i} in 7.12. In particular, since 𝒥\mathcal{J} is a Grassmann necklace, SS must either be an interval of the form [d,n][d,n], or empty.

Next we claim that, once we verify that SS is an interval of the form [d,n][d,n] or is empty, then it follows automatically that 𝒥\mathcal{J}, as constructed in 7.11, is a Grassmann necklace.

Lemma 7.13.

Let (1)=(I1(1),,In(1))\mathcal{I}^{(1)}=\left(I^{(1)}_{1},\ldots,I^{(1)}_{n}\right) and (2)=(I1(2),,In(2))\mathcal{I}^{(2)}=\left(I^{(2)}_{1},\ldots,I^{(2)}_{n}\right) be Grassmann necklaces of types (r,n)(r,n) and (r+1,n)(r+1,n), respectively. Construct 𝒥=(J1,,Jn+1)\mathcal{J}=(J_{1},\ldots,J_{n+1}) as in 7.11. Let S={i[n]|Ii(1)(n+1)<iIi(2)}S=\left\{i\in[n]|I^{(1)}_{i}\cup(n+1)<_{i}I^{(2)}_{i}\right\}. If S=[d,n]S=[d,n] for some dnd\leq n or S=S=\emptyset, then 𝒥\mathcal{J} is a Grassmann necklace.

Proof.

It is clear from the definition that 𝒥\mathcal{J} satisfies the Grassmann necklace condition for each pair of consecutive sets JiJ_{i} and Ji+1J_{i+1} except for when i=k1i=k-1, i=ni=n and i=n+1i=n+1 (where we label sets cyclically so that Jn+2=J1J_{n+2}=J_{1}).

If SS\neq\emptyset, then Jn=In(1){n+1}J_{n}=I^{(1)}_{n}\cup\{n+1\}. This makes it clear that the Grassmann necklace condition holds for JnJ_{n} and Jn+1J_{n+1}. Also, sing the fact that Ii(1)Ii(2)I^{(1)}_{i}\subset I^{(2)}_{i} for all ii, it is not hard to verify the Grassmann necklace condition for Jn+1J_{n+1} and J1J_{1}.

This leaves us to check the condition for Jk1J_{k-1} and JkJ_{k}. In this case, Jk1=Ik1(2)J_{k-1}=I^{(2)}_{k-1} and Jk=Ik(1){n+1}J_{k}=I^{(1)}_{k}\cup\{n+1\}. Our goal is to show that Jk=(Jk1{k1}){a}J_{k}=(J_{k-1}\setminus\{k-1\})\cup\{a\} for some a[n+1]a\in[n+1]. It is immediately obvious that we necessarily have a=n+1a=n+1. Thus, we are left to show that Ik(1){n+1}=(Ik1(2){k1}){n+1}I^{(1)}_{k}\cup\{n+1\}=(I^{(2)}_{k-1}\setminus\{k-1\})\cup\{n+1\}, or that Ik(1)=Ik1(2){k1}I^{(1)}_{k}=I^{(2)}_{k-1}\setminus\{k-1\}.

Let aia_{i} be defined by Ii(1)=(Ii1(1){i1}){ai}I^{(1)}_{i}=(I^{(1)}_{i-1}\setminus\{i-1\})\cup\{a_{i}\}, let bib_{i} be defined by Ii(2)=(Ii1(2){i1}){bi}I^{(2)}_{i}=(I^{(2)}_{i-1}\setminus\{i-1\})\cup\{b_{i}\} and let cic_{i} be defined by Ii(2)=Ii(1){ci}I^{(2)}_{i}=I^{(1)}_{i}\cup\{c_{i}\}. We observe that Ik(1)=(Ik1(1){k1}){ak}=(Ik1(2){ck1,k1}){ak}I^{(1)}_{k}=(I^{(1)}_{k-1}\setminus\{k-1\})\cup\{a_{k}\}=(I^{(2)}_{k-1}\setminus\{c_{k-1},k-1\})\cup\{a_{k}\}. Also, Ik(1)=Ik(2){ck}=(Ik1(2){ck,k1}){bk}I^{(1)}_{k}=I^{(2)}_{k}\setminus\{c_{k}\}=(I^{(2)}_{k-1}\setminus\{c_{k},k-1\})\cup\{b_{k}\}. Comparing these two equalities, we conclude that either ak=ck1a_{k}=c_{k-1} and bk=ckb_{k}=c_{k}, or ck1=ckc_{k-1}=c_{k} and ak=bka_{k}=b_{k}. The first case is what we want to prove, so let us show by contradiction that the second case cannot occur.

