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

Amalgams of matroids, fibre products
and tropical graph correspondences

Dmitry Mineev Bar-Ilan University, Ramat Gan, Israel [email protected]
Abstract.

We prove that the proper amalgam of matroids M1M_{1} and M2M_{2} along their common restriction NN exists if and only if the tropical fibre product of Bergman fans B(M1)×B(N)B(M2){B(M_{1})\times_{B(N)}B(M_{2})} is positive. We introduce tropical correspondences between Bergman fans as tropical subcycles in their product, similar to correspondences in algebraic geometry, and define a “graph correspondence” of the map of lattices. We prove that graph construction is a functor for the “covering” maps of lattices, exploiting a generalization of Bergman fan which we call a “Flag fan”.

1. Introduction

One of the features of matroid theory is that it makes use of numerous inductive constructions such as deletions, contractions and extensions (see, for example, [Oxl11], Chapter 7). In this light, various attempts to “glue” matroids together were studied for a long time. The most basic construction having its roots in graphic matroids is the parallel connection, which sometimes plays even bigger role than the direct sum ([SW23]). Generalizations of parallel connection, called amalgams, were studied in the 1970s and 1980s ([BK88], [PT82]). Among those, the free amalgam stands out as the one satisfying a certain universal property in the category of matroids with morphisms being what is called weak maps. A slightly stronger notion of the free amalgam, the proper amalgam, is naturally called “generalized parallel connection” in some cases.

It turns out, though, that in general amalgamation is complicated — for a given pair of matroids that we try to glue, there may be no free amalgam, or the free amalgam may not be proper, or no amalgams may exist at all. Several criteria guaranteeing the existence of the proper amalgam were established ([Oxl11], Section 11.4), but even the weakest of them is not a necessary condition. Amalgams continue to be investigated — for example, the so-called sticky matroid conjecture was apparently resolved recently ([Bon09], [Shi21]).

Over the last thirty years, tropical geometry, initiated as a combinatorial tool of algebraic geometry inspired by the formalism of toric geometry, has established many connections to matroid theory. From the tropical point of view, a Bergman fan — the tropical variety corresponding to a loopless matroid — can be considered an analog of the smooth tangent cone. Bergman fans are widely used in many of the latest developments ([AHK15], [MRS19], [BEST]).

In particular, in the beginning of the 2010s, a well-behaved intersection theory of the tropical cycles in Bergman fans was developed simultaneously by Allermann, Esterov, Francois, Rau, and Shaw ([Est11], [Sha13], [AR09], [FR12]). This intersection theory possesses many similarities with the intersection theory in algebraic geometry, going as far as the projection formula. This machinery was sufficient to define a tropical subcycle in the product of Bergman fans called the tropical fibre product ([FH13], [Cav+16]), which the authors used to study moduli spaces of curves. They considered the case when the support of the tropical fibre product equals the set-theoretic fibre product of the supports of the factors, finding several sufficient conditions on matroids for this case to occur.

As it turns out, though, the tropical fibre product is worth investigating in wider generality. This becomes evident when considering a simple Example 2.22 (2).

Theorem A (4.1).

The tropical fibre product equals the Bergman fan of the proper amalgam if the proper amalgam exists, and has cones with negative weights otherwise.

Thus, unlike all criteria for the existence of the proper amalgam known to us, Theorem 4.1 is an equivalence. It is also constructive and, moreover, algorithmic in a certain sense, as shown in Section 3.

With tropical counterpart of the proper amalgam in our hands, we seek a categorical justification of its name — the fibre product. As it is justly pointed out in [FH13], the tropical fibre product is not a categorical pullback even in the underlying category of sets. Thus, we aim to extend the notion of morphism of tropical fans to something that does not have to be a map between their supports. This leads naturally to the definition of the tropical correspondences category (Definition 5.4). It formalizes the intuition behind “not exactly graphs” inspired by Example 5.1.

The whole category of tropical correspondences between Bergman fans is too unwieldy. In particular, since we allow the cones of the cycles to have negative weights, we cannot even require correspondences with negative weights to be irreducible. Thus, in Sections 3 and 5 we develop a subcategory of “covering graph correspondences” (Definition 5.6, Lemma 5.8) which is convenient from the combinatorial point of view. These covering graph correspondences between Bergman fans contain strictly more information than weak maps between the groundsets (Remark 5.17).

Theorem B (5.13).
{weak lattice map}{graph correspondence}\{\mbox{weak lattice map}\}\to\{\mbox{graph correspondence}\}

is a functor from the category of covering lattice maps of flats of simple matroids with the usual composition to the category of tropical correspondences between Bergman fans.

The paper is organized as follows. In Section 2, we recall necessary notions from matroid theory and tropical geometry, including the tropical fibre product from [FH13] — Definition 2.21. In Section 3, we introduce the generalization of Bergman fan called a Flag fan and observe its nice behavior with respect to Weil divisors of rational functions generalizing characteristic functions of modular cuts (Lemma 3.7). Section 4 is solely devoted to the proof of Theorem 4.1. In Section 5, we introduce and study tropical correspondences, construct covering graph correspondences, establish their useful combinatorial description (Theorem 5.12), and verify the functoriality (Theorem 5.13). We conclude with an open Question 5.19. Resolving it positively will achieve the best scenario still plausible — that there is a full forgetful functor from the reasonable subcategory of tropical correspondences onto the category of weak maps.

I am grateful to Evgeniya Akhmedova, Alexander Esterov, Tali Kaufman, Anna–Maria Raukh, and Kris Shaw for useful discussions. I want to thank Alexander Zinov for helping me with software calculations which often preceded proofs.

2. Preliminaries

2.1. Matroid theory

In this subsection we recall the necessary notions from the combinatorial matroid theory.

Definition 2.1.

A matroid MM on the groundset EE is a collection of subsets (M)=2E\mathcal{F}(M)=\mathcal{F}\subset 2^{E} called flats with the following properties:

  • EE\in\mathcal{F};

  • F1,F2F1F2F_{1},F_{2}\in\mathcal{F}\Rightarrow F_{1}\cap F_{2}\in\mathcal{F};

  • F:EF=FFFF\forall F\in\mathcal{F}\colon E\setminus F=\bigsqcup_{F^{\prime}\gtrdot F}F^{\prime}\setminus F, where XYX\gtrdot Y is read “XX covers YY” and means XY,Z:XZYX\supsetneq Y,\;\nexists Z\colon X\supsetneq Z\supsetneq Y.

A matroid is called loopless if \varnothing\in\mathcal{F}. It is called simple if, additionally, all flats covering \varnothing are one-element subsets (in this case all one-element subsets are flats). All matroids in this paper are assumed loopless. A restriction of the matroid MM on the groundset EE onto the subset TET\subset E, denoted by M|TM|_{T}, is an image of the map 2T\mathcal{F}\to 2^{T} given by FFTF\mapsto F\cap T. An extension of MM onto EEE^{\prime}\supset E is a matroid MM^{\prime} on EE^{\prime} such that M|E=MM^{\prime}|_{E}=M. A contraction of F(M)F\in\mathcal{F}(M) is a matroid denoted by M/FM/F on the groundset EFE\setminus F with flats FFF^{\prime}\setminus F for FF(M)F\subset F^{\prime}\in\mathcal{F}(M).

There is a large number of equivalent definitions of a matroid, usually called cryptomorphisms. Many of them can be found in [Oxl11], Chapter 1. It can be shown that \mathcal{F} forms a ranked poset with respect to the partial order given by set inclusion. Since intersection of flats is also a flat, there exists an operator of closure cl:2E2E\operatorname{cl}\colon 2^{E}\to 2^{E}, where clM(X)=FXF\operatorname{cl}_{M}(X)=\cap_{F\supset X}F. Moreover, for each pair of flats F1F_{1} and F2F_{2} there exists a unique minimal element FF\in\mathcal{F} such that FF1,F2F\supset F_{1},F_{2}, called join and denoted by F1F2F_{1}\vee F_{2}. The rank function can be extended from \mathcal{F} to the whole 2E2^{E} by defining

rkMX:=minFXrkF.\operatorname{rk}_{M}X\vcentcolon=\min_{F\supset X}{\operatorname{rk}F}.

A subset XEX\subset E is called independent if rkX=|X|\operatorname{rk}X=|X|. The rank of MM is defined to be rkME\operatorname{rk}_{M}E. The hyperplanes are flats of rank rkM1\operatorname{rk}M-1.

Throughout the paper, we sometimes describe matroids via pictures instead of listing their flats. The points correspond to the elements of the groundset, while lines and planes correspond to flats. Those are instances of matroids representable over \mathbb{R} drawn in the projectivization of the vector space of representation (more on that in [Oxl11], Chapter 1).

Next we prove a simple fact needed in Sections 4 and 5:

Lemma 2.2.

The difference of ranks can only decrease when adding elements. More precisely, if ABA\supset B, then

rkM(AC)rkM(BC)rkMArkMB.\operatorname{rk}_{M}(A\cup C)-\operatorname{rk}_{M}(B\cup C)\leqslant\operatorname{rk}_{M}A-\operatorname{rk}_{M}B.
Proof.

We can add elements of CC one by one, so take any xCx\in C and let us prove that

rkM(A{x})rkM(B{x})rkMArkMB.\operatorname{rk}_{M}(A\cup\{x\})-\operatorname{rk}_{M}(B\cup\{x\})\leqslant\operatorname{rk}_{M}A-\operatorname{rk}_{M}B.

The rank function can only increase by 1 when adding an element ([Oxl11], Section 1), and this happens exactly when this element does not belong to the closure of the set, so what we need follows from the implication

xclAxclB,x\notin\operatorname{cl}A\Rightarrow x\notin\operatorname{cl}B,

which is true, since ABclAclBA\supset B\Rightarrow\operatorname{cl}A\supset\operatorname{cl}B. ∎

Definition 2.3.

Consider sets E1,E2,T=E1E2E_{1},E_{2},T=E_{1}\cap E_{2} and matroids M1M_{1} on E1E_{1}, M2M_{2} on E2E_{2}, NN on TT such that M1|T=N=M2|TM_{1}|_{T}=N=M_{2}|_{T}. A matroid MM on E1E2E_{1}\cup E_{2} is called an amalgam of M1M_{1} and M2M_{2} along NN if M|E1=M1M|_{E_{1}}=M_{1} and M|E2=M2M|_{E_{2}}=M_{2}.

An amalgam MM is called free if for every amalgam MM^{\prime} and every independent set XX in it, XX is also independent in MM.

An amalgam MM is called proper if, for every F(M)F\in\mathcal{F}(M) the following holds:

rkM(F)=η(F):=rkM1(FE1)+rkM2(FE2)rkN(FT).\operatorname{rk}_{M}(F)=\eta(F)\vcentcolon=\operatorname{rk}_{M_{1}}(F\cap E_{1})+\operatorname{rk}_{M_{2}}(F\cap E_{2})-\operatorname{rk}_{N}(F\cap T).

Both free and proper amalgams are unique, if they exist. This is not always the case, though: sometimes there is no free amalgam, sometimes no amalgams at all (see [Oxl11], Section 11.4). It can be shown that the proper amalgam is always free ([Oxl11], Propositions 11.4.2 and 11.4.3).

Remark 2.4.

In [Oxl11] Proposition 11.4.2 defines proper amalgam as an amalgam where the rank of any set XEX\subset E satisfies rkX=min{η(Y):YX}\operatorname{rk}X=\min{\{\eta(Y)\colon Y\supseteq X\}}, and Proposition 11.4.3 establishes equivalence with the definition that we give in Definition 2.3.

The free amalgam is not always proper, though, as shown by the following

Example 2.5.
Refer to caption
Figure 1. MM is the free amalgam but not the proper.

For M1,M2,NM_{1},M_{2},N as shown on Figure 1 MM is the only amalgam and hence free, but

1=rkM{5,6}rkM1{5}+rkM2{6}rkN=21=\operatorname{rk}_{M}\{5,6\}\neq\operatorname{rk}_{M_{1}}\{5\}+\operatorname{rk}_{M_{2}}\{6\}-\operatorname{rk}_{N}\varnothing=2
Definition 2.6.

A map f:E1E2f\colon E_{1}\to E_{2} between groundsets of matroids M1M_{1} and M2M_{2} is called a strong map if f1(F)(M1)f^{-1}(F)\in\mathcal{F}(M_{1}) for F(M2)F\in\mathcal{F}(M_{2}). A strong map is called an embedding if M2|f(E1)=M1M_{2}|_{f(E_{1})}=M_{1}.

A map f:E1E2f\colon E_{1}\to E_{2} between groundsets of matroids M1M_{1} and M2M_{2} is called a weak map if for any XE1rkM1(X)rkM2(f(X))X\in E_{1}\operatorname{rk}_{M_{1}}(X)\geqslant\operatorname{rk}_{M_{2}}(f(X)). Equivalently ([Oxl11], Proposition 7.3.11), f1(X)f^{-1}(X) is independent in M1M_{1} for any independent set XX of M2M_{2}. We will sometimes write f:M1M2f\colon M_{1}\to M_{2} instead of f:E1E2f\colon E_{1}\to E_{2}.

It can be shown that strong maps are also weak maps ([Oxl11], Corollary 7.3.12). Note that the free amalgam MM has the following universal property in the category of matroids with weak maps. If ι1,2:NM1,2\iota_{1,2}\colon N\to M_{1,2} and j1,2:M1,2Mj_{1,2}\colon M_{1,2}\to M are embeddings, then

M,embeddings j1,2:M1,2M such that j1ι1=j2ι2! weak map f:MM such that fj1=j1,fj2=j2.\begin{split}\forall M^{\prime},\;\mbox{embeddings }\;j^{\prime}_{1,2}\colon M_{1,2}\to M^{\prime}\mbox{ such that }j^{\prime}_{1}\circ\iota_{1}=j^{\prime}_{2}\circ\iota_{2}\\ \exists!\;\mbox{ weak map }\;f\colon M\to M^{\prime}\mbox{ such that }f\circ j_{1}=j^{\prime}_{1},f\circ j_{2}=j^{\prime}_{2}.\end{split} (2.1)

It is not, strictly speaking, a universal property of the pushout, because allowing j1,2j^{\prime}_{1,2} to be any weak maps may lead to rank of NN dropping, and then MM^{\prime} may become “freer” than any amalgam.

Definition 2.7.

A pair of flats F1,F2(M)F_{1},F_{2}\in\mathcal{F}(M) is called a modular pair if

rkF1+rkF2=rk(F1F2)+rk(F1F2).\operatorname{rk}{F_{1}}+\operatorname{rk}{F_{2}}=\operatorname{rk}{(F_{1}\vee F_{2})}+\operatorname{rk}{(F_{1}\wedge F_{2})}.

The flat F(M)F\in\mathcal{F}(M) is called modular if it forms a modular pair with any F(M)F^{\prime}\in\mathcal{F}(M).

Definition 2.8.

A subset \mathcal{M} of flats of MM on groundset EE is called a modular cut if it satisfies the following properties:

  • EE\in\mathcal{M};

  • F1,F2F1F2F_{1}\in\mathcal{M},F_{2}\ni F_{1}\Rightarrow F_{2}\in\mathcal{M};

  • if F1,F2F_{1},F_{2}\in\mathcal{M} are a modular pair, then F1F2F_{1}\wedge F_{2}\in\mathcal{M}.

2.2. Tropical geometry

In this subsection we recall the tropical counterpart of matroid theory involving the techniques needed for our results.

Definition 2.9.