Assume ck=ck1c_{k}=c_{k-1} and ak=bka_{k}=b_{k}. By assumption, Ik1(1){ck1}=Ik1(2)<k1Ik1(1){n+1}I^{(1)}_{k-1}\cup\{c_{k-1}\}=I^{(2)}_{k-1}<_{k-1}I^{(1)}_{k-1}\cup\{n+1\} and Ik(1){n+1}<kIk(2)=Ik(1){ck}I^{(1)}_{k}\cup\{n+1\}<_{k}I^{(2)}_{k}=I^{(1)}_{k}\cup\{c_{k}\}. Thus, ck1<k1n+1c_{k-1}<_{k-1}n+1 and ck>kn+1c_{k}>_{k}n+1. Since ck=ck1c_{k}=c_{k-1}, this means they are both equal to k1k-1. However, if ck=k1c_{k}=k-1, then M2M_{2} has k1k-1 as a coloop. it follows that bk=k1b_{k}=k-1, which means ak=k1a_{k}=k-1 as well. Thus, in this case, Ik(1)=Ik1(1)=Ik1(2){k1}I_{k}^{(1)}=I_{k-1}^{(1)}=I_{k-1}^{(2)}\setminus\{k-1\}, as desired.

Finally, if A=A=\emptyset, we can check that the Grassmann necklace condition holds for Jn+1J_{n+1} and J1J_{1} as before. The we are just left to verify this condition for JnJ_{n} and Jn+1J_{n+1}. We can apply the same logic but with Jk1J_{k-1} replaced by Jn=In(2)J_{n}=I_{n}^{(2)} and JkJ_{k} replaced by Jn+1=I(1){n+1}J_{n+1}=I^{(1)}\cup\{n+1\}. Specifically, we find I1(1)=(In(2){cn,n}){a1}=In(2){c1,n}){b1}I_{1}^{(1)}=(I_{n}^{(2)}\setminus\{c_{n},n\})\cup\{a_{1}\}=I_{n}^{(2)}\setminus\{c_{1},n\})\cup\{b_{1}\}. We then must show that it is impossible for c1=cnc_{1}=c_{n} and a1=b1a_{1}=b_{1}. However, In(1){cn}=In(2)<nIn(1){n+1}I_{n}^{(1)}\cup\{c_{n}\}=I_{n}^{(2)}<_{n}I_{n}^{(1)}\cup\{n+1\}. Moreover, it is always true that I1(1){n+1}<n+1I1(1){c1}=I2(1)I_{1}^{(1)}\cup\{n+1\}<_{n+1}I_{1}^{(1)}\cup\{c_{1}\}=I_{2}^{(1)}. Using c1=cnc_{1}=c_{n}, we then find cn<n(n+1)c_{n}<_{n}(n+1) and cn>n+1(n+1)c_{n}>_{n+1}(n+1) which means that c1=cn=nc_{1}=c_{n}=n and we can conclude as in the previous paragraph. ∎

Combining 7.11, 7.12 and 7.13, we obtain the following:

Theorem 7.14.

Fix positroids M1M_{1} and M2M_{2} on [n][n] of ranks rr and r+1r+1, respectively. Let =M1=(I1,,In)\mathcal{I}=\mathcal{I}_{M_{1}}=\left(I_{1},\ldots,I_{n}\right) and 𝒥=M2=(J1,,Jn)\mathcal{J}=\mathcal{I}_{M_{2}}=\left(J_{1},\ldots,J_{n}\right) be their Grassmann necklaces. We now set S={i[n]Ii{n+1}iJi}S=\left\{i\in[n]\mid I_{i}\cup\{n+1\}\leq_{i}J_{i}\right\}, where i\leq_{i} denotes the i\leq_{i} Gale order on [n+1][n+1]. Define ai=maxi(JiI1)a_{i}=\max_{i}\left(J_{i}\setminus I_{1}\right) and bi=mini(I1Ii)b_{i}=\min_{i}\left(I_{1}\setminus I_{i}\right). Then M1M_{1} and M2M_{2} form a positively oriented quotient if and only if the following conditions hold:

  1. (1)

    For i[n]i\in[n], IiJiI_{i}\subset J_{i}.