A (reduced or projective) Bergman fan of matroid MM on groundset EE is a collection of maximal cones σ𝐅\sigma_{\mathbf{F}} and their faces in E/(1,,1)\mathbb{R}^{E}/\langle(1,\ldots,1)\rangle for each flag of flats 𝐅\mathbf{F} consisting of =F0F1Fr=E\varnothing=F_{0}\lessdot F_{1}\lessdot\ldots\lessdot F_{r}=E with Fi(M)F_{i}\in\mathcal{F}(M) defined as

σ𝐅=eF1,,eFr1,\sigma_{\mathbf{F}}=\langle e_{F_{1}},\ldots,e_{F_{r-1}}\rangle,

where eXe_{X} stands for the image in E/(1,,1)\mathbb{R}^{E}/\langle(1,\ldots,1)\rangle of the characteristic function of XEX\subset E in E\mathbb{R}^{E}, i.e., eX(x)=1e_{X}(x)=1 if xXx\in X and eX(x)=0e_{X}(x)=0 otherwise.

A non-reduced or affine Bergman fan, denoted by B(M)B(M), is the preimage of the projective Bergman fan in E\mathbb{R}^{E}.

An affine Bergman fan is an instance of the main object we work with in this paper.

Definition 2.10 ([AR09], Definition 2.6).

A kk-dimensional tropical fan XX in n=n\mathbb{R}^{n}=\mathbb{Z}^{n}\otimes_{\mathbb{Z}}\mathbb{R} is a collection of cones σ\sigma of dimension kk, generated by the vectors from the lattice n\mathbb{Z}^{n}, with integer weights ω(σ)\omega(\sigma) such that the balancing condition holds for each (k1)(k-1)-dimensional face τ\tau:

στω(σ)uσ/τ=0V/Vτ,\sum_{\sigma\supset\tau}\omega(\sigma)\cdot u_{\sigma/\tau}=0\in V/V_{\tau},

where VρV_{\rho} is the smallest subspace of n\mathbb{R}^{n} containing cone ρ\rho, and uρ/ρu_{\rho/\rho^{\prime}} is a normal vector — the generator of the one-dimensional \mathbb{Z}-module (Vρn)/(Vρn)(V_{\rho}\cap\mathbb{Z}^{n})/(V_{\rho^{\prime}}\cap\mathbb{Z}^{n}) chosen such that it “looks inside” ρ\rho. The support of the fan XX, denoted by |X||X|, is the set of points in n\mathbb{R}^{n} that belong to some cone.

In all our considerations all the cones are going to be simplicial. Moreover, the only generator of cone ρ\rho missing from the generators of ρ\rho^{\prime} is always going to map to uρ/ρu_{\rho/\rho^{\prime}} when taking quotient by VρV_{\rho^{\prime}} — not some multiple of uρ/ρu_{\rho/\rho^{\prime}}, which means that this generator can be taken as a representative in all calculations.

The maximal cones of the affine Bergman fan are the following:

eF1,,eFr1,eFr=(1,,1),eF1,,eFr1,eFr=(1,,1),\langle e_{F_{1}},\ldots,e_{F_{r-1}},e_{F_{r}}=(1,\ldots,1)\rangle,\langle e_{F_{1}},\ldots,e_{F_{r-1}},-e_{F_{r}}=(-1,\ldots,-1)\rangle,

and the weights are all equal to 1 (slightly abusing notations, we now mean by eFe_{F}’s actual characteristic vectors in E\mathbb{R}^{E}, not their classes in the quotient E/(1,,1)\mathbb{R}^{E}/\langle(1,\ldots,1)\rangle). Thus, one can check that the balancing condition for any (r1)(r-1)-dimensional cone containing the all-ones or minus all-ones boils down to the third axiom (the covering axiom) of Definition 2.1; and the balancing condition for a (r1)(r-1)-dimensional cone not containing the all-ones or minus all-ones is trivially satisfied, as any such cone is contained in just two rr-dimensional cones which cancel each other out in the balancing condition.

It is convenient to consider the refinements of tropical fans ([AR09], Definition 2.8), which are essentially the tropical fans with the same support and the same weight for the interior points of maximal cones. This leads to

Definition 2.11 ([AR09], Definitions 2.12, 2.15).

A tropical cycle is a class of equivalence of tropical fans with respect to equivalence relation “to have a common refinement”. A tropical subcycle ZZ of the tropical cycle XX is a tropical cycle whose support is a subset: |Z||X||Z|\subset|X|.

Apart from the very general Definition 5.4, all the cones in the subcycles will be the faces of the simplicial maximal cones of the fans they lie in.

Considering tropical cycles allows to define new tropical cycles without explicitly listing all the cones, but describing the support and the weights of the interior points of the maximal cones instead. This is useful to us when introducing products and stars.

Definition 2.12.

The product of tropical fans XVXX\in V_{X} and YVYY\in V_{Y} is a tropical cycle X×YVX×VYX\times Y\in V_{X}\times V_{Y} such that |X×Y|=|X|×|Y||X\times Y|=|X|\times|Y|, and the weight of (x,y)X×Y(x,y)\in X\times Y is equal to the product of xXx\in X and yYy\in Y for x,yx,y interior points of the maximal cones of X,YX,Y.

It is known that the tropical cycles B(MN)B(M\oplus N) and B(M)×B(N)B(M)\times B(N) coincide ([FR12], Lemma 2.1), but the fans are different — B(MN)B(M\oplus N) is a refinement of B(M)×B(N)B(M)\times B(N) (see also Lemma 3.10).

Definition 2.13.

Let XnX\subset\mathbb{R}^{n} be a tropical fan and pnp\in\mathbb{R}^{n} a point (not necessarily in XX). The star of pp in XX denoted by StarX(p)\mathrm{Star}_{X}(p) is a tropical fan with support

{vn:c>0 0<c<cp+cv|X|}\{v\in\mathbb{R}^{n}\colon\exists\;c>0\;\forall\;0<c^{\prime}<c\;\;p+c^{\prime}\cdot v\in|X|\}

and the weight of vv equal to the weight of p+cvp+c^{\prime}\cdot v for interior points of maximal cones. In particular, the star is empty if and only if pp does not belong to the support of XX.

If B(M)B(M) is a Bergman fan, and pp is an interior point of the cone of any dimension generated by the flag =F0F1Fk=E\varnothing=F_{0}\subsetneq F_{1}\subsetneq\ldots F_{k}=E of flats Fi(M)F_{i}\in\mathcal{F}(M), then the explicit description of the StarB(M)(p)\mathrm{Star}_{B(M)}(p) is known (see, for example, [FR12], Lemma 2.2):

B(M|F1/F0)×B(M|F2/F1)×B(M|Fk/Fk1).B(M|_{F_{1}}/F_{0})\times B(M|_{F_{2}}/F_{1})\ldots\times B(M|_{F_{k}}/F_{k-1}).

The crucial ingredient of the tropical intersection theory is the possibility to cut out Weil divisors of what is called “rational functions” on tropical fans.

Definition 2.14 ([AR09], Section 3).

A rational function on the tropical fan is a continuous function on the union of its maximal cones which is linear on each of them. A Weil divisor of the rational function φ\varphi on the kk-dimensional tropical fan XX is a (k1)(k-1)-dimensional tropical subfan denoted by φX\varphi\cdot X with the weight on the (k1)(k-1)-dimensional cone τ\tau of XX given by

ωφX(τ)=στωX(σ)φ(vσ/τ)φ(στωX(σ)vσ/τ),\omega_{\varphi\cdot X}(\tau)=\sum_{\sigma\supset\tau}\omega_{X}(\sigma)\varphi(v_{\sigma/\tau})-\varphi\left(\sum_{\sigma\supset\tau}\omega_{X}(\sigma)v_{\sigma/\tau}\right),

where vσ/τv_{\sigma/\tau} is any representative of the primitive vector uσ/τu_{\sigma/\tau} in the ambient space. One shows that this definition does not depend on the choice of representatives.

More elegant, but not as useful in calculations, is the equivalent definition of associated Weil divisor — that it is the n\mathbb{R}^{n}-section of the only way to “balance” the graph of φ\varphi in n+1=n×\mathbb{R}^{n+1}=\mathbb{R}^{n}\times\mathbb{R}, i.e., to add cones containing direction along the last coordinate corresponding to the value of φ\varphi with some weights, so that the disbalance brought by the non-linearity of φ\varphi on the hyperface junctions of maximal cones is fixed. Thus, the divisor has zero weights if the function is globally linear (mind that the converse is not true, but the counter-examples are not going to occur in this paper).

In the general setting, it is common to take advantage of the opportunity to refine the fan, then take the fine rational function which is not conewise linear before the refinement and consider this to be a rational function on the tropical cycle given by the equivalence class of the fan. In our setting, though, values on the primitive vectors of the generating rays of cones (extended conewise linearly) always suffice for the functions at play; so we do not concern ourselves with the details and refer for them to [AR09]. We may also sometimes abuse the notation and say “value on F” meaning value on the primitive vector of the one-dimensional cone eF\langle e_{F}\rangle which is eFe_{F} itself.

Finally, because of the technical necessity to work with affine fans, all our functions are assumed to be linear on the lineality space (1,1)\langle(1\ldots,1)\rangle, which means that φ((1,1))=φ((1,1))\varphi((-1\ldots,-1))=-\varphi((1\ldots,1)). This guarantees that all the maximal cones of the Weil divisors contain either eEe_{E} or eE-e_{E} as well. Henceforth, when verifying something about the fan, we do it only for positive cones, assuming that negative cones are treated exactly the same.

The most ubiquitous rational function α\alpha is described in the following

Definition 2.15.

A truncation of the Bergman fan B(M)B(M) is the Weil divisor of the function α\alpha which is equal to 1-1 on EE and 0 on every other ray. It can be shown (a simple partial case of Lemma 3.1 ahead) that the resulting fan is again a Bergman fan of the truncated matroid, denoted by Tr(M)\mathrm{\mathop{Tr}}(M), whose flats are the same as of MM except for the hyperplanes.

Cutting out a divisor of α\alpha is analogous in tropical geometry to taking a general hyperplane section in algebraic geometry, and thus

Definition 2.16.

The degree of the kk-dimensional affine tropical fan XX is the weight of the cone eE\langle e_{E}\rangle in the fan ααk1X\underbrace{\alpha\cdots\alpha}_{k-1}\cdot X.

It is easy to see that Bergman fans have degree 1. Moreover,

Theorem 2.17 ([Fin13], Theorem 6.5).

Tropical fan of degree 1 with positive weights is a Bergman fan of matroid.

Theorem 2.17 was proven in [Fin13] in more general setting, for tropical varieties, not only for fans, and may be considered an overkill for our needs, where we could use techniques developed in Section 3 instead, but we believe it is a beautiful result, and we are happy to apply it when relevant. Note that without requiring the weights to be positive this is not true, of course, as we are going to see many times.

To define tropical fibre product and to work with tropical intersections effectively, it remains to recall the definition of tropical morphism, the constructions of the pullback and the push-forward.

Definition 2.18 ([AR09], Definition 4.1).

A morphism f:XYf\colon X\to Y between tropical fans XmX\subset\mathbb{Z}^{m}\otimes_{\mathbb{Z}}\mathbb{R} and YnY\subset\mathbb{Z}^{n}\otimes_{\mathbb{Z}}\mathbb{R} is a map from the support of XX to the support of YY induced by a \mathbb{Z}-linear map f~:mn\tilde{f}\colon\mathbb{Z}^{m}\to\mathbb{Z}^{n}.

In full generality, the maximal cone σ\sigma of XX maps to some subset f(σ)σf(\sigma)\subset\sigma^{\prime} of YY. For instance, the embedding of the refinement into the unrefined fan is a tropical morphism. For tropical morphisms we consider, though, such thing is not going to happen, as generating rays of cones of XX always map to the generating rays of cones of YY. This observation also makes the next definitions simpler.

Definition 2.19 ([AR09], Proposition 4.7).

Let f:XYf\colon X\to Y be a morphism of tropical fans, and let φ\varphi be a rational function on YY with φY\varphi\cdot Y its Weil divisor. Then the pullback of φ\varphi is a function f(φ):Xf^{*}(\varphi)\colon X\to\mathbb{R} defined as f(φ)(x)=φ(f(x))f^{*}(\varphi)(x)=\varphi(f(x)), and the pullback of the Weil divisor φY\varphi\cdot Y is the Weil divisor f(φ)Xf^{*}(\varphi)\cdot X.

One shows that pullback is a linear map from divisors on YY to divisors on XX.

Definition 2.20.

Let f:XYf\colon X\to Y be a morphism of tropical fans lying in spaces m\mathbb{Z}^{m}\otimes_{\mathbb{Z}}\mathbb{R} and n\mathbb{Z}^{n}\otimes_{\mathbb{Z}}\mathbb{R}, respectively. Let ZXZ\subset X be a subcycle. Then, the push-forward f(Z)f_{*}(Z) is defined as a tropical subcycle of YY consisting of maximal cones f(σ)f(\sigma) for σ\sigma the cone of XX if ff is injective on σ\sigma (and their faces). The weight on the maximal cone should be defined by the formula

ωfZ(σ)=f(σ)=σωZ(σ)λ(σ,σ),\omega_{f_{*}Z}(\sigma^{\prime})=\sum_{f(\sigma)=\sigma^{\prime}}\omega_{Z}(\sigma)\cdot\lambda(\sigma^{\prime},\sigma),

where λ(σ,σ)\lambda(\sigma^{\prime},\sigma) is the index of the sublattice f(Vσm)f(V_{\sigma}\cap\mathbb{Z}^{m}) in the lattice f(Vσn)f(V_{\sigma^{\prime}}\cap\mathbb{Z}^{n}) (in the notation of Definition 2.10).

Fortunately, again, in our setting the integer lattice map is always surjective, therefore, the index of the sublattice is always 11. The sum in the definition of the weight is important, though, as there can be several maximal cones of XX mapping to the same maximal cone of YY.

Now we are almost ready to recall the main notion for Theorem 4.1 — the tropical fibre product. Its only ingredient not covered so far is the construction of “diagonal” rational functions from Definition 3.3, which arise as the partial case of more general construction from Lemma 3.1. For the clarity of exposition, the definition is postponed to Section 3, so the reader can either look at functions φi\varphi_{i} as a black box for now, or comprehend their definition as a separate construction.

Definition 2.21 ([FH13], Definition 3.6).

Let M1M_{1} and M2M_{2} be matroids on groundsets E1E_{1} and E2E_{2}, respectively. Let NN be their common restriction, i.e., a matroid on groundset T=E1E2T=E_{1}\cap E_{2} such that M1|T=N=M2|TM_{1}|_{T}=N=M_{2}|_{T}. Let π1:B(M1)B(N)\pi_{1}\colon B(M_{1})\to B(N) and π2:B(M2)B(N)\pi_{2}\colon B(M_{2})\to B(N) be projections induced by the embeddings TE1T\hookrightarrow E_{1} and TE2T\hookrightarrow E_{2}. Let φ1φrB(N)×B(N)=ΔB(N)×B(N)\varphi_{1}\cdots\varphi_{r}\cdot B(N)\times B(N)=\Delta\subset B(N)\times B(N), like in Definition 3.3, and let π=π1×π2\pi=\pi_{1}\times\pi_{2}. Then, the tropical fibre product B(M1)×π1,B(N),π2B(M2)B(M_{1})\times_{\pi_{1},B(N),\pi_{2}}B(M_{2}) is a cycle in B(M1)×B(M2)B(M_{1})\times B(M_{2}) defined as

π(φ1)π(φr)B(M1)×B(M2)B(M1)×B(M2).\pi^{*}(\varphi_{1})\cdots\pi^{*}(\varphi_{r})\cdot B(M_{1})\times B(M_{2})\subset B(M_{1})\times B(M_{2}).
Example 2.22.

We consider three instances of the tropical fibre product to familiarize the reader with the notion:

  1. (1)

    Assume TT is a modular flat of one of the matroids, say, M1M_{1}. In this case, π1:B(M1)B(N)\pi_{1}\colon B(M_{1})\to B(N) is not only surjective, but locally surjective, a term defined in [FH13]. It means that for any p|B(M1)|p\in|B(M_{1})|, the restriction of π1\pi_{1} on some neighborhood of pp is surjective onto some neighborhood of π1(p)|B(N)|\pi_{1}(p)\in|B(N)|. This, in its turn, is equivalent to the projection π1\pi_{1} being surjective on StarB(N)(π1(p))\mathrm{Star}_{B(N)}(\pi_{1}(p)). By description of stars of Bergman fans (paragraph after Definition 2.13) we need to show that for any pair of flats F′′FF^{\prime\prime}\supset F^{\prime} of M1M_{1} the restriction map

    (M1|F′′/F)(M1|TF′′/(TF)),FFT\mathcal{F}(M_{1}|_{F^{\prime\prime}}/F^{\prime})\to\mathcal{F}(M_{1}|_{T\cap F^{\prime\prime}}/(T\cap F^{\prime})),\;F\mapsto F\cap T

    is surjective. We show that for the flat FNF_{N} of NN containing FTF^{\prime}\cap T the preimage is, for example, clM1(FFN)\operatorname{cl}_{M_{1}}(F^{\prime}\cup F_{N}). Indeed, using modularity of TT,

    rkM1(FFN)=rkM1F+rkM1FNrkM1FT,\operatorname{rk}_{M_{1}}(F^{\prime}\vee F_{N})=\operatorname{rk}_{M_{1}}{F^{\prime}}+\operatorname{rk}_{M_{1}}{F_{N}}-\operatorname{rk}_{M_{1}}{F^{\prime}\cap T},

    and then

    rkM1((FFN)T)=rkM1(FFN)+rkM1TrkM1FT=rkM1F++rkM1FNrkM1FT+rkM1TrkM1FrkM1T+rkM1FT=rkM1FN,\begin{split}\operatorname{rk}_{M_{1}}((F^{\prime}\vee F_{N})\cap T)=\operatorname{rk}_{M_{1}}(F^{\prime}\vee F_{N})+\operatorname{rk}_{M_{1}}{T}-\operatorname{rk}_{M_{1}}{F^{\prime}\vee T}=\operatorname{rk}_{M_{1}}{F^{\prime}}+\\ +\operatorname{rk}_{M_{1}}{F_{N}}-\operatorname{rk}_{M_{1}}{F^{\prime}\cap T}+\operatorname{rk}_{M_{1}}{T}-\operatorname{rk}_{M_{1}}{F^{\prime}}-\operatorname{rk}_{M_{1}}{T}+\operatorname{rk}_{M_{1}}{F^{\prime}\cap T}=\operatorname{rk}_{M_{1}}{F_{N}},\end{split}

    so (FFN)T=FN(F^{\prime}\vee F_{N})\cap T=F_{N} as claimed. In Theorem 3.9 of [FH13] it is shown that when one of the projections is locally surjective, the tropical fibre product is a tropical fan with positive weights. Moreover, its support is equal to the set-theoretic fibre product of B(M1)B(M_{1}) and B(M2)B(M_{2}) over B(N)B(N).

    Refer to caption
    Figure 2. B(M)B(M) is the tropical fibre product B(M1)×π1,B(N),π2B(M2)B(M_{1})\times_{\pi_{1},B(N),\pi_{2}}B(M_{2}).
  2. (2)

    Consider M1,M2,NM_{1},M_{2},N as shown on Figure 2. In this case one verifies that the tropical fibre product coincides with B(M)B(M) (the elements 11 and 22 of MM should be thought of as pairs of respective parallel elements of M1M_{1} and M2M_{2} — more precisely in Theorem 4.1). Observe that MM is the proper amalgam of M1M_{1} and M2M_{2}. Mind that the set-theoretic product B(M1)×π1,B(N),π2B(M2)B(M_{1})\times_{\pi_{1},B(N),\pi_{2}}B(M_{2}) is not even a pure-dimensional fan and, therefore, certainly not tropical. It happens because, unlike part (1) of this example, the projections B(M1)B(N)B(M_{1})\to B(N) and B(M2)B(N)B(M_{2})\to B(N) are not locally surjective on the rays corresponding to flats {3}\{3\} and {4}\{4\} respectively. Projective fans illustrating this phenomenon are shown on Figure 3.

    Refer to caption
    Figure 3. Set-theoretic fibre product is in black and gray, B(M)B(M) is in red.
  3. (3)

    Consider M1,M2,NM_{1},M_{2},N as in Example 2.5. Then, it can be calculated that, after taking two out of three Weil divisors, π(φ2)π(φ1)(B(M1)×B(M2))=B(M)\pi^{*}(\varphi_{2})\cdot\pi^{*}(\varphi_{1})\cdot(B(M_{1})\times B(M_{2}))=B(M^{\prime}), where MM^{\prime} is shown on Figure 4. Then, π(φ3)({11,12,31,32,5,6})=π(φ3)({21,22,41,42,5,6})=1\pi^{*}(\varphi_{3})(\{1_{1},1_{2},3_{1},3_{2},5,6\})=\pi^{*}(\varphi_{3})(\{2_{1},2_{2},4_{1},4_{2},5,6\})=-1, but π(φ3)({5,6})=0\pi^{*}(\varphi_{3})(\{5,6\})=0, despite {11,12,31,32,5,6}\{1_{1},1_{2},3_{1},3_{2},5,6\} and {21,22,41,42,5,6}\{2_{1},2_{2},4_{1},4_{2},5,6\} being a modular pair. Thus, π(φ3)\pi^{*}(\varphi_{3}) is not defined by the modular cut of MM^{\prime} (see Lemma 3.1), and the tropical fibre product has two cones with weight 1-1, corresponding to the flags {5}{5,6}E1E2\varnothing\subset\{5\}\subset\{5,6\}\subset E_{1}\sqcup E_{2} and {6}{5,6}E1E2\varnothing\subset\{6\}\subset\{5,6\}\subset E_{1}\sqcup E_{2}.

    Refer to caption
    Figure 4. Indices discern between elements of E1E_{1} and E2E_{2} in E1E2E_{1}\sqcup E_{2}.
Lemma 2.23.

Restriction commutes with truncation (preserving dimension). More precisely, let MM be a matroid on groundset EE with rkM=r\operatorname{rk}{M}=r. let TET\subset E and let M|T=NM|_{T}=N with rkN=s\operatorname{rk}N=s. Let π:B(M)B(N)\pi\colon B(M)\to B(N) be a projection induced by the embedding TET\hookrightarrow E. Let αM\alpha_{M} be a truncation rational function in E\mathbb{R}^{E} defined in 2.15, and let ZM=αMriB(M)Z_{M}=\alpha_{M}^{r-i}\cdot B(M) be an ii-dimensional truncation of B(M)B(M). Similarly, let αN\alpha_{N} be a truncation function in T\mathbb{R}^{T}, and let ZN=αNsiB(N)Z_{N}=\alpha_{N}^{s-i}\cdot B(N) be an ii-dimensional truncation of B(N)B(N). Then, π(ZM)=ZN\pi_{*}(Z_{M})=Z_{N}.

Proof.

Since ZM=B(Trri(M))Z_{M}=B(\mathop{\mathrm{Tr}}^{r-i}(M)), maximal cones of ZMZ_{M} are generated by characteristic rays of flags of the form F1Fi1Fi=E\varnothing\lessdot F_{1}\lessdot\ldots\lessdot F_{i-1}\leqslant F_{i}=E. The image of such a cone under π\pi is ii-dimensional if and only if all FiTF_{i}\cap T are different flats of B(Trsi(N))B(\mathop{\mathrm{Tr}}^{s-i}(N)). Any such flag can be restored from the flag F1Fi1Fi=T\varnothing\lessdot F^{\prime}_{1}\lessdot\ldots\lessdot F^{\prime}_{i-1}\leqslant F^{\prime}_{i}=T of flats in NN by defining Fi=clMFiF_{i}=\operatorname{cl}_{M}{F^{\prime}_{i}}. Thus, there is a bijection between the maximal cones of ZNZ_{N} and maximal cones of ZMZ_{M} with zero intersection with kerπ\ker{\pi}. All the indices of the respective lattices are equal to 11, as always, as each lattice is generated by the primitive vectors of the generating rays. ∎

Finally, we will need a few more statements providing tying tropical intersections in related objects — the domain and the range of the tropical morphism; or the tropical fan and its star.

Theorem 2.24 (Projection formula, [AR09], Proposition 4.8).

Let f:XYf\colon X\to Y be a morphism of tropical fans. Let CC be a subcycle of XX, and φ\varphi a rational function on YY. Then,

φf(C)=f(f(φ)C).\varphi\cdot f_{*}(C)=f_{*}(f^{*}(\varphi)\cdot C).

If φ\varphi is a rational function on XX, then the rational function φp\varphi^{p} on the star StarX(p)\mathrm{Star}_{X}(p) is defined as the linear extension of the restriction of φ\varphi onto the small neighborhood of pp.

Lemma 2.25.

Taking the star commutes with cutting out Weil divisor. More precisely, if XX is a tropical fan, pp a point and φ\varphi a rational function, then

StarφX(ρ)=φρStarX(ρ).\mathrm{Star}_{\varphi\cdot X}(\rho)=\varphi^{\rho}\cdot\mathrm{Star}_{X}(\rho).
Lemma 2.26 ([FR12], Definition 8.1).

Given morphism of Bergman fans f:XYf\colon X\to Y, if ZYZ\subset Y is a subcycle obtained from taking Weil divisors of functions φi\varphi_{i}, then the support of f(φi)X\prod f^{*}(\varphi_{i})\cdot X is contained in f1(|Z|)f^{-1}(|Z|).

Remark 2.27.

Let us explain Lemma 2.26 via a more general setting of intersection theory in Bergman fans. In [FR12], Section 4, the intersection CDC\cdot D of arbitrary subcycles of Bergman fans is defined using diagonal construction from Definition 3.3 below. It is shown there in Theorem 4.5 that the support of CDC\cdot D is contained in the intersection of the supports of CC and DD. Then, in Definition 8.1, a pullback of the arbitrary cycle DB(N)D\subset B(N) with respect to the morphism f:B(M)B(N)f\colon B(M)\to B(N) is defined as the intersection in B(M)×B(N)B(M)\times B(N) of the cycle Γf\Gamma_{f} — the graph of ff defined as the push-forward; and the cycle B(M)×DB(M)\times D; this intersection then push-forwarded onto the first factor XX of X×YX\times Y. This construction of the pullback coincides with the one we use for those cycles DB(N)D\subset B(N) that are cut out by rational functions (see Example 8.2 of [FR12]). Hence, the comment after Definition 8.1 of [FR12] claims that the support of fDf^{*}D is contained in the preimage of the support of DD.

3. Flag fans

In this section we introduce the generalization of Bergman fan which we call a Flag fan. This class of tropical fans is stable under taking Weil divisors with respect to certain rational functions which we call (simply, not modular) cut functions.

We begin by recalling the standard setting of modular cuts and tropical modifications. The following lemma is a summary of results established at approximately the same time by Shaw, Francois and Rau in [Sha13] and [FR12].

Lemma 3.1.

Let MM be a matroid, B(M)B(M) its Bergman fan, (M)\mathcal{M}\subset\mathcal{F}(M) a modular cut. Define rational function

φ(eF)={1,if F,0,otherwise.\varphi_{\mathcal{M}}(e_{F})=\begin{cases}-1,&\mbox{if $F\in\mathcal{M}$},\\ 0,&\mbox{otherwise.}\end{cases}

Assume \mathcal{M} does not contain flats of rank 1. Then φB(M)=B(N)\varphi_{\mathcal{M}}\cdot B(M)=B(N), where F(N)F\in\mathcal{F}(N) if and only if

{F(M)[FF for any FF\begin{cases}F\in\mathcal{F}(M)\\ \left[\begin{array}[]{ll}F\in\mathcal{M}\\ F^{\prime}\notin\mathcal{M}\mbox{ for any }F^{\prime}\gtrdot F\end{array}\right.\end{cases}

Moreover, the balanced graph of φ:B(M)\varphi_{\mathcal{M}}\colon B(M)\to\mathbb{R}, denoted by ΓφE×\Gamma_{\varphi}\in\mathbb{R}^{E}\times\mathbb{R}, which is the union of the set-theoretic graph and cones containing generator along the last coordinate (see Definition 2.14), is equal to the Bergman fan B(M)B(M^{\prime}), where MM^{\prime} is a one-element extension of MM on the groundset E{e}E\cup\{e\}, and F{e}(M)F\cup\{e\}\in\mathcal{F}(M^{\prime}) if and only if FF\in\mathcal{M}. Matroid NN is, in its turn, the contraction of the element ee of MM^{\prime} (see Definition 2.1).

In the other direction, if |B(N)||B(M)||B(N)|\subset|B(M)| is a subfan of codimension 11, then there exists a modular cut \mathcal{M} such that φB(M)=B(N)\varphi_{\mathcal{M}}\cdot B(M)=B(N). Function φ\varphi_{\mathcal{M}} can be defined on the rays of B(M)B(M) as

φ(eF)=rkNFrkMF.\varphi_{\mathcal{M}}(e_{F})=\operatorname{rk}_{N}{F}-\operatorname{rk}_{M}{F}.

More generally, if |B(N)||B(M)||B(N)|\subset|B(M)| is a subfan of codimension ss, then there exists a sequence of functions φi=φi\varphi_{i}=\varphi_{\mathcal{M}_{i}} corresponding to the modular cuts i\mathcal{M}_{i} such that i=1sφiB(M)=B(N)\prod_{i=1}^{s}{\varphi_{i}}\cdot B(M)=B(N). Mind that j\mathcal{M}_{j} is a modular cut of i=1j1φiB(M)\prod_{i=1}^{j-1}{\varphi_{i}}\cdot B(M), not necessarily of B(M)B(M). Functions φi\varphi_{i} can be defined on the rays of B(M)B(M) as

φi(eF)={0,if rkN(F)+sirkM(F)1,otherwise.\varphi_{i}(e_{F})=\begin{cases}0,&\mbox{if $\operatorname{rk}_{N}(F)+s-i\geqslant\operatorname{rk}_{M}(F)$}\\ -1,&\mbox{otherwise.}\end{cases} (3.1)

In other words,

i=1sφi(F)=rkNFrkMF, and 1φi(F)φj(F)0 for i>j.\sum_{i=1}^{s}\varphi_{i}(F)=\operatorname{rk}_{N}{F}-\operatorname{rk}_{M}{F},\;\mbox{ and }\;-1\leqslant\varphi_{i}(F)\leqslant\varphi_{j}(F)\leqslant 0\;\mbox{ for }\;i>j.
Remark 3.2.

As shown in [AR09], Proposition 3.7 (b), the order of taking Weil divisors with respect to functions φi\varphi_{i} does not matter. We are going to stick to the non-increasing value order, though, as it is easier to track the result. For example, if, say, φ2\varphi_{2} is not a characteristic function of a modular cut of the initial matroid MM, but of a modular cut on φ1B(M)\varphi_{1}\cdot B(M), then φ1φ2B(M)=φ2φ1B(M)\varphi_{1}\cdot\varphi_{2}\cdot B(M)=\varphi_{2}\cdot\varphi_{1}\cdot B(M) has positive weights, while φ2B(M)\varphi_{2}\cdot B(M) does not.

The next definition we recall is the partial case of Lemma 3.1 and the main advancement of [FR12] — cutting out the set-theoretic diagonal allowed to define intersection of arbitrary subcycles of Bergman fans.

Definition 3.3 ([FR12], Corollary 4.2).

Given a Bergman fan B(N)B(N), the functions φN,i=φi\varphi_{N,i}=\varphi_{i} cutting the diagonal subcycle ΔB(N)×B(N)\Delta\subset B(N)\times B(N) are the piecewise linear functions on B(N)×B(N)B(N)\times B(N) such that φ1φrB(N)×B(N)=ΔB(N)×B(N)\varphi_{1}\cdots\varphi_{r}\cdot B(N)\times B(N)=\Delta\subset B(N)\times B(N). They are defined by their values on rays F=F1F2F=F_{1}\sqcup F_{2} by

φi(eF)={0, if rkN(F1F2)+rirkN(F1)+rkN(F2)1, otherwise.\varphi_{i}(e_{F})=\begin{cases}0,\mbox{ if $\operatorname{rk}_{N}(F_{1}\cup F_{2})+r-i\geqslant\operatorname{rk}_{N}(F_{1})+\operatorname{rk}_{N}(F_{2})$}\\ -1,\mbox{ otherwise.}\end{cases}

Let us now relax the requirements for the function by cancelling the modularity condition.