  2. (2)

    SS is an interval of the form [d,n][d,n] or S=S=\emptyset.

  3. (3)

    For iSi\notin S, Ii=Ji{ai}I_{i}=J_{i}\setminus\{a_{i}\}.

  4. (4)

    For iSi\in S, Ji=Ii{bi}J_{i}=I_{i}\cup\{b_{i}\}.

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 33 and 44.

Conversely, if the conditions in the theorem statement hold, then by 7.13, applying the construction of 7.11 to \mathcal{I} and 𝒥\mathcal{J} yields another Grassmann necklace 𝒦\mathcal{K} on [n+1][n+1] such that n+1n+1 is neither a loop nor a coloop of the positroid corresponding to 𝒦\mathcal{K}. Then, conditions 33 and 44 guarantee that applying the construction of 7.12 to 𝒦\mathcal{K} will recover \mathcal{I} and 𝒥\mathcal{J}. The result of applying 7.12 to the Grassmann necklace of a positroid MM with n+1n+1 neither a loop nor a coloop is the pair of Grassmann necklaces corresponding to M/(n+1)M/(n+1) and M(n+1)M\setminus(n+1), which form a positively oriented quotient. ∎

Example 7.15.

Let =(123,235,356,456,561,613)\mathcal{I}=(123,235,356,456,561,613) and 𝒥=(1235,2356,3456,4562,5612,6123)\mathcal{J}=(1235,2356,3456,4562,5612,6123). Then A={4,5,6}A=\{4,5,6\} is an interval with upper endpoint n=6n=6. Note that a1=5a_{1}=5, a2=6a_{2}=6 and a3=6a_{3}=6, while b4=1b_{4}=1, b5=2b_{5}=2 and b6=2b_{6}=2. The positroids with these Grassmann necklaces do not form a positively oriented quotient since it is false that I3=J3{a3}I_{3}=J_{3}\setminus\{a_{3}\}.

However, if we start with Grassmann necklaces =(123,235,345,456,561,613)\mathcal{I}=(123,235,345,456,561,613) and 𝒥=(1235,2356,3456,4562,5612,6123)\mathcal{J}=(1235,2356,3456,4562,5612,6123), then the values of the aia_{i} and bib_{i} are unchanged. It is straightforward to verify that the conditions of 7.14 hold and so the positroids corresponding to \mathcal{I} and 𝒥\mathcal{J} 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 (Ma,Ma+1,,Mb)(M_{a},M_{a+1},\ldots,M_{b}) is a sequence of positroids of ranks a,a+1,,ba,a+1,\dots,b. Then (Ma,Ma+1,,Mb)(M_{a},M_{a+1},\ldots,M_{b}) is a flag positroid if and only if for ai<ba\leq i<b, the pair of positroids (Mi,Mi+1)(M_{i},M_{i+1}) satisfy the conditions of 7.14.

Proof.

By 7.10, it suffices to check that each such pair forms a positively oriented quotient, which is precisely the content of 7.14. ∎

8. Fan structures for and coherent subdivisions from TrGrd;n>0\operatorname{TrGr}_{d;n}^{>0} and TrFln>0\operatorname{TrFl}_{n}^{>0}

In this section we make some brief remarks about the various fan structures for TrFl𝐫;n>0\operatorname{TrFl}_{\mathbf{r};n}^{>0} and coherent subdivisions from points of TrFl𝐫;n>0\operatorname{TrFl}_{\mathbf{r};n}^{>0}. 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 TrFl4>0\operatorname{TrFl}_{4}^{>0}.

8.1. Fan structures

There are multiple possibly different natural fan structures for TrFl𝐫;n>0\operatorname{TrFl}_{\mathbf{r};n}^{>0}:

  1. (i)

    The Plücker fan (induced by the three-term tropical Plücker relations).

  2. (ii)

    The secondary fan (induced according to the coherent subdivision as in 8.3).

  3. (iii)

    The Gröbner fan (induced according to the initial ideal of the ideal 𝒫𝐫;n\langle\mathscr{P}_{\mathbf{r};n}\rangle).