Definition 3.4.

Consider the pair (𝒢,ω𝒢)(\mathcal{G},\omega_{\mathcal{G}}) that consists of the ranked poset 𝒢\mathcal{G} of subsets of EE (,E𝒢\varnothing,E\in\mathcal{G}, also if X𝒢XX\leqslant_{\mathcal{G}}X^{\prime}, then XXX\subset X^{\prime}, but not vice versa) and of the edge weight function w𝒢:{(F,F)𝒢|F𝒢F}w_{\mathcal{G}}\colon\{(F,F^{\prime})\in\mathcal{G}|\;F\lessdot_{\mathcal{G}}F\}\to\mathbb{Z}. Then the Flag fan corresponding to the pair (𝒢,ω𝒢)(\mathcal{G},\omega_{\mathcal{G}}) is a tropical fan in E\mathbb{R}^{E} such that each cone is generated by the characteristic vectors of the set of pairwise comparable elements of 𝒢\mathcal{G} (which we call a flag, as for flats of matroids), and the weight on the maximal cones is given by

w(F0Fr)=i=0r1w𝒢(Fi,Fi+1).w(F_{0}\lessdot\ldots\lessdot F_{r})=\prod_{i=0}^{r-1}w_{\mathcal{G}}(F_{i},F_{i+1}).

As usual, the copies of cones with eEe_{E} replaced with eE-e_{E} are added with the same weights.

Definition 3.5.

A cut on a ranked poset 𝒢\mathcal{G} is a subset of flats 𝒜\mathcal{A} such that if FF𝒜F^{\prime}\supset F\in\mathcal{A}, then F𝒜F^{\prime}\in\mathcal{A}. A cut function φ𝒜\varphi_{\mathcal{A}} on the Flag fan of 𝒢\mathcal{G} is a piecewise linear continuation of χ𝒜-\chi_{\mathcal{A}}, a function equal to 1-1 on the rays eFe_{F} for F𝒜F\in\mathcal{A} and equal to 0 on the other rays (except, as always, eE-e_{E}, where the value must be linear extension of φ𝒜(eE)\varphi_{\mathcal{A}}(e_{E})).

Example 3.6.

Here are a few examples of Flag fans to get used to them:

  1. (1)

    Bergman fan B(M)B(M) is a Flag fan corresponding to the pair ((M),ω(M))(\mathcal{F}(M),\omega_{\mathcal{F}(M)}), where
    ω(M)(F,F)=1\omega_{\mathcal{F}(M)}(F,F^{\prime})=1 for any FMFF\lessdot_{M}F^{\prime}.

  2. (2)

    Consider 𝒢\mathcal{G} consisting of ,{1,2,3}\varnothing,\{1,2,3\} and sets {1},{1,2},{1,3}\{1\},\{1,2\},\{1,3\} all covering \varnothing and all covered by {1,2,3}\{1,2,3\}. Note that {1}{1,2},{1,3}\{1\}\subset\{1,2\},\{1,3\}, but they all have rank 11 in 𝒢\mathcal{G}. Let all ω𝒢(X,X)=1\omega_{\mathcal{G}}(X,X^{\prime})=1 except for ω𝒢({1},{1,2,3})=1\omega_{\mathcal{G}}(\{1\},\{1,2,3\})=-1. It is easy to verify that the balancing condition is satisfied, so that we indeed get a tropical fan. Besides being the simplest non-Bergman Flag fan, it is nice to keep in mind because a very similar fan will appear naturally in Example 5.1.

  3. (3)

    A tropical fibre product from Example 2.22 (3) is a Flag fan. The Hasse diagram of its poset 𝒢\mathcal{G} is shown on Figure 5, and all the weights are 11 except for ω𝒢({5,6},E1E2)=1\omega_{\mathcal{G}}(\{5,6\},E_{1}\sqcup E_{2})=-1.

    Refer to caption
    Figure 5. The only covering edge with negative weight is in red.
Lemma 3.7.

If XX is a Flag fan and φ=φ𝒜\varphi=\varphi_{\mathcal{A}} a cut function on XX, then Y=φ𝒜XY=\varphi_{\mathcal{A}}\cdot X is a Flag fan.

Proof.

The idea of proof is straightforward: we take Definition 2.14 and verify the weights on the hyperfaces of maximal cones of XX. Maximal cone σ𝐅\sigma_{\mathbf{F}} corresponds to the maximal flag 𝐅\mathbf{F} of 𝒢\mathcal{G} consisting of =F0F1Fr=E\varnothing=F_{0}\lessdot F_{1}\lessdot\ldots\lessdot F_{r}=E with Fi𝒢F_{i}\in\mathcal{G}, and each of its hyperfaces corresponds to 𝐅\mathbf{F} with one missing set. Since φ\varphi is a monotonously non-increasing function, and sets FiF_{i} are pairwise comparable, there exists unique ii such that φ(Fi)=0\varphi(F_{i})=0 and φ(Fi+1)=1\varphi(F_{i+1})=-1. Consider τ=τ𝐅\tau=\tau_{\mathbf{F}^{\prime}} where 𝐅=𝐅{Fj}\mathbf{F}^{\prime}=\mathbf{F}\setminus\{F_{j}\}. Recall that the unique generator of the maximal cone σ\sigma not belonging to the face τ\tau can be chosen as vσ/τv_{\sigma/\tau}. Denote by 𝐅FF\mathbf{F}_{F\to F^{\prime}} the flag 𝐅\mathbf{F} with FF replaced by FF^{\prime}. We get

ωY(τ)=στωX(σ)φ(vσ/τ)φ(στωX(σ)vσ/τ)==Fj+1FFj1ωX(σ𝐅FjF)φ(eF)φ(Fj+1FFj1ωX(σ𝐅FjF)eF)==Fj+1FFj1kj1,jw𝒢(Fk,Fk+1)w𝒢(Fj1,F)w𝒢(F,Fj+1)φ(eF)φ(Fj+1FFj1kj1,jw𝒢(Fk,Fk+1)w𝒢(Fj1,F)w𝒢(F,Fj+1)eF)==kj1,jw𝒢(Fk,Fk+1)(Fj+1FFj1w𝒢(Fj1,F)w𝒢(F,Fj+1)φ(eF)φ(Fj+1FFj1w𝒢(Fj1,F)w𝒢(F,Fj+1)eF))==kj1,jw𝒢(Fk,Fk+1)deg(φ|XFj1,Fj+1XFj1,Fj+1),\begin{split}\omega_{Y}(\tau)=\sum_{\sigma\supset\tau}\omega_{X}(\sigma)\varphi(v_{\sigma/\tau})-\varphi\left(\sum_{\sigma\supset\tau}\omega_{X}(\sigma)v_{\sigma/\tau}\right)=\\ =\sum_{F_{j+1}\gtrdot F^{\prime}\gtrdot F_{j-1}}\omega_{X}(\sigma_{\mathbf{F}_{F_{j}\to F^{\prime}}})\varphi(e_{F^{\prime}})-\varphi\left(\sum_{F_{j+1}\gtrdot F^{\prime}\gtrdot F_{j-1}}\omega_{X}(\sigma_{\mathbf{F}_{F_{j}\to F^{\prime}}})e_{F^{\prime}}\right)=\\ =\sum_{F_{j+1}\gtrdot F^{\prime}\gtrdot F_{j-1}}\prod_{k\neq j-1,j}w_{\mathcal{G}}(F_{k},F_{k+1})\cdot w_{\mathcal{G}}(F_{j-1},F^{\prime})\cdot w_{\mathcal{G}}(F^{\prime},F_{j+1})\cdot\varphi(e_{F^{\prime}})-\\ -\varphi\left(\sum_{F_{j+1}\gtrdot F^{\prime}\gtrdot F_{j-1}}\prod_{k\neq j-1,j}w_{\mathcal{G}}(F_{k},F_{k+1})\cdot w_{\mathcal{G}}(F_{j-1},F^{\prime})\cdot w_{\mathcal{G}}(F^{\prime},F_{j+1})\cdot e_{F^{\prime}}\right)=\\ =\prod_{k\neq j-1,j}w_{\mathcal{G}}(F_{k},F_{k+1})\left(\sum_{F_{j+1}\gtrdot F^{\prime}\gtrdot F_{j-1}}w_{\mathcal{G}}(F_{j-1},F^{\prime})\cdot w_{\mathcal{G}}(F^{\prime},F_{j+1})\cdot\varphi(e_{F^{\prime}})\right.-\\ -\varphi\left.\left(\sum_{F_{j+1}\gtrdot F^{\prime}\gtrdot F_{j-1}}w_{\mathcal{G}}(F_{j-1},F^{\prime})\cdot w_{\mathcal{G}}(F^{\prime},F_{j+1})\cdot e_{F^{\prime}}\right)\right)=\\ =\prod_{k\neq j-1,j}w_{\mathcal{G}}(F_{k},F_{k+1})\cdot\deg{(\varphi|_{X_{F_{j-1},F_{j+1}}}\cdot X_{F_{j-1},F_{j+1}})},\end{split}

where XFj1,Fj+1X_{F_{j-1},F_{j+1}} in Fj+1Fj1\mathbb{R}^{F_{j+1}\setminus F_{j-1}} is a Flag fan of the segment [Fj1,Fj+1]𝒢[F_{j-1},F_{j+1}]\subset\mathcal{G} with induced edge weights. Thus, the weight ωY(τ)\omega_{Y}(\tau) can be calculated locally — besides weights of covers of σ\sigma, it depends only on the edge weights of [Fj1,Fj+1][F_{j-1},F_{j+1}] and values of φ\varphi on that segment of 𝒢\mathcal{G}.

The first thing we notice is that if a hyperface τ=τ𝐅\tau=\tau_{\mathbf{F}^{\prime}} contains both FiF_{i} and Fi+1F_{i+1} — that is, the missing level is neither ii-th nor (i+1)(i+1)-st (say, it is jj-th), then ωY(τ)=0\omega_{Y}(\tau)=0. Indeed, in this case φ|XFj1,Fj+1\varphi|_{X_{F_{j-1},F_{j+1}}} is constant 0 or constant 1-1, therefore linear, not just piecewise linear, so its Weil divisor is zero.

Note that the modified pair (𝒢φ,ω𝒢φ)(\mathcal{G}_{\varphi},\omega_{\mathcal{G}_{\varphi}}), where 𝒢φ\mathcal{G}_{\varphi} has the same covering relations \gtrdot with the same weights as 𝒢\mathcal{G}, except those with different values of φ\varphi, and additional covers between Fj1𝒢φFj+1F_{j-1}\lessdot_{\mathcal{G}_{\varphi}}F_{j+1} if deg(φ|XFj1,Fj+1XFj1,Fj+1)0\deg{(\varphi|_{X_{F_{j-1},F_{j+1}}}\cdot X_{F_{j-1},F_{j+1}})}\neq 0 with those non-zero weights, is also ranked: rk𝒢φ(F)=rk𝒢(F)+φ(F)\operatorname{rk}_{\mathcal{G}_{\varphi}}(F)=\operatorname{rk}_{\mathcal{G}}(F)+\varphi(F), and YY is the Flag fan of this new pair. Indeed, every maximal cone of YY with non-zero weight has generators with all possible ranks in 𝒢φ\mathcal{G}_{\varphi} — from 11 to r1r-1. ∎

Corollary 3.8 (Algorithmic description).

To construct pair (𝒢φ,ω𝒢φ)(\mathcal{G}_{\varphi},\omega_{\mathcal{G}_{\varphi}}) from pair (𝒢,ω𝒢)(\mathcal{G},\omega_{\mathcal{G}}), perform the following steps:

  1. (1)

    Delete all covers F𝒢FF\lessdot_{\mathcal{G}}F^{\prime} if φ(F)φ(F)\varphi(F)\neq\varphi(F^{\prime});

  2. (2)

    For all segments [Fj1,Fj+1][F_{j-1},F_{j+1}] such that φ(Fj1)φ(Fj+1)\varphi(F_{j-1})\neq\varphi(F_{j+1}), add covers Fj1𝒢φFj+1F_{j-1}\lessdot_{\mathcal{G}_{\varphi}}F_{j+1} with ω𝒢φ(Fj1,Fj+1)=deg(φ|XFj1,Fj+1XFj1,Fj+1)\omega_{\mathcal{G}_{\varphi}}(F_{j-1},F_{j+1})=\deg{(\varphi|_{X_{F_{j-1},F_{j+1}}}\cdot X_{F_{j-1},F_{j+1}})} if it is not 0;

  3. (3)

    Delete those vertices which do not lie in the resulting graph of covers in the same connected component as \varnothing and EE (if \varnothing and EE are in different connected components, φX\varphi\cdot X is zero).

Corollary 3.9.

Tropical fibre product from Definition 2.21 is a Flag fan.

Proof.

Functions π(φi)\pi^{*}(\varphi_{i}) are pullbacks of cut functions on B(N)×B(N)B(N)\times B(N) with respect to the map π\pi which preserves inclusion of the sets, therefore, they are also cut functions. Thus, the claim follows from Lemma 3.7 applied multiple times. ∎

Lemma 3.10.

If XX is a Flag fan corresponding to the pair (𝒢,ω𝒢)(\mathcal{G},\omega_{\mathcal{G}}), and YY is a Flag fan corresponding to the pair (,ω)(\mathcal{H},\omega_{\mathcal{H}}), then X×YX\times Y is a Flag fan corresponding to the pair (𝒢×,ω𝒢×)(\mathcal{G}\times\mathcal{H},\omega_{\mathcal{G}\times\mathcal{H}}), where

ω𝒢×((X1,Y),(X2,Y))=ω𝒢(X1,X2),ω𝒢×((X,Y1),(X,Y2))=ω(Y1,Y2).\omega_{\mathcal{G}\times\mathcal{H}}((X_{1},Y),(X_{2},Y))=\omega_{\mathcal{G}}(X_{1},X_{2}),\;\omega_{\mathcal{G}\times\mathcal{H}}((X,Y_{1}),(X,Y_{2}))=\omega_{\mathcal{H}}(Y_{1},Y_{2}).
Proof.

The cone corresponding to the pair of maximal flags GiG_{i} in 𝒢\mathcal{G} and HiH_{i} \mathcal{H} is the union of all the cones corresponding to the flags in 𝒢×\mathcal{G}\times\mathcal{H} of the following form:

(G0,H0)(Gi1,Hj1)(Gik,Hjk),(G_{0},H_{0})\subsetneq(G_{i_{1}},H_{j_{1}})\subsetneq\ldots\subsetneq(G_{i_{k}},H_{j_{k}}),

where k=rk𝒢+rkk=\operatorname{rk}{\mathcal{G}}+\operatorname{rk}{\mathcal{H}}, and (is,js)=(is1,js1)+(1,0)(i_{s},j_{s})=(i_{s-1},j_{s-1})+(1,0) or (is,js)=(is1,js1)+(0,1)(i_{s},j_{s})=(i_{s-1},j_{s-1})+(0,1). Thus the supports coincide, and the weight of each such maximal cone is equal to the product of the respective weights of the maximal cones in the factors, since each covering edge of both 𝒢\mathcal{G} and \mathcal{H} is used in this flag exactly once. ∎

4. Tropical fibre product

Theorem 4.1.