  4. (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 Fl𝐫;n>0\operatorname{Fl}_{\mathbf{r};n}^{>0}, such as an 𝒳\mathcal{X}-cluster chart.

  5. (v)

    (If the cluster algebra associated to Fl𝐫;n\operatorname{Fl}_{\mathbf{r};n} 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 Gr2,n\operatorname{Gr}_{2,n}, fan (iv) (which coincides with (v)) is isomorphic to the cluster complex444See [FZ03] for background on the cluster complex. of type An3A_{n-3}; for Gr3,6\operatorname{Gr}_{3,6} and Gr3,7\operatorname{Gr}_{3,7}, fan (iv) is isomorphic to a coarsening of the corresponding cluster complex, while fan (v) is isomorphic to the cluster complex (of types D4D_{4} and E6E_{6}, 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 FF-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 TrGr2,n\operatorname{TrGr}_{2,n}, 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 TrGr2,n>0\operatorname{TrGr}_{2,n}^{>0}. For TrGr3,6>0\operatorname{TrGr}_{3,6}^{>0}, 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 n=3n=3, the fan TrFln>0\operatorname{TrFl}_{n}^{>0} modulo its lineality space is a one-dimensional fan, and all fan structures coincide. For TrFln\operatorname{TrFl}_{n} (before taking the positive part), one can find computations of the fan (iii) for n=4n=4 and n=5n=5 in [BLMM17, §3], the fan (i) and its relation to (iii) for n=4n=4 in [BEZ21, Example 5.2.3], and the fan (ii) and its relation to (iii) for n=4n=4 in [JLLO, §5]. Returning to the positive tropicalization, [Bos22, Section 5.1] computed the fan structure (iii) for TrFl4>0\operatorname{TrFl}_{4}^{>0}, and found it was dual to the three-dimensional associahedron; in particular, there are 1414 maximal cones and the ff-vector is (14,21,9,1)(14,21,9,1). Using the positive parameterization of [Bor22] (a graphical version of the parameterizations of [MR04]) for TrFln>0\operatorname{TrFl}_{n}^{>0}, we computed the polyhedral complex underlying (iv) for n=4n=4 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 ff-vector (13,20,9,1)(13,20,9,1). We also computed (v) after incorporating the additional non-Plücker cluster variable p2p134p1p234p_{2}p_{134}-p_{1}p_{234}. Combining these, we find that for n=4n=4, (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 TrFl4>0\operatorname{TrFl}_{4}^{>0} is dual to the three-dimensional associahedron is consistent with [SW05, Conjecture 8.1] and the fact that Fl4\operatorname{Fl}_{4} has a cluster algebra structure of finite type A3A_{3} [GLS08, Table 1], whose cluster complex is dual to the associahedron.

We now give a graphical way to think about the fan structure on TrFl4>0\operatorname{TrFl}_{4}^{>0}, building on the ideas of [SW05] and [BEZ21, Example 5.2.3].

112233445555^{\prime}112233445555^{\prime}112233445555^{\prime}112233445555^{\prime}112233445555^{\prime}112233445555^{\prime}112233445555^{\prime}112233445555^{\prime}112233445555^{\prime}112233445555^{\prime}112233445555^{\prime}112233445555^{\prime}112233445555^{\prime}112233445555^{\prime}112233445555^{\prime}112233445555^{\prime}112233445555^{\prime}112233445555^{\prime}
Figure 2. The fan structure (ii)=(iii) of TrFl4>0\operatorname{TrFl}_{4}^{>0}.
Example 8.1.