Let M1M_{1} and M2M_{2} be matroids on groundsets E1E_{1} and E2E_{2}, respectively. Let NN be their common restriction, i.e., a matroid on groundset T=E1E2T=E_{1}\cap E_{2} such that M1|T=N=M2|TM_{1}|_{T}=N=M_{2}|_{T}. Let π1:B(M1)B(N)\pi_{1}\colon B(M_{1})\to B(N) and π2:B(M2)B(N)\pi_{2}\colon B(M_{2})\to B(N) be projections induced by the inclusions TE1T\hookrightarrow E_{1} and TE2T\hookrightarrow E_{2}. Let X=B(M1)×π1,B(N),π2B(M2)X=B(M_{1})\times_{\pi_{1},B(N),\pi_{2}}B(M_{2}) be the tropical fibre product of B(M1)B(M_{1}) and B(M2)B(M_{2}) over B(N)B(N), as in Definition 2.21. Then:

  1. (1)

    XX is a tropical fan of degree 1;

  2. (2)

    If there exists a proper amalgam MM of M1M_{1} and M2M_{2} over NN, as in Definition 2.3, then B(M)B(M1)×B(M2)B(M)\to B(M_{1})\times B(M_{2}) induced by E1E2E1E2E_{1}\sqcup E_{2}\to E_{1}\cup E_{2} is an isomorphism on XX;

  3. (3)

    If all weights of XX are positive, then the proper amalgam MM of M1M_{1} and M2M_{2} over NN exists.

Proof.

We begin with Claim 1. Let r1=rkM1r_{1}=\operatorname{rk}M_{1}, r2=rkM2r_{2}=\operatorname{rk}M_{2}, r0=rkNr_{0}=\operatorname{rk}N. By construction, XE1E2X\subset\mathbb{R}^{E_{1}\sqcup E_{2}} is a tropical fan of dimension r=r1+r2r0r=r_{1}+r_{2}-r_{0}, therefore, to verify that it has degree 1 we need to prove that XH=1eE1E2X\cdot H=1\cdot\langle e_{E_{1}\sqcup E_{2}}\rangle, where H=αr1B(M1M2)H=\alpha^{r-1}\cdot B(M_{1}\oplus M_{2}). By the definition of tropical fibre product, we have X=π(φ1)π(φr0)B(M1M2)X=\pi^{*}(\varphi_{1})\cdots\pi^{*}(\varphi_{r_{0}})\cdot B(M_{1}\oplus M_{2}). Therefore, by commutativity of cutting out Weil divisors ([FR12], Theorem 4.5 (6)), we need π(φ1)π(φr0)H=1eE1E2\pi^{*}(\varphi_{1})\cdots\pi^{*}(\varphi_{r_{0}})\cdot H=1\cdot\langle e_{E_{1}\sqcup E_{2}}\rangle.

Observe that π(H)=αr01B(NN)\pi_{*}(H)=\alpha^{r_{0}-1}\cdot B(N\oplus N), since truncation commutes with restriction by Lemma 2.23. We also know that φ1,,φr0\varphi_{1},\ldots,\varphi_{r_{0}} cut out the diagonal ΔB(NN)\Delta\subset B(N\oplus N), which is a Bergman fan of a matroid and hence has degree 1, so φ1φr0π(H)=1eT1T2\varphi_{1}\cdots\varphi_{r_{0}}\cdot\pi_{*}(H)=1\cdot\langle e_{T_{1}\sqcup T_{2}}\rangle in B(NN)B(N\oplus N). The desired equality then follows from consecutively applying the projection formula 2.24. Indeed, the first time we use it, it yields π(π(φ1)H)=φ1π(H)\pi_{*}(\pi^{*}(\varphi_{1})\cdot H)=\varphi_{1}\cdot\pi_{*}(H). Next, π(π(φ2)π(φ1)H)=φ2π(π(φ1)H)=φ2φ1π(H)\pi_{*}(\pi^{*}(\varphi_{2})\cdot\pi^{*}(\varphi_{1})\cdot H)=\varphi_{2}\cdot\pi_{*}(\pi^{*}(\varphi_{1})\cdot H)=\varphi_{2}\cdot\varphi_{1}\cdot\pi_{*}(H), and so forth, until we get π(π(φ1)π(φr0)H)=φ1φr0π(H)=1eT1T2\pi_{*}(\pi^{*}(\varphi_{1})\cdots\pi^{*}(\varphi_{r_{0}})\cdot H)=\varphi_{1}\cdots\varphi_{r_{0}}\cdot\pi_{*}(H)=1\cdot\langle e_{T_{1}\sqcup T_{2}}\rangle, and the claim follows from the fact that eT1T2\langle e_{T_{1}\sqcup T_{2}}\rangle is the only 11-dimensional cone an affine fan can have, and that π\pi_{*} sums weights over the preimages of the cone.

We proceed to verify Claim 2. We will prove that XX coincides with the Bergman fan of the matroid MM^{\prime} on groundset E1E2E_{1}\sqcup E_{2} which is obtained from the proper amalgam MM on groundset E1E2E_{1}\cup E_{2} by replacing each element of TT with two parallel copies. It is immediate from the definition of the Bergman fan that B(M)B(M^{\prime}) and B(M)B(M) are isomorphic with isomorphism induced by E1E2E1E2E_{1}\sqcup E_{2}\to E_{1}\cup E_{2}. Denote the copies of TT in E1E2E_{1}\sqcup E_{2} by T1T_{1} and T2T_{2}.

Observe that B(M)B(M1M2)B(M^{\prime})\subset B(M_{1}\oplus M_{2}) — since each flat of MM^{\prime} restricts to a flat of both M1M_{1} and M2M_{2} (as M1M_{1} and M2M_{2} are restrictions of MM^{\prime}) and since flags of flats of MM^{\prime} are also flags of flats in M1M2M_{1}\oplus M_{2}. Therefore, by equation 3.1 from Lemma 3.1, B(M)B(M^{\prime}) is cut by functions ψ1,,ψr0\psi_{1},\ldots,\psi_{r_{0}} given by

ψi(eF)={0,if rkM(F)+r0irkM1M2(F)1,otherwise.\psi_{i}(e_{F})=\begin{cases}0,&\mbox{if $\operatorname{rk}_{M^{\prime}}(F)+r_{0}-i\geqslant\operatorname{rk}_{M_{1}\oplus M_{2}}(F)$}\\ -1,&\mbox{otherwise.}\end{cases}

We are going to show that ψi\psi_{i} coincides with π(φi)\pi^{*}(\varphi_{i}) on all the rays of ψi1ψ1B(M1M2)\psi_{i-1}\cdots\psi_{1}\cdot B(M_{1}\oplus M_{2}), a Bergman fan of the matroid that we denote by B(M~i1)B(\tilde{M}_{i-1}). Therefore, iterated tropical modifications with respect to functions ψi\psi_{i}, which yield B(M)B(M^{\prime}), and with respect to functions π(φi)\pi^{*}(\varphi_{i}), which yield XX, must also coincide.

Consider FE1E2F\subset E_{1}\sqcup E_{2} such that eFψi1ψ1B(M1M2)e_{F}\subset\psi_{i-1}\cdots\psi_{1}\cdot B(M_{1}\oplus M_{2}), and denote

X1=FE1;X2=FE2;Y1=FT1;Y2=FT2.X_{1}=F\cap E_{1};\;X_{2}=F\cap E_{2};\;Y_{1}=F\cap T_{1};\;Y_{2}=F\cap T_{2}.

Then, rkM1M2(F)=rkM1(X1)+rkM2(X2)\operatorname{rk}_{M_{1}\oplus M_{2}}(F)=\operatorname{rk}_{M_{1}}(X_{1})+\operatorname{rk}_{M_{2}}(X_{2}), and

rkM(F)=min{η(Y):YF}η(F)=rkM1(X1Y2)+rkM2(X2Y1)rkN(Y1Y2),\operatorname{rk}_{M^{\prime}}(F)=\min{\{\eta(Y)\colon Y\supseteq F\}}\leqslant\eta(F)=\operatorname{rk}_{M_{1}}(X_{1}\cup Y_{2})+\operatorname{rk}_{M_{2}}(X_{2}\cup Y_{1})-\operatorname{rk}_{N}(Y_{1}\cup Y_{2}),

with equality attained if FF is a flat of MM^{\prime}. Similarly, by Definition 3.3,

π(φi)(eF)={0,if rkN(Y1Y2)+r0irkN(Y1)+rkN(Y2)1,otherwise.\pi^{*}(\varphi_{i})(e_{F})=\begin{cases}0,&\mbox{if $\operatorname{rk}_{N}(Y_{1}\cup Y_{2})+r_{0}-i\geqslant\operatorname{rk}_{N}(Y_{1})+\operatorname{rk}_{N}(Y_{2})$}\\ -1,&\mbox{otherwise.}\end{cases}

Therefore, to show that π(φi)(eF)=ψi(eF)\pi^{*}(\varphi_{i})(e_{F})=\psi_{i}(e_{F}) for all 0ir00\leqslant i\leqslant r_{0}, we need to verify that

rkM1(X1Y2)+rkM2(X2Y1)rkN(Y1Y2)rkM1(X1)rkM2(X2)==rkN(Y1Y2)rkN(Y1)rkN(Y2),\begin{split}\operatorname{rk}_{M_{1}}(X_{1}\cup Y_{2})+\operatorname{rk}_{M_{2}}(X_{2}\cup Y_{1})-\operatorname{rk}_{N}(Y_{1}\cup Y_{2})-\operatorname{rk}_{M_{1}}(X_{1})-\operatorname{rk}_{M_{2}}(X_{2})=\\ =\operatorname{rk}_{N}(Y_{1}\cup Y_{2})-\operatorname{rk}_{N}(Y_{1})-\operatorname{rk}_{N}(Y_{2}),\end{split} (4.1)

and that η(F)=rkM(F)\eta(F)=\operatorname{rk}_{M^{\prime}}(F), since both π(φj)\pi^{*}(\varphi_{j}) and ψj\psi_{j} can only decrease as ii grows so φi=ψi\varphi_{i}=\psi_{i} if iφi=iψi\sum_{i}{\varphi_{i}}=\sum_{i}{\psi_{i}}. Neither is true for arbitrary FF in the flat lattice of the initial matroid M1M2M_{1}\oplus M_{2}, but both are true “while it matters”. More precisely, we will show that if one of these equalities fails to hold for ii-th function, then eFe_{F} is not a ray of B(M~i1)B(\tilde{M}_{i-1}).

After rearranging equation 4.1 splits into two analogous rank identities:

rkM1(X1Y2)rkM1(X1)=rkN(Y1Y2)rkN(Y1);rkM2(X2Y1)rkM1(X2)=rkN(Y1Y2)rkN(Y2).\begin{split}\operatorname{rk}_{M_{1}}(X_{1}\cup Y_{2})-\operatorname{rk}_{M_{1}}(X_{1})=\operatorname{rk}_{N}(Y_{1}\cup Y_{2})-\operatorname{rk}_{N}(Y_{1});\\ \operatorname{rk}_{M_{2}}(X_{2}\cup Y_{1})-\operatorname{rk}_{M_{1}}(X_{2})=\operatorname{rk}_{N}(Y_{1}\cup Y_{2})-\operatorname{rk}_{N}(Y_{2}).\end{split} (4.2)

First, observe that the inequality \leqslant holds in both identities of equation 4.2 by Lemma 2.2 and due to the fact that NN is the restriction of M1M_{1} and M2M_{2}, so the ranks of the subsets of TT are the same in NN as in M1M_{1} or M2M_{2}. Moreover, if either one of these inequalities turns out to be strict, or if η(F)<rkM(F)\eta(F)<\operatorname{rk}_{M^{\prime}}(F), then FF is not a flat of MM^{\prime} (if it is, then Y1=Y2Y_{1}=Y_{2}, and equation 4.2 holds trivially).

Take the earliest step ii which creates the discrepancy between ψ\psi and π(φ)\pi^{*}(\varphi). Namely, assume that for each j<ij<i and each ray eFe_{F} of B(M~j1)B(\tilde{M}_{j-1}) the equality ψj(F)=π(φj)(F)\psi_{j}(F)=\pi^{*}(\varphi_{j})(F) holds. Now, both sequences of numbers ψj(F)\psi_{j}(F) and π(φj)(F)\pi^{*}(\varphi_{j})(F) are monotonously non-increasing as jj grows, and

jψj(F)=rkM(F)rkM1M2(F)<rkN(Y1Y2)rkN(Y1)rkN(Y2)=jπ(φj)(F),\sum_{j}\psi_{j}(F)=\operatorname{rk}_{M^{\prime}}(F)-\operatorname{rk}_{M_{1}\oplus M_{2}}(F)<\operatorname{rk}_{N}(Y_{1}\cup Y_{2})-\operatorname{rk}_{N}(Y_{1})-\operatorname{rk}_{N}(Y_{2})=\sum_{j}\pi^{*}(\varphi_{j})(F),

which means that 1=ψi(F)<π(φi)(F)=0-1=\psi_{i}(F)<\pi^{*}(\varphi_{i})(F)=0. Our tactics is now to show that FF cannot actually be a flat of M~i1\tilde{M}_{i-1}, since there must exist a covering flat FF^{\prime} of FF in M~i2\tilde{M}_{i-2} which belongs to the modular cut which yields M~i1\tilde{M}_{i-1}, and, therefore, eFe_{F} does not belong to the support of B(M~i1)B(\tilde{M}_{i-1}) by Lemma 3.1.

Since FF is not a flat of MM^{\prime}, we can define F′′=clMFFF^{\prime\prime}=\operatorname{cl}_{M^{\prime}}{F}\supsetneq F. Now, both FF and F′′F^{\prime\prime} are flats of M~i2\tilde{M}_{i-2} — the former by assumption, the latter because it is even the flat of MM^{\prime} which is at the end of the process deleting some of the flats. Take any FFF^{\prime}\gtrdot F in (M~i2)\mathcal{F}(\tilde{M}_{i-2}) such that F<F′′F^{\prime}<F^{\prime\prime}. Since both X1X_{1} and X2X_{2} are flats of M1M_{1} and M2M_{2} respectively, we have rkM1M2F>rkM1M2F\operatorname{rk}_{M_{1}\oplus M_{2}}F^{\prime}>\operatorname{rk}_{M_{1}\oplus M_{2}}F. Therefore, for FF^{\prime} the difference between the initial rank rkM1M2(F)\operatorname{rk}_{M_{1}\oplus M_{2}}(F^{\prime}) and the final rank rkM(F)\operatorname{rk}_{M^{\prime}}(F^{\prime}) is strictly larger than for FF, so jψj(F)<jψj(F)\sum_{j}\psi_{j}(F^{\prime})<\sum_{j}\psi_{j}(F).

Using monotonicity of ψ\psi again, we get ψi1(F)=1\psi_{i-1}(F^{\prime})=-1 (i.e., ψ\psi already was 1-1 on FF^{\prime} on the previous step), but, at the same time, ψi1(F)=π(φi1)(F)\psi_{i-1}(F)=\pi^{*}(\varphi_{i-1})(F) since ii-th step is the first with the discrepancy, which in its turn, by monotonicity of πφ\pi^{*}{\varphi}, is not less than π(φi)(F)=0\pi^{*}(\varphi_{i})(F)=0. For the sake of clarity these considerations are gathered in Figure 6. So, we have that FF is a flat under the (i1)(i-1)-st cut and cannot be a flat of M~i1\tilde{M}_{i-1} by Lemma 3.1. This completes the proof of Claim 2.

Refer to caption
Figure 6. Relations between values of π(φ)\pi^{*}(\varphi) and ψ\psi.

We will now verify Claim 3. By Claim 1 and Theorem 2.17 we have that X=B(M)X=B(M) for some matroid MM, and we want to show that MM is actually the proper amalgam of M1M_{1} and M2M_{2} along NN (with added parallel copies). The reason we cannot simply follow the equivalences constructed for Claim 2 in the opposite direction is that we have no analogues of Corollary 3.6 from [FR12]. More precisely, we do not know what happens to the fan ψiψ1B(M1M2)\psi_{i}\cdots\psi_{1}\cdot B(M_{1}\oplus M_{2}) once it ceases to be a Bergman fan for the first time. Theoretically, it can emerge effective again (which actually happens, as our verification of the degree practically means that after enough truncations the fan becomes 1eE1\cdot\langle e_{E}\rangle, with ranks of flats being not the ones prescribed by the definition of proper amalgam.