A planar tree on [n][n] is an unrooted tree drawn in the plane with nn leaves labeled by 1,2,,n1,2,\dots,n (in counterclockwise order). By [SW05], TrGr2;n>0\operatorname{TrGr}_{2;n}^{>0} 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 ii and jj encodes the positive tropical Plücker coordinate wijw_{ij} of a point in the corresponding cone. In particular, it is easy to see that the negative distances wijw_{ij} 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 μ\mu whose underlying matroid is the uniform matroid U2,4U_{2,4}, the tropical linear spaces trop(μ)\operatorname{trop}(\mu) and trop(μ)\operatorname{trop}(\mu^{*}) associated to μ\mu and its dual μ\mu^{*} are translates of each other. This allows us to identify points 𝝁=(μ1,μ2,μ3){\boldsymbol{\mu}}=(\mu_{1},\mu_{2},\mu_{3}) of TrFl4>0\operatorname{TrFl}_{4}^{>0} with planar trees on the vertices {1,2,3,4,5,5}\{1,2,3,4,5,5^{\prime}\} such that the vertices {1,2,3,4,5}\{1,2,3,4,5\} and separately the vertices {1,2,3,4,5}\{1,2,3,4,5^{\prime}\} appear in counterclockwise order. To see this, note that (using the same idea as Construction #1 from Section 7) we can identify (μ1,μ2)(\mu_{1},\mu_{2}), with Plücker coordinates (w1,,w4;w12,,w34)(w_{1},\dots,w_{4};w_{12},\dots,w_{34}), with an element (wab)(w_{ab}) of TrGr2,5>0\operatorname{TrGr}_{2,5}^{>0}: we simply set wa5:=waw_{a5}:=w_{a} for 1a41\leq a\leq 4. Similarly, we identify (μ2,μ3)(\mu_{2},\mu_{3}), where μ3\mu_{3} has Plücker coordinates (w123,,w234)(w_{123},\dots,w_{234}), with an element of TrGr2,5>0\operatorname{TrGr}_{2,5}^{>0}: we simply set wd5:=wabcw_{d5^{\prime}}:=w_{abc}, where {a,b,c}:=[4]{d}\{a,b,c\}:=[4]\setminus\{d\}.

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 dist(x1,blue)=dist(x2,red)\operatorname{dist}(x_{1},blue)=\operatorname{dist}(x_{2},red), where x1x_{1} and x2x_{2} are the two black trivalent nodes in the tree on [4][4]. To obtain the fan structure (v), instead of the squiggly line, the square face is subdivided along the other diagonal.

Using the computation of TrFl5\operatorname{TrFl}_{5} in [BLMM17], available at https://github.com/Saralamboglia/Toric-Degenerations/blob/master/Flag5.rtf and 3.12, we further computed that TrFl5+\operatorname{TrFl}_{5}^{+} 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 TrFl5+\operatorname{TrFl}_{5}^{+} 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 Fl𝐫;n\operatorname{Fl}_{\mathbf{r};n} is the Grassmannian Grd,n\operatorname{Gr}_{d,n} and the support 𝝁¯\underline{{\boldsymbol{\mu}}} is the uniform matroid, A gives rise to the following corollary (which was first proved in [LPW20] and [AHLS20]).

Corollary 8.2.

Let 𝝁=(μd)(𝕋([n]d)){\boldsymbol{\mu}}=(\mu_{d})\in\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{d}}\right), and suppose it has no \infty coordinates. Then the following statements are equivalent.

  • 𝝁TrGrd,n>0{\boldsymbol{\mu}}\in\operatorname{TrGr}_{d,n}^{>0}, that is, 𝝁{\boldsymbol{\mu}} lies in the strictly positive tropical Grassmannian.

  • Every face in the coherent subdivision 𝒟𝝁\mathcal{D}_{\boldsymbol{\mu}} of the hypersimplex Δd,n\Delta_{d,n} induced by 𝝁{\boldsymbol{\mu}} 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 Δd,n\Delta_{d,n} achieve equality in Speyer’s ff-vector theorem; in particular, they all consist of (n2d1){n-2\choose d-1} facets [SW21, Corollary 6.7].

When Fl𝐫;n\operatorname{Fl}_{\mathbf{r};n} is the complete flag variety Fln\operatorname{Fl}_{n}, and the support 𝝁¯\underline{{\boldsymbol{\mu}}} is the uniform flag matroid, A gives rise to the following corollary, which appeared in [JLLO, Theorem 20].

Corollary 8.3.

Let 𝝁=(μ1,,μn)i=ab(𝕋([n]i)){\boldsymbol{\mu}}=(\mu_{1},\ldots,\mu_{n})\in\prod_{i=a}^{b}\mathbb{P}\left(\mathbb{T}^{\binom{[n]}{i}}\right), and suppose it has no \infty coordinates. Then the following statements are equivalent.

  • 𝝁TrFln>0{\boldsymbol{\mu}}\in\operatorname{TrFl}_{n}^{>0}, that is, 𝝁{\boldsymbol{\mu}} lies in the strictly positive tropical flag variety.

  • Every face in the coherent subdivision 𝒟𝝁\mathcal{D}_{\boldsymbol{\mu}} of the permutohedron Permn\operatorname{Perm}_{n} induced by 𝝁{\boldsymbol{\mu}} 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 Permn\operatorname{Perm}_{n} into Bruhat interval polytopes. Furthermore, do they all have the same ff-vector?

Explicit computations for TrFl4\operatorname{TrFl}_{4} show that the answer to the second question is no. We find that TrFl4\operatorname{TrFl}_{4} with the fan structure (iii) (which agrees with (ii) by [JLLO, §5]) has 7878 maximal cones. We choose a point in the relative interior of each of the 7878 cones to use as a height function (thinking of points in TrFl4\operatorname{TrFl}_{4} as weights on the vertices of Perm4\operatorname{Perm}_{4} as in Item (c) of A), then use Sage to compute the corresponding coherent subdivision of Perm4\operatorname{Perm}_{4}. As expected, precisely 1414 of the 7878 cones induce subdivisions of Perm4\operatorname{Perm}_{4} into Bruhat interval polytopes, see Table 1.

Height function (P1,P2,P3,P4(P_{1},P_{2},P_{3},P_{4}; P12,P13,P_{12},P_{13},
P14,P23,P24,P34P_{14},P_{23},P_{24},P_{34}; P123,P124,P134,P234)P_{123},P_{124},P_{134},P_{234})
Bruhat interval polytopes
in subdivision
ff-vector
(15,1,7,7;4,2,2,(15,-1,-7,-7;4,-2,-2,
2,2,4;7,7,1,15)-2,-2,4;-7,-7,-1,15)
P3214,4321,P3124,4231,P2314,3421,P_{3214,4321},P_{3124,4231},P_{2314,3421},
P2134,3241,P1324,2431,P1234,2341P_{2134,3241},P_{1324,2431},P_{1234,2341}
(24,46,29,6)(24,46,29,6)
(15,3,9,9;4,8,8,(15,3,-9,-9;4,-8,-8,
4,4,20;1,1,1,3)-4,-4,20;-1,-1,-1,3)
P2413,4321,P3124,4231,P2314,4231,P_{2413,4321},P_{3124,4231},P_{2314,4231},
P2134,3241,P1324,2431,P1234,2341P_{2134,3241},P_{1324,2431},P_{1234,2341}
(15,7,1,7;2,4,2,(15,-7,-1,-7;-2,4,-2,
2,4,2;7,1,7,15)-2,4,-2;-7,-1,-7,15)
P3142,4321,P3124,4312,P2143,3421,P_{3142,4321},P_{3124,4312},P_{2143,3421},
P2134,3412,P1243,2431,P1234,2413P_{2134,3412},P_{1243,2431},P_{1234,2413}
(1,1,1,3;4,8,4,(-1,-1,-1,3;4,-8,-4,
8,4,20;15,3,9,9)-8,-4,20;15,3,-9,-9)
P2413,4321,P1423,4231,P1342,4231,P_{2413,4321},P_{1423,4231},P_{1342,4231},
P1324,4213,P1243,4132,P1234,4123P_{1324,4213},P_{1243,4132},P_{1234,4123}
(7,7,1,15;4,2,2,(-7,-7,-1,15;4,-2,-2,
2,2,4;15,1,,7,7)-2,-2,4;15,-1,,-7,-7)
P1432,4321,P1423,4312,P1342,4231,P_{1432,4321},P_{1423,4312},P_{1342,4231},
P1324,4213,P1243,4132,P1234,4123P_{1324,4213},P_{1243,4132},P_{1234,4123}
(1,7,7,15;2,2,4,(-1,-7,-7,15;-2,-2,4,
4,2,2;15,7,7,1)4,-2,-2;15,-7,-7,-1)
P3142,4321,P2143,4312,P2134,4213,P_{3142,4321},P_{2143,4312},P_{2134,4213},
P1342,3421,P1243,3412,P1234,2413P_{1342,3421},P_{1243,3412},P_{1234,2413}
(9,9,3,15;20,4,8,(-9,-9,3,15;20,-4,-8,