By Corollary 3.9, though, XX is a Flag fan corresponding to some pair (𝒢,ω𝒢)(\mathcal{G},\omega_{\mathcal{G}}). All the rays FF of XX have Y1=Y2=YY_{1}=Y_{2}=Y by Lemma 2.26, so

rk𝒢(F)=rkM1M2(F)+π(φi)(F)=rkM1(X1)+rkM2(X2)rkN(Y)=η(F).\operatorname{rk}_{\mathcal{G}}(F)=\operatorname{rk}_{M_{1}\oplus M_{2}}(F)+\sum\pi^{*}(\varphi_{i})(F)=\operatorname{rk}_{M_{1}}(X_{1})+\operatorname{rk}_{M_{2}}(X_{2})-\operatorname{rk}_{N}(Y)=\eta(F).

Summarizing, X=B(M)X=B(M) for some MM by Theorem 2.17, and for each flat FF of MM we have rkM(F)=η(F)\operatorname{rk}_{M}(F)=\eta(F), which means that, by Definition 2.3, MM is a proper amalgam of M1M_{1} and M2M_{2} along NN. ∎

5. Tropical graph correspondences

In this section we develop first steps in tropical correspondence theory. The motivation is to obtain a category of tropical fans where the tropical fibre product is actually a pullback. While this precise formulation is not achieved (and is likely not possible to achieve, see Remark 5.17), worthwhile statements are established in the process.

5.1. Introducing notions

We begin with the following unsatisfactory observation. Consider the tropical fibre product from Example 2.22 (2) and a fan B(M)B(M^{\prime}), where MM^{\prime} is another amalgam of M1M_{1} and M2M_{2} with {3}\{3\} and {4}\{4\} being parallel elements. Then, neither of the supports of B(M)=B(M1)×B(N)B(M2)B(M)=B(M_{1})\times_{B(N)}B(M_{2}) and B(M)B(M^{\prime}) is contained in the other, and it is not difficult to check (see Figure 7 for projective Bergman fans of MM and MM^{\prime}) that there are no tropical morphisms between them making the diagram commute:

B(M)\textstyle{B(M^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q1\scriptstyle{q_{1}}q2\scriptstyle{q_{2}}\scriptstyle{\nexists}B(M)\textstyle{B(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p1\scriptstyle{p_{1}}p2\scriptstyle{p_{2}}B(M2)\textstyle{B(M_{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π2\scriptstyle{\pi_{2}}B(M1)\textstyle{B(M_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π1\scriptstyle{\pi_{1}}B(N)\textstyle{B(N)}
Refer to caption
Figure 7. B(M)B(M^{\prime}) is in blue, B(M)B(M) is in red.

Something can be done, though:

Example 5.1.

Denote the groundset of MM^{\prime} by E={1,2,3,4}E^{\prime}=\{1^{\prime},2^{\prime},3^{\prime},4^{\prime}\}, the groundset of MM by E={1,2,3,4}E=\{1,2,3,4\}, and consider a Flag fan ΓB(M)×B(M)\Gamma\subset B(M^{\prime})\times B(M), where

𝒢={,EE,{1,1},{2,2},{3,4,3},{3,4,4},{3,4}},\mathcal{G}=\{\varnothing,E\cup E^{\prime},\{1^{\prime},1\},\{2^{\prime},2\},\{3^{\prime},4^{\prime},3\},\{3^{\prime},4^{\prime},4\},\{3^{\prime},4^{\prime}\}\},

and all edge weights are 1 except for ({3,4},EE)(\{3^{\prime},4^{\prime}\},E\cup E^{\prime}) which is 1-1. It is easy to see that Γ\Gamma is actually a tropical fan and a subfan in B(M)×B(M)B(M^{\prime})\times B(M).

The push-forward of Γ\Gamma onto B(M)×B(M1)B(M^{\prime})\times B(M_{1}) (along id×p1\operatorname{id}\times p_{1}) is an actual graph of tropical morphism q1:B(M)B(M1)q_{1}\colon B(M^{\prime})\to B(M_{1}) (as {3,4,4},{3,4}\{3^{\prime},4^{\prime},4\},\{3^{\prime},4^{\prime}\} both map to {3,4}\{3^{\prime},4^{\prime}\} and cancel each other out, the latter having weight 1-1). The same goes for q2:B(M)B(M2)q_{2}\colon B(M^{\prime})\to B(M_{2}).

We proceed to build on this example.

Definition 5.2.

Given two simple matroids M1,M2M_{1},M_{2} and their lattices of flats 1=(M1),2=(M2)\mathcal{F}_{1}=\mathcal{F}(M_{1}),\mathcal{F}_{2}=\mathcal{F}(M_{2}), the lattice morphism is a map f:12f\colon\mathcal{F}_{1}\to\mathcal{F}_{2} such that if F>1FF>_{\mathcal{F}_{1}}F^{\prime}, then f(F)2f(F)f(F)\geqslant_{\mathcal{F}_{2}}f(F^{\prime}).

We call a lattice morphism ff a weak lattice map if f(F)f(F)\neq\varnothing for FF\neq\varnothing and rkM1FrkM2(f(F))\operatorname{rk}_{M_{1}}F\geqslant\operatorname{rk}_{M_{2}}(f(F)).

We call a weak lattice map covering if whenever F1FF\gtrdot_{\mathcal{F}_{1}}F^{\prime}, either f(F)2f(F)f(F)\gtrdot_{\mathcal{F}_{2}}f(F^{\prime}) or f(F)=f(F)f(F)=f(F^{\prime}).

Note that covering lattice maps are automatically weak provided that f(F)f(F)\neq\varnothing for FF\neq\varnothing.

Simple matroids with weak lattice maps form a category 𝐒𝐌𝐚𝐭𝐫𝐖𝐋\mathbf{SMatrWL}, and covering maps form a wide subcategory 𝐒𝐌𝐚𝐭𝐫𝐖𝐋\mathbf{SMatrWL_{\gtrdot}}. The category of simple matroids with weak maps of matroids (Definition 2.6) is denoted by 𝐒𝐌𝐚𝐭𝐫𝐖\mathbf{SMatrW}. There is a forgetful functor Pt:𝐒𝐌𝐚𝐭𝐫𝐖𝐋𝐒𝐌𝐚𝐭𝐫𝐖\mathrm{Pt}\colon\mathbf{SMatrWL}\to\mathbf{SMatrW}: it maps simple matroids to themselves, and, given f:12f\colon\mathcal{F}_{1}\to\mathcal{F}_{2}, the map Pt(f):E1E2\mathrm{Pt}(f)\colon E_{1}\to E_{2} is defined via Pt(f)(x)=f({x})\mathrm{Pt}(f)(x)=f(\{x\}). The definition is correct since rkM2f({x})=1\operatorname{rk}_{M_{2}}f(\{x\})=1, and all rank-1 flats of M2M_{2} are one-element subsets since M2M_{2} is simple.

The map Pt(f):E1E2\mathrm{Pt}(f)\colon E_{1}\to E_{2} is actually a weak map: if XE1X\subset E_{1}, then

rkM2Pt(f)(X)=rkM2(clM2xXf({x}))=rkM2xXf({x})rkM2f(xX{x})rkM1X.\operatorname{rk}_{M_{2}}\mathrm{Pt}(f)(X)=\operatorname{rk}_{M_{2}}{\left(\operatorname{cl}_{M_{2}}{\bigcup_{x\in X}f(\{x\})}\right)}=\operatorname{rk}_{M_{2}}{\bigvee_{x\in X}f(\{x\})}\leqslant\operatorname{rk}_{M_{2}}{f\left(\bigvee_{x\in X}\{x\}\right)}\leqslant\operatorname{rk}_{M_{1}}X.

In the examples of this section matroids shown are sometimes not simple for the sake of clarity, but what is assumed every time is their simplifications, where each flat of rank 11 is replaced by a single element.

Not only functor Pt\mathrm{Pt} is not faithful, it also does not have the right inverse, as shown by the following

Example 5.3.

On Figure 8 there are three different weak lattice maps from the matroid on the left side to the matroid on the right side. The only difference between them is the image of the flat {3,4}\{3,4\} — in the top map it is {1,3,4}\{1,3,4\}, in the bottom map it is {2,3,4}\{2,3,4\}, and in the central map it is {3,4}\{3,4\} again. Note that the central map is not a covering lattice map ({1,2,3,4}\{1,2,3,4\} covers {3,4}\{3,4\} in the left matroid, but not in the right one), while the top and the bottom are compositions of covering lattice maps and, therefore, covering lattice maps themselves. For each of the four top and bottom arrows there is only one lattice map corresponding to the identity map on the groundset. Thus, no matter which lattice map we consider to be the image of the central identity map on the groundset, the map will not be functorial.

Refer to caption
Figure 8. Three weak lattice maps with the same weak map of groundset.

Next, we define tropical correspondences between Bergman fans. The whole construction is almost verbatim to the correspondences between, say, smooth projective varieties over \mathbb{C}, with the (already noted in tropical intersection theory) exception that those are defined on classes of cycles modulo rational equivalence, while groups of tropical cycles are not quotients.

Definition 5.4.

A tropical correspondence Γ\Gamma between Bergman fans BX=B(M1)B_{X}=B(M_{1}) and BY=B(M2)B_{Y}=B(M_{2}) denoted by Γ:BXBY\Gamma\colon B_{X}\vdash B_{Y} is a tropical subcycle in BX×BYB_{X}\times B_{Y}. The identity correspondence is the diagonal ΔXBX×BX\Delta_{X}\subset B_{X}\times B_{X}. The composition of correspondences ΓXY:BXBY\Gamma_{XY}\colon B_{X}\vdash B_{Y} and ΓYZ:BYBZ\Gamma_{YZ}\colon B_{Y}\vdash B_{Z} is given by the following formula:

ΓYZΓXY:=(πXZ)(πXY(ΓXY)πYZ(ΓYZ))BX×BZ,\Gamma_{YZ}\circ\Gamma_{XY}\vcentcolon=(\pi_{XZ})_{*}(\pi_{XY}^{*}(\Gamma_{XY})\cdot\pi_{YZ}^{*}(\Gamma_{YZ}))\subset B_{X}\times B_{Z},

where πXY:BX×BY×BZBX×BY\pi_{XY}\colon B_{X}\times B_{Y}\times B_{Z}\to B_{X}\times B_{Y} is a projection, and similarly πXZ,πYZ\pi_{XZ},\pi_{YZ}.

Lemma 5.5.

Bergman fans of simple matroids with tropical correspondences form a category (that we denote with 𝐒𝐌𝐚𝐭𝐫𝐂𝐨𝐫\mathbf{SMatrCor}):

  • ΔX:BXBX\Delta_{X}\colon B_{X}\vdash B_{X} is a unit: for any Γ:BXBYΓΔX=Γ\Gamma\colon B_{X}\vdash B_{Y}\;\Gamma\circ\Delta_{X}=\Gamma, and for any Γ:BYBXΔXΓ=Γ\Gamma\colon B_{Y}\vdash B_{X}\;\Delta_{X}\circ\Gamma=\Gamma;

  • Composition is associative: for any Γ12:X1X2,Γ23:X2X3,Γ34:X3X4\Gamma_{12}\colon X_{1}\vdash X_{2},\Gamma_{23}\colon X_{2}\vdash X_{3},\Gamma_{34}\colon X_{3}\vdash X_{4}, we have

    Γ34(Γ23Γ12)=(Γ34Γ23)Γ12.\Gamma_{34}\circ(\Gamma_{23}\circ\Gamma_{12})=(\Gamma_{34}\circ\Gamma_{23})\circ\Gamma_{12}.
Proof.

This is standard: see, for example, Proposition 16.1.1 of [Ful98]. The only non-trivial ingredient needed is the projection formula — Theorem 2.24. Note that Proposition 1.7 of [Ful98], that is used as a justification of one of the equalities in the proof, is actually applied to show that

(π1231234)(π1241234)=(π12123)(π12124),(\pi^{1234}_{123})_{*}\circ(\pi^{1234}_{124})^{*}=(\pi^{123}_{12})^{*}\circ(\pi^{124}_{12})_{*},

where πj1jli1ik\pi^{i_{1}\ldots i_{k}}_{j_{1}\ldots j_{l}} denotes the projection Xi1××XikXj1××XjlX_{i_{1}}\times\ldots\times X_{i_{k}}\to X_{j_{1}}\times\ldots\times X_{j_{l}}. This particular case is obviously true in 𝐒𝐌𝐚𝐭𝐫𝐂𝐨𝐫\mathbf{SMatrCor}. ∎

Category 𝐒𝐌𝐚𝐭𝐫𝐂𝐨𝐫\mathbf{SMatrCor} has too many morphisms to be useful for us here. In the category of correspondences between smooth projective varieties there is a subcategory of graphs of morphisms of varieties. We want to imitate this subcategory, but our starting point (see Example 5.1 and discussion above it) is that actual graphs of tropical morphisms are not sufficient. Therefore, we establish a larger subcategory which we expect to behave like graphs.

Definition 5.6.

Given matroids M1,M2M_{1},M_{2} and a weak lattice map f:12f\colon\mathcal{F}_{1}\to\mathcal{F}_{2}, we define a graph rank function γ(f):(M1M2)0\gamma(f)\colon\mathcal{F}(M_{1}\oplus M_{2})\to\mathbb{Z}_{\leqslant 0} by the rule:

γ(f)((F1,F2))=rkM2(f(F1)F2)rkM1F1rkM2F2.\gamma(f)((F_{1},F_{2}))=\operatorname{rk}_{M_{2}}(f(F_{1})\cup F_{2})-\operatorname{rk}_{M_{1}}F_{1}-\operatorname{rk}_{M_{2}}F_{2}.

Similar to Lemma 3.1, a graph rank function γ(f)\gamma(f) is decomposed as a sum of γi(f):(M1M2){1,0},1irkM1\gamma_{i}(f)\colon\mathcal{F}(M_{1}\oplus M_{2})\to\{-1,0\},1\leqslant i\leqslant\operatorname{rk}{M_{1}} in such a way that value on each F=(F1,F2)F=(F_{1},F_{2}) does not increase as ii grows. More precisely,

γi(f)((F1,F2))={0, if rkM2(f(F1)F2)+rkM1irkM1F1+rkM2F21, otherwise.\gamma_{i}(f)((F_{1},F_{2}))=\begin{cases}0,\mbox{ if $\operatorname{rk}_{M_{2}}(f(F_{1})\cup F_{2})+\operatorname{rk}{M_{1}}-i\geqslant\operatorname{rk}_{M_{1}}F_{1}+\operatorname{rk}_{M_{2}}F_{2}$}\\ -1,\mbox{ otherwise.}\end{cases}

A graph correspondence Γf\Gamma_{f} is given by

i=1rkM1γi(f)(B(M1)×B(M2)).\prod_{i=1}^{\operatorname{rk}M_{1}}\gamma_{i}(f)\cdot(B(M_{1})\times B(M_{2})).
Example 5.7.

Here are a few instances of Definition 5.6:

  1. (1)

    For the identity lattice map ι:(N)(N)\iota\colon\mathcal{F}(N)\to\mathcal{F}(N) we get γi(ι)=φi\gamma_{i}(\iota)=\varphi_{i} from Definition 3.3, and thus Γι=ΔB(N)×B(N)\Gamma_{\iota}=\Delta\subset B(N)\times B(N).

  2. (2)

    More generally, if f:E1E2f\colon E_{1}\to E_{2} is a strong map between groundsets of matroids M1M_{1} and M2M_{2}, as in Definition 2.6, and f:E2E1f^{*}\colon\mathbb{R}^{E_{2}}\to\mathbb{R}^{E_{1}} is the dual map, then the set-theoretical graph

    {(v,f(v)),vB(M2)}B(M2)×B(M1)\{(v,f^{*}(v)),v\in B(M_{2})\}\subset B(M_{2})\times B(M_{1})

    is a Bergman fan B(Mf)B(M_{f}) of matroid MfM_{f} on the groundset E1E2E_{1}\sqcup E_{2}. Matroid MfM_{f} is isomorphic to M2M_{2}, with xE1x\in E_{1} being parallel element to f(x)E2f(x)\in E_{2}. By Lemma 3.1, since |B(Mf)||B(M1M2)||B(M_{f})|\subset|B(M_{1}\oplus M_{2})|, it is cut by the rank functions, and it is easy to see that they coincide with γi(f)\gamma_{i}(f_{\mathcal{F}}), where f:(M1)(M2)f_{\mathcal{F}}\colon\mathcal{F}(M_{1})\to\mathcal{F}(M_{2}) is the lattice map induced by ff. Thus, Γf\Gamma_{f_{\mathcal{F}}} equals set-theoretic graph of ff^{*}.

  3. (3)

    If MM and MM^{\prime} are as in Example 5.1, and f:(M)(M)f\colon\mathcal{F}(M)\to\mathcal{F}(M^{\prime}) is given by f(F)=clM(F)f(F)=\operatorname{cl}_{M^{\prime}}(F), then the Flag fan ΓB(M)×B(M)\Gamma\subset B(M^{\prime})\times B(M) that we constructed in the example equals Γf\Gamma_{f}.

We are going to see that Example 5.7 (3) is not a coincidence.

Lemma 5.8.

The graph correspondence Γf\Gamma_{f} of the covering lattice map ff is a Flag fan.

Proof.

We need to verify that each γi(f)\gamma_{i}(f) is monotonous on flats of M1M2M_{1}\oplus M_{2} and, therefore, on each of the Flag fans i=1jγi(f)(B(M1)×B(M2))\prod_{i=1}^{j}\gamma_{i}(f)\cdot(B(M_{1})\times B(M_{2})) on the way (their posets are the subposets of the initial lattice of flats). Then, the claim follows from Lemma 3.7.

Since γi(f)\gamma_{i}(f) are monotonously non-increasing as ii grows, and iγi(f)=γ(f)\sum_{i}\gamma_{i}(f)=\gamma(f), we want to show that, if FFF^{\prime}\gtrdot F, then γ(f)(F)γ(f)(F)\gamma(f)(F^{\prime})\leqslant\gamma(f)(F). The covering relations FM1M2FF^{\prime}\gtrdot_{M_{1}\oplus M_{2}}F can be of two kinds: F=(F1,F2)(F1,F2)=FF^{\prime}=(F^{\prime}_{1},F_{2})\gtrdot(F_{1},F_{2})=F, where F1M1F1F^{\prime}_{1}\gtrdot_{M_{1}}F_{1}, and F=(F1,F2)(F1,F2)=FF^{\prime}=(F_{1},F^{\prime}_{2})\gtrdot(F_{1},F_{2})=F, where F2M2F2F^{\prime}_{2}\gtrdot_{M_{2}}F_{2}. As γ(f)((F1,F2))=rkM2(f(F1)F2)rkM1F1rkM2F2\gamma(f)((F_{1},F_{2}))=\operatorname{rk}_{M_{2}}(f(F_{1})\cup F_{2})-\operatorname{rk}_{M_{1}}F_{1}-\operatorname{rk}_{M_{2}}F_{2}, and exactly one of the subtrahends increases by 11 when replacing FF with FF^{\prime}, we want to show that

rkM2(f(F1)F2)rkM2(f(F1)F2)+1 and rkM2(f(F1)F2)rkM2(f(F1)F2)+1.\operatorname{rk}_{M_{2}}(f(F^{\prime}_{1})\cup F_{2})\leqslant\operatorname{rk}_{M_{2}}(f(F_{1})\cup F_{2})+1\;\mbox{ and }\;\operatorname{rk}_{M_{2}}(f(F_{1})\cup F^{\prime}_{2})\leqslant\operatorname{rk}_{M_{2}}(f(F_{1})\cup F_{2})+1.

Since ff is a covering lattice map, rkM2f(F1)rkM2f(F1)+1\operatorname{rk}_{M_{2}}{f(F^{\prime}_{1})}\leqslant\operatorname{rk}_{M_{2}}{f(F_{1})}+1, so both inequalities boil down to the fact that in any matroid MM with groundset EE and for any AEA\subset E

FMFrkM(FA)rkM(FA)1,F^{\prime}\gtrdot_{M}F\Rightarrow\operatorname{rk}_{M}(F^{\prime}\cup A)-\operatorname{rk}_{M}(F\cup A)\leqslant 1,

which is simply Lemma 2.2. ∎

5.2. Graph functor

In this subsection we prove the second main result of the paper — that graph correspondence from Definition 5.6 provides a functor from 𝐒𝐌𝐚𝐭𝐫𝐖𝐋\mathbf{SMatrWL_{\gtrdot}} to 𝐒𝐌𝐚𝐭𝐫𝐂𝐨𝐫\mathbf{SMatrCor} (Theorem 5.13). In order to do that, we first describe the edge weights of graph correspondence Γf\Gamma_{f} combinatorially, in terms of ff, in Theorem 5.12, using that Γf\Gamma_{f} is a Flag fan due to Lemma 5.8. It requires some preparation. Recall the following

Definition 5.9.

A Möbius function μ\mu on poset 𝒢\mathcal{G} is defined for pairs A𝒢BA\leqslant_{\mathcal{G}}B recursively via

μ(A,A)=1,μ(A,B)=AC<Bμ(A,C).\mu(A,A)=1,\;\;\mu(A,B)=-\sum_{A\leqslant C<B}\mu(A,C).

One sees that function μ\mu is designed in such a way that ACBμ(A,C)=δ(A,B)\sum_{A\leqslant C\leqslant B}\mu(A,C)=\delta(A,B), where δ\delta is a delta-function, equal to 11 if A=BA=B and 0 otherwise.

We are going to use the following simple property of μ\mu, which can be found, for example, in [Sta86] (Ch. 3, Exercise 88).

Lemma 5.10.

For a poset 𝒫\mathcal{P} with the minimal element (denoted by \varnothing for convenience)

a,b𝒫μ(a,b)=1.\sum_{a,b\in\mathcal{P}}{\mu(a,b)}=1.
Proof.

We have

a,b𝒫μ(a,b)=b𝒫abμ(a,b)=b𝒫abμ(a,b)=μ(,)=1,\sum_{a,b\in\mathcal{P}}{\mu(a,b)}=\sum_{b\in\mathcal{P}}{\sum_{a\leqslant b}{\mu(a,b)}}=\sum_{b\in\mathcal{P}}{\sum_{\varnothing\leqslant a\leqslant b}{\mu(a,b)}}=\mu(\varnothing,\varnothing)=1,

where the penultimate equality follows from ACBμ(A,C)=δ(A,B)\sum_{A\leqslant C\leqslant B}\mu(A,C)=\delta(A,B). ∎

Remark 5.11.

We are going to use Lemma 5.10 several times, so, to avoid excessive combinatorial abstraction, we define η(x)=η𝒫(x)=b𝒫xμ(x,b)\eta(x)=\eta_{\mathcal{P}}(x)=\sum_{b\geqslant_{\mathcal{P}}x}\mu(x,b). Lemma 5.10 can be rewritten, then, as x𝒫η(x)=1\sum_{x\in\mathcal{P}}{\eta(x)}=1 whenever 𝒫\mathcal{P} has the minimal element. Observe that, if 𝒫\mathcal{P} is a meet-semilattice, η(x)\eta(x) is the coefficient of xx in the inclusion-exculsion formula written for the maximal elements A1,,AnA_{1},\ldots,A_{n} of the 𝒫\mathcal{P}. Indeed,

η(x)=bxμ(x,b)=1i1<<ikn(1)kj=1kAijbxμ(x,b)=1i1<<ikn(1)kδ(j=1kAij,x).\eta(x)=\sum_{b\geqslant x}\mu(x,b)=\sum_{1\leqslant i_{1}<\ldots<i_{k}\leqslant n}{(-1)^{k}\sum_{\bigcap_{j=1}^{k}A_{i_{j}}\geqslant b\geqslant x}{\mu(x,b)}}=\sum_{1\leqslant i_{1}<\ldots<i_{k}\leqslant n}{(-1)^{k}\cdot\delta\left(\bigcap_{j=1}^{k}A_{i_{j}},x\right)}.

Before formulating Theorem 5.12 rigorously, let us give its informal explanation. Basically, it claims that the complete flags of the graph correspondence Γf\Gamma_{f} are of the form

(X1,Y1)(E(M1),E(M2)),\varnothing\lessdot(X_{1},Y_{1})\lessdot\ldots\lessdot(E(M_{1}),E(M_{2})),

where Y1E(M2)\varnothing\lessdot Y_{1}\ldots\lessdot E(M_{2}) is a complete flag in M2M_{2}, images f(Xi)f(X_{i}) lie under YiY_{i} in (M2)\mathcal{F}(M_{2}) (but not necessarily equal to them), and the weights are balanced in such a way that the whole thing is a tropical fan and its push-forward onto B(M2)B(M_{2}) is B(M2)B(M_{2}). This way resembles inclusion-exclusion principle (see Remark 5.11).

Theorem 5.12.

Let Γf:B(M1)B(M2)\Gamma_{f}\colon B(M_{1})\vdash B(M_{2}) be a graph correspondence of the covering lattice map f:(M1)(M2)f\colon\mathcal{F}(M_{1})\to\mathcal{F}(M_{2}). Let (𝒢,ω𝒢)(\mathcal{G},\omega_{\mathcal{G}}) be the pair for which Γf\Gamma_{f} is a Flag fan. Then, all the edges of 𝒢\mathcal{G} are as follows:

(X1,Y1)𝒢(X2,Y2), where Y1M2Y2,f(X1)M2Y1 and X1M1X2.(X_{1},Y_{1})\lessdot_{\mathcal{G}}(X_{2},Y_{2}),\mbox{ where }\;Y_{1}\lessdot_{M_{2}}Y_{2},\;f(X_{1})\leqslant_{M_{2}}Y_{1}\mbox{ and }X_{1}\leqslant_{M_{1}}X_{2}.

Denote by (X1,X2,Y1)\mathcal{F}(X_{1},X_{2},Y_{1}) the subposet of (M1)\mathcal{F}(M_{1}) consisting of FF such that

X1M1FM1X2 and f(F)M2Y1.X_{1}\leqslant_{M_{1}}F\leqslant_{M_{1}}X_{2}\;\mbox{ and }\;f(F)\leqslant_{M_{2}}Y_{1}.

Then,

ω𝒢((X1,Y1),(X2,Y2))=η(X1,X2,Y1)(X1).\omega_{\mathcal{G}}((X_{1},Y_{1}),(X_{2},Y_{2}))=\eta_{\mathcal{F}(X_{1},X_{2},Y_{1})}(X_{1}).
Proof.

Denote the Flag fan γi(f)γ1(f)(B(M1)×B(M2))\gamma_{i}(f)\cdots\gamma_{1}(f)\cdot(B(M_{1})\times B(M_{2})) by Γf,i=Γi\Gamma_{f,i}=\Gamma_{i}, and the corresponding poset and weights pair by (𝒢i,ω𝒢i)(\mathcal{G}_{i},\omega_{\mathcal{G}_{i}}). It turns out to be convenient to digress for a moment from considering edges and to focus on the vertices of 𝒢\mathcal{G}. Notice that, if we show that only pairs (X,Y)(M1M2)(X,Y)\in\mathcal{F}(M_{1}\oplus M_{2}) such that f(X)M2Yf(X)\leqslant_{M_{2}}Y can belong to 𝒢\mathcal{G}, then the first claim follows. Indeed, the inequality X1M1X2X_{1}\leqslant_{M_{1}}X_{2} follows from the fact that 𝒢\mathcal{G} is the subposet of the initial lattice (M1M2)\mathcal{F}(M_{1}\oplus M_{2}), and Y1M2Y2Y_{1}\lessdot_{M_{2}}Y_{2} is obtained from the comparison of ranks in 𝒢\mathcal{G}. More precisely, observe that γ(f)((X,Y))=rkM1X\gamma(f)((X,Y))=-\operatorname{rk}_{M_{1}}{X} since Y=Yf(X)Y=Y\vee f(X), and their ranks in Definition 5.6 cancel each other out. Thus,

rk𝒢(X,Y)=rkM1M2(X,Y)+γ(f)((X,Y))=rkM2Y,\operatorname{rk}_{\mathcal{G}}{(X,Y)}=\operatorname{rk}_{M_{1}\oplus M_{2}}{(X,Y)}+\gamma(f)((X,Y))=\operatorname{rk}_{M_{2}}{Y},

which means that if Y2>M2Y1Y_{2}>_{M_{2}}Y_{1} but rkM2Y2rkM2Y1>1\operatorname{rk}_{M_{2}}{Y_{2}}-\operatorname{rk}_{M_{2}}{Y_{1}}>1, then there can be no edge in 𝒢\mathcal{G} between (X1,Y1)(X_{1},Y_{1}) and (X2,Y2)(X_{2},Y_{2}). But then there can be no edges between (X1,Y)(X_{1},Y) and (X2,Y)(X_{2},Y) in 𝒢\mathcal{G} as well, because any increasing chain must reach (E(M1),E(M2))(E(M_{1}),E(M_{2})) from \varnothing in rk𝒢=rkM2\operatorname{rk}{\mathcal{G}}=\operatorname{rk}{M_{2}} steps, with the rank in M2M_{2} not allowed to jump, so it must grow steadily by 1 on each edge of 𝒢\mathcal{G}.

Assume the contrary: there exists (X,Y)𝒢(X,Y)\in\mathcal{G} with Yf(X)YY\neq f(X)\vee Y. Take any maximal vertex with this property. Observe that, by definition of γ(f)\gamma(f), rk𝒢(X,Y)=rk𝒢(X,f(X)Y)\operatorname{rk}_{\mathcal{G}}{(X,Y)}=\operatorname{rk}_{\mathcal{G}}{(X,f(X)\vee Y)}, meaning that the pair (X,f(X)Y)(X,f(X)\vee Y) has its rank decreased by 11 with each next 𝒢i\mathcal{G}_{i} until minimal ii such that rk𝒢i(X,Y)=rk𝒢i(X,f(X)Y)\operatorname{rk}_{\mathcal{G}_{i}}{(X,Y)}=\operatorname{rk}_{\mathcal{G}_{i}}{(X,f(X)\vee Y)}, and then they continue the descent together. Consider the last step before (X,f(X)Y)(X,f(X)\vee Y) reaches (X,Y)(X,Y), namely, maximal ii that γi(f)((X,Y))=0\gamma_{i}(f)((X,Y))=0. We will prove that the localization γi(f)e(X,Y)\gamma_{i}(f)^{e_{(X,Y)}} is linear on the star of StarΓi1e(X,Y)\mathrm{Star}_{\Gamma_{i-1}}e_{(X,Y)}, so, by Lemma 2.25, (X,Y)(X,Y) does not belong to the next poset 𝒢i\mathcal{G}_{i} and, consequently, to 𝒢\mathcal{G}.