4,8,4;3,1,1,1)-4,-8,4;3,-1,-1,-1)
P1432,4321,P1423,4312,P1342,4231,P_{1432,4321},P_{1423,4312},P_{1342,4231},
P1324,4213,P1324,4132,P1234,3142P_{1324,4213},P_{1324,4132},P_{1234,3142}
(11,7,7,3;6,6,4,(11,-7,-7,3;-6,-6,4,
4,2,2;11,7,7,3)4,2,2;11,-7,-7,3)
P3142,4321,P2143,4312,P2134,4213,P_{3142,4321},P_{2143,4312},P_{2134,4213},
P2143,3421,P1243,2431,P1234,2413P_{2143,3421},P_{1243,2431},P_{1234,2413}
(3,3,3,3;20,10,10,(3,3,-3,-3;20,-10,-10,
10,10,20;3,3,3,3)-10,-10,20;-3,-3,3,3)
P2413,4321,P3124,4231,P2314,4231,P_{2413,4321},P_{3124,4231},P_{2314,4231},
P1324,2431,P1324,3241,P1234,3142P_{1324,2431},P_{1324,3241},P_{1234,3142}
(3,1,1,1;20,4,4,(3,-1,-1,-1;20,-4,-4,
8,8,4;9,9,3,15)-8,-8,4;-9,-9,3,15)
P3214,4321,P3124,4231,P2314,3421,P_{3214,4321},P_{3124,4231},P_{2314,3421},
P1324,3241,P1324,2431,P1234,3142P_{1324,3241},P_{1324,2431},P_{1234,3142}
(3,3,3,3;20,10,10,(-3,-3,3,3;20,-10,-10,
10,10,20;3,3,3,3)-10,-10,20;3,3,-3,-3)
P2413,4321,P1423,4231,P1342,4231,P_{2413,4321},P_{1423,4231},P_{1342,4231},
P1324,4132,P1324,4213,P1234,3142P_{1324,4132},P_{1324,4213},P_{1234,3142}
(3,7,7,11;2,2,4,(3,-7,-7,11;2,2,4,
4,6,6;3,7,7,11)4,-6,-6;3,-7,-7,11)
P3142,4321,P3124,4312,P1342,3421,P_{3142,4321},P_{3124,4312},P_{1342,3421},
P2134,3412,P1243,3412,P1234,2413P_{2134,3412},P_{1243,3412},P_{1234,2413}
(11,1,7,3;2,8,4,(11,-1,-7,-3;-2,-8,-4,
4,0,18;11,1,7,3)-4,0,18;11,-1,-7,-3)
P2413,4321,P2143,4231,P2134,4213,P_{2413,4321},P_{2143,4231},P_{2134,4213},
P1243,2431,P1234,2413P_{1243,2431},P_{1234,2413}
(24,45,27,5)(24,45,27,5)
(3,7,1,11;18,0,4,(-3,-7,-1,11;18,0,-4,
4,8,2;3,7,1,11)-4,-8,-2;-3,-7,-1,11)
P3142,4321,P3124,4312,P1342,3421P_{3142,4321},P_{3124,4312},P_{1342,3421}
P1324,3412,P1234,3142P_{1324,3412},P_{1234,3142}
Table 1. Table documenting the 1414 finest coherent subdivisions of Perm4\operatorname{Perm}_{4} into Bruhat interval polytopes. There are two possible ff-vectors, each of which can be realized in multiple ways.

Of the 1414 coherent subdivisions coming from maximal cones of TrFl4>0\operatorname{TrFl}_{4}^{>0}, 1212 of them contain 66 facets, while the other 22 contain 55 facets. Table 1 lists the facets and ff-vectors of each of these 1414 subdivisions. Note that each Bruhat interval polytope Pv,wP_{v,w} which appears as a facet satisfies (w)(v)=3\ell(w)-\ell(v)=3. Thus, any Bruhat interval polytope Pv,wP_{v^{\prime},w} properly contained inside Pv,wP_{v,w} would have the property that (w)(v)2\ell(w^{\prime})-\ell(v^{\prime})\leq 2, and hence dim(Pv,w)2\dim(P_{v^{\prime},w^{\prime}})\leq 2. Since Perm4\operatorname{Perm}_{4} is 33-dimensional, all 1414 of these subdivisions are finest subdivisions.

We note that the 1212 finest subdivisions whose ff-vector is (24,46,29,6)(24,46,29,6) 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 Permn\operatorname{Perm}_{n} into (n1)!(n-1)! Bruhat interval polytopes

{Pu,v|u=(u1,un) with un=n, and v=(v1,,vn) with vi=ui+1 modulo n}.\{P_{u,v}\ |\ u=(u_{1}\dots,u_{n})\text{ with }u_{n}=n,\text{ and }v=(v_{1},\dots,v_{n})\text{ with }v_{i}=u_{i}+1\text{ modulo }n\}.

The first subdivision in Table 1 has this form.

We can further study the ff-vectors of subdivisions of TrFl4>0\operatorname{TrFl}_{4}^{>0} which are coarsest (without being trivial), rather than finest. In this case, we observe three different ff-vectors, each of which occurs in multiple subdivisions. The detailed results of our explicit computations on coarsest subdivisions can be found in Table 2.

Height function (P1,P2,P3,P4;P12,P13,(P_{1},P_{2},P_{3},P_{4};P_{12},P_{13},
P14,P23,P24,P34;P123,P124,P134,P234)P_{14},P_{23},P_{24},P_{34};P_{123},P_{124},P_{134},P_{234})
Bruhat interval polytopes
in subdivision
ff-vector
(1,1,1,0;1,1,0,1,0,0;0,0,0,0)(-1,-1,-1,0;-1,-1,0,-1,0,0;0,0,0,0)
P1243,4321,P1234,4213P_{1243,4321},P_{1234,4213}
(24,39,18,2)(24,39,18,2)
(1,1,1,0;0,0,0,0,0,0;0,0,0,0)(-1,-1,-1,0;0,0,0,0,0,0;0,0,0,0)
P1342,4321,P1234,4312P_{1342,4321},P_{1234,4312}
(1,0,0,0;0,0,0,0,0,0;0,0,0,0)(1,0,0,0;0,0,0,0,0,0;0,0,0,0)
P2134,4321,P1234,2431P_{2134,4321},P_{1234,2431}
(1,0,0,0;0,0,0,1,1,1;0,0,0,0)(1,0,0,0;0,0,0,1,1,1;0,0,0,0)
P3124,4321,P1234,3421P_{3124,4321},P_{1234,3421}
(0,0,0,0;1,1,1,1,1,0;0,0,0,0)(0,0,0,0;-1,-1,-1,-1,-1,0;0,0,0,0)
P2413,4321,P1234,4231P_{2413,4321},P_{1234,4231}
(24,40,19,2)(24,40,19,2)
(0,0,0,0;1,0,0,0,0,0;0,0,0,0)(0,0,0,0;1,0,0,0,0,0;0,0,0,0)
P1324,4321,P1234,3142P_{1324,4321},P_{1234,3142}
(1,1,0,0;1,1,1,1,1,0;0,0,0,0)(-1,-1,0,0;-1,-1,-1,-1,-1,0;0,0,0,0)
P1423,4321,P1342,4231,P_{1423,4321},P_{1342,4231},
P1324,4213,P1234,4132P_{1324,4213},P_{1234,4132}
(24,42,23,4)(24,42,23,4)
(0,1,1,0;0,0,1,0,0,0;0,0,0,0)(0,-1,-1,0;0,0,1,0,0,0;0,0,0,0)
P3142,4321,P1243,3421,P_{3142,4321},P_{1243,3421},
P2134,4312,P1234,2413P_{2134,4312},P_{1234,2413}
(1,1,0,0;1,0,0,0,0,0;0,0,0,0)(1,1,0,0;1,0,0,0,0,0;0,0,0,0)
P2314,4321,P1324,2431,P_{2314,4321},P_{1324,2431},
P3124,4231,P1234,3241P_{3124,4231},P_{1234,3241}
Table 2. Table documenting the 99 coarsest coherent subdivisions of Perm4\operatorname{Perm}_{4} into Bruhat interval polytopes. There are three possible ff-vectors, each of which can be realized in multiple ways.

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 MM^{\natural}-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 G/PG/P. Math. Res. Lett., 13(5-6):775–786, 2006.
  • [RW08] Konstanze Rietsch and Lauren Williams. The totally nonnegative part of G/PG/P 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.