We show that γi(f)e(X,Y)=χy\gamma_{i}(f)^{e_{(X,Y)}}=-\chi_{y}, where y(f(X)Y)YE(M2)y\in(f(X)\vee Y)\setminus Y\subset E(M_{2}) is any element. Since the characteristic function χy\chi_{y} is linear on the StarΓi1e(X,Y)\mathrm{Star}_{\Gamma_{i-1}}e_{(X,Y)}, the claim follows. Note that we are not interested in the whole fan structure of the star, only that the rays are the elements (X,Y)(X^{\prime},Y^{\prime}) comparable with (X,Y)(X,Y). Those of them which are below (X,Y)(X,Y) have γi(f)e(X,Y)\gamma_{i}(f)^{e_{(X,Y)}} equal to 0. For (X,Y)𝒢i1(X,Y)(X^{\prime},Y^{\prime})\geqslant_{\mathcal{G}_{i-1}}(X,Y) we need to verify the equivalence:

yYγi(f)((X,Y))=1.y\in Y^{\prime}\Leftrightarrow\gamma_{i}(f)((X^{\prime},Y^{\prime}))=-1.

If yYy\notin Y^{\prime}, then, in particular, f(X)Yf(X)YyYf(X^{\prime})\vee Y^{\prime}\supset f(X)\vee Y\ni y\notin Y^{\prime}. Since γi(f)((X,Y))\gamma_{i}(f)((X^{\prime},Y^{\prime})) is already 1-1, the next γj(f)((X,Y))\gamma_{j}(f)((X^{\prime},Y^{\prime})) are going to be 1-1 for all jij\geqslant i, so the vertex (X,Y)(X^{\prime},Y^{\prime}) never disappears and thus belongs to 𝒢\mathcal{G}, which contradicts (X,Y)(X,Y) being a maximal element of 𝒢\mathcal{G} with Yf(X)YY\neq f(X)\vee Y. This proves \Leftarrow.

If yYy\in Y^{\prime}, then

γ(f)((X,Y))γ(f)((X,Y))=rkM2(f(X)Y)rkM1XrkM2Y<<rkM2(f(X)Y)rkM1XrkM2Y=γ(f)((X,Y)),\begin{split}\gamma(f)((X^{\prime},Y^{\prime}))\leqslant\gamma(f)((X,Y^{\prime}))=\operatorname{rk}_{M_{2}}(f(X)\cup Y^{\prime})-\operatorname{rk}_{M_{1}}{X}-\operatorname{rk}_{M_{2}}{Y^{\prime}}<\\ <\operatorname{rk}_{M_{2}}(f(X)\cup Y)-\operatorname{rk}_{M_{1}}{X}-\operatorname{rk}_{M_{2}}{Y}=\gamma(f)((X,Y)),\end{split} (5.1)

which implies γi(f)((X,Y))=1\gamma_{i}(f)((X^{\prime},Y^{\prime}))=-1 since γi+1(f)((X,Y))=1\gamma_{i+1}(f)((X,Y))=-1. The strict inequality in equation 5.1 follows from Lemma 2.2 and the fact that yf(X)Yy\in f(X)\vee Y. This proves \Rightarrow.

It remains to calculate ω𝒢((X1,Y1),(X2,Y2))\omega_{\mathcal{G}}((X_{1},Y_{1}),(X_{2},Y_{2})) in the case where both vertices belong to 𝒢\mathcal{G}. We will perform it by induction on the number of step ii when the edge ((X1,Y1),(X2,Y2))((X_{1},Y_{1}),(X_{2},Y_{2})) appears in 𝒢i\mathcal{G}_{i} (step 2 of Corollary 3.8). This is made possible by the fact that, as we are going to see, all the edge weights on which ω𝒢((X1,Y1),(X2,Y2))\omega_{\mathcal{G}}((X_{1},Y_{1}),(X_{2},Y_{2})) depends are determined for earlier ii.

Poset (X1,X2,Y1)\mathcal{F}(X_{1},X_{2},Y_{1}) does not always have the top element (which is precisely the reason for the weights to get complicated), but it always has the bottom element X1X_{1}. Therefore, in the base of induction (X1,X2,Y1)\mathcal{F}(X_{1},X_{2},Y_{1}) consists of a single element X1X_{1}. Indeed, if there is another X(X1,X2,Y1)X^{\prime}\in\mathcal{F}(X_{1},X_{2},Y_{1}), the edge ((X,Y1),(X2,Y2))((X^{\prime},Y_{1}),(X_{2},Y_{2})) would appear before, since

γ(f)((X,Y1))=rkM1X<rkM1X1=γ(f)((X1,Y1)).\gamma(f)((X^{\prime},Y_{1}))=-\operatorname{rk}_{M_{1}}{X^{\prime}}<-\operatorname{rk}_{M_{1}}{X_{1}}=\gamma(f)((X_{1},Y_{1})).

We aim to show that ω𝒢((X1,Y1),(X2,Y2))=1\omega_{\mathcal{G}}((X_{1},Y_{1}),(X_{2},Y_{2}))=1.

The edge between vertices (X1,Y1)(X_{1},Y_{1}) and (X2,Y2)(X_{2},Y_{2}) appears when taking the Weil divisor of γi(f)\gamma_{i}(f) for such ii that γi(f)(((X1,Y1)))=0\gamma_{i}(f)(((X_{1},Y_{1})))=0, γi(f)(((X2,Y2)))=1\gamma_{i}(f)(((X_{2},Y_{2})))=-1 and γi+1(f)(((X1,Y1)))=1\gamma_{i+1}(f)(((X_{1},Y_{1})))=-1. As we have seen in Corollary 3.8, the weight on the newly-created edge is equal to

deg(γi(f)|(Γi1)(X1,Y1),(X2,Y2)(Γi1)(X1,Y1),(X2,Y2)),\deg{(\gamma_{i}(f)|_{(\Gamma_{i-1})_{(X_{1},Y_{1}),(X_{2},Y_{2})}}\cdot(\Gamma_{i-1})_{(X_{1},Y_{1}),(X_{2},Y_{2})})},

which, after expanding according to Definition 2.14, yields

γi(f)((X1,Y1)𝒢i1F𝒢i1(X2,Y2)ω𝒢i1((X1,Y1),F)ω𝒢i1(F,(X2,Y2))eF)++(X1,Y1)𝒢i1F𝒢i1(X2,Y2)ω𝒢i1((X1,Y1),F)ω𝒢i1(F,(X2,Y2))γi(f)(F).\begin{split}-\gamma_{i}(f)\left(\sum_{(X_{1},Y_{1})\lessdot_{\mathcal{G}_{i-1}}F^{\prime}\lessdot_{\mathcal{G}_{i-1}}(X_{2},Y_{2})}\omega_{\mathcal{G}_{i-1}}((X_{1},Y_{1}),F^{\prime})\omega_{\mathcal{G}_{i-1}}(F^{\prime},(X_{2},Y_{2}))\cdot e_{F^{\prime}}\right)+\\ +\sum_{(X_{1},Y_{1})\lessdot_{\mathcal{G}_{i-1}}F^{\prime}\lessdot_{\mathcal{G}_{i-1}}(X_{2},Y_{2})}\omega_{\mathcal{G}_{i-1}}((X_{1},Y_{1}),F^{\prime})\omega_{\mathcal{G}_{i-1}}(F^{\prime},(X_{2},Y_{2}))\cdot\gamma_{i}(f)(F^{\prime}).\end{split} (5.2)

Due to the balancing condition written for the hyperface of Γi1\Gamma_{i-1} generated by any chain containing (X1,Y1)(X_{1},Y_{1}) and (X2,Y2)(X_{2},Y_{2}) with missing FF^{\prime}, the sum in the large brackets is just some ce(X2,Y2)c\cdot e_{(X_{2},Y_{2})}. Then, since γi(f)((X2,Y2))=1\gamma_{i}(f)((X_{2},Y_{2}))=-1, the first summand is just cc. We want to show that γi(f)(F)=0\gamma_{i}(f)(F^{\prime})=0 for all FF^{\prime} in the sum, so that the second term is 0, and that c=1c=1.

Recall that Y2M2Y1Y_{2}\gtrdot_{M_{2}}Y_{1}, thus every FF^{\prime} has the form (X,Y1)(X^{\prime},Y_{1}) or (X,Y2)(X^{\prime},Y_{2}). In the former case, X2M1X>M1X1X_{2}\geqslant_{M_{1}}X^{\prime}>_{M_{1}}X_{1}, but (X1,X2,Y1)\mathcal{F}(X_{1},X_{2},Y_{1}) consists of X1X_{1} only, therefore, Y1f(X)Y1Y_{1}\neq f(X^{\prime})\vee Y_{1}. Thus, if γi(f)((X,Y1))=1\gamma_{i}(f)((X^{\prime},Y_{1}))=-1, we get a contradiction (this vertex should not exist already in 𝒢i1\mathcal{G}_{i-1}). In the latter case, since f(X)M2f(X2)M2Y2f(X^{\prime})\leqslant_{M_{2}}f(X_{2})\leqslant_{M_{2}}Y_{2}, values of γ(f)\gamma(f) on both (X,Y2)(X^{\prime},Y_{2}) and (X2,Y2)(X_{2},Y_{2}) are equal to the ranks rkM1X,rkM1X2\operatorname{rk}_{M_{1}}{X^{\prime}},\operatorname{rk}_{M_{1}}{X_{2}}, respectively, and γi(f)\gamma_{i}(f) cannot coincide on them unless they are of the same rank in 𝒢i1\mathcal{G}_{i-1}, in which case (X2,Y2)𝒢i1(X,Y2)(X_{2},Y_{2})\gtrdot_{\mathcal{G}_{i-1}}(X^{\prime},Y_{2}) could not occur.

To see that c=1c=1, focus on those FF^{\prime} of the form (X,Y2)(X^{\prime},Y_{2}) (in other words, count only the coordinate of any element yY2Y1y\in Y_{2}\setminus Y_{1} in the vector ce(X2,Y2)c\cdot e_{(X_{2},Y_{2})}). As we have just seen, if X>X1X^{\prime}>X_{1}, then (X2,Y2)(X_{2},Y_{2}) cannot cover (X,Y2)(X^{\prime},Y_{2}) in 𝒢i1\mathcal{G}_{i-1}, so there is only one such F=(X1,Y2)F^{\prime}=(X_{1},Y_{2}). Now, ω𝒢i1((X1,Y1),(X1,Y2))=1\omega_{\mathcal{G}_{i-1}}((X_{1},Y_{1}),(X_{1},Y_{2}))=1, because this is the original covering edge from (M1M2)\mathcal{F}(M_{1}\oplus M_{2}), where all the weights are 11, so it remains to prove that ω𝒢i1((X1,Y2),(X2,Y2))=1\omega_{\mathcal{G}_{i-1}}((X_{1},Y_{2}),(X_{2},Y_{2}))=1 as well. This is easy, though: γ(f)\gamma(f) coincides with rkM1-\operatorname{rk}_{M_{1}} on the whole segment [(X1,Y2),(X2,Y2)][(X_{1},Y_{2}),(X_{2},Y_{2})] of (M1M2)\mathcal{F}(M_{1}\oplus M_{2}), and thus all newly-created edges always have weight 11 there. Indeed, for each ii value of γi(f)(X′′,Y2)\gamma_{i}(f)(X^{\prime\prime},Y_{2}) equals 1-1 if X′′X^{\prime\prime} is above certain rank in M1M_{1}, so the value of each γi(f)\gamma_{i}(f) on the level of missing flats is 0, thus they never impact the weight of the new edge. It follows that there is only one FF^{\prime} containing Y2Y_{2}, and its weight is 1, so c=1c=1 as claimed. This completes the base of induction.

Let us now find ω𝒢((X1,Y1),(X2,Y2))\omega_{\mathcal{G}}((X_{1},Y_{1}),(X_{2},Y_{2})) in the case of arbitrary (X1,X2,Y1)\mathcal{F}(X_{1},X_{2},Y_{1}). We begin as previously, fixing ii such that this edge appears when taking the Weil divisor of γi(f)\gamma_{i}(f), and writing equation 5.2 for its weight. This time, though, we already know that

ω𝒢i1((X,Y1),(X2,Y2))=η(X,X2,Y1)(X)\omega_{\mathcal{G}_{i-1}}((X^{\prime},Y_{1}),(X_{2},Y_{2}))=\eta_{\mathcal{F}(X^{\prime},X_{2},Y_{1})}(X^{\prime})

for X>M1X1X^{\prime}>_{M_{1}}X_{1} that belong to (X1,X2,Y1)\mathcal{F}(X_{1},X_{2},Y_{1}) by the induction assumption. We have also seen that ω𝒢i1((X1,Y1),(X,Y1))=1\omega_{\mathcal{G}_{i-1}}((X_{1},Y_{1}),(X^{\prime},Y_{1}))=1 (as in the previous paragraph, γ(f)\gamma(f) coincides with rkM1-\operatorname{rk}_{M_{1}} on the whole segment [(X1,Y1),(X,Y1)][(X_{1},Y_{1}),(X^{\prime},Y_{1})]). Thus, the second term of equation 5.2 becomes

XX(X1,X2,Y1)(η(X,X2,Y1)(X)).-\sum_{X\neq X^{\prime}\in\mathcal{F}(X_{1},X_{2},Y_{1})}\left(\eta_{\mathcal{F}(X^{\prime},X_{2},Y_{1})}(X^{\prime})\right).

The first term still equals 11, since c=1c=1 by exactly the same argument as before — (X1,Y2)(X_{1},Y_{2}) is the only vertex between (X1,Y1)(X_{1},Y_{1}) and (X2,Y2)(X_{2},Y_{2}) containing elements of Y2Y1Y_{2}\setminus Y_{1}, and both edges connecting it to (X1,Y1)(X_{1},Y_{1}) and (X2,Y2)(X_{2},Y_{2}) have weight 11 in ω𝒢i1\omega_{\mathcal{G}_{i-1}}. Thus, we need

η(X1,X2,Y1)(X1)=1XX(X1,X2,Y1)(η(X,X2,Y1)(X)),\eta_{\mathcal{F}(X_{1},X_{2},Y_{1})}(X_{1})=1-\sum_{X\neq X^{\prime}\in\mathcal{F}(X_{1},X_{2},Y_{1})}\left(\eta_{\mathcal{F}(X^{\prime},X_{2},Y_{1})}(X^{\prime})\right),

which rearranges to

X(X1,X2,Y1)η(X,X2,Y1)(X)=1,\sum_{X^{\prime}\in\mathcal{F}(X_{1},X_{2},Y_{1})}\eta_{\mathcal{F}(X^{\prime},X_{2},Y_{1})}(X^{\prime})=1,

and this is simply Lemma 5.10. ∎

We can now use Theorem 5.12 to obtain

Theorem 5.13.

In the notation of Definition 5.6, a map between Hom-sets of 𝐒𝐌𝐚𝐭𝐫𝐖𝐋\mathbf{SMatrWL_{\gtrdot}} and 𝐒𝐌𝐚𝐭𝐫𝐂𝐨𝐫\mathbf{SMatrCor} given by fΓff\mapsto\Gamma_{f} is a functor: if f:(M1)(M2)f\colon\mathcal{F}(M_{1})\to\mathcal{F}(M_{2}) and g:(M2)(M3)g\colon\mathcal{F}(M_{2})\to\mathcal{F}(M_{3}) are covering lattice maps between simple matroids, then

Γgf=ΓgΓf.\Gamma_{g\circ f}=\Gamma_{g}\circ\Gamma_{f}.
Proof.

Despite seemingly cumbersome construction of ΓgΓf\Gamma_{g}\circ\Gamma_{f} and the fact that in general we have no idea whether the push-forward of the Flag fan is again a Flag fan, or, even if it is, what are its (𝒢,ω𝒢)(\mathcal{G},\omega_{\mathcal{G}}) — despite all that, we already possess almost all the necessary tools to verify the claim.

For simplicity, denote B(M1)=BX,B(M2)=BY,B(M3)=BZB(M_{1})=B_{X},B(M_{2})=B_{Y},B(M_{3})=B_{Z}. Thus,

ΓfBX×BY,ΓgBY×BZ,ΓgfBX×BZ,\Gamma_{f}\subset B_{X}\times B_{Y},\;\Gamma_{g}\subset B_{Y}\times B_{Z},\;\Gamma_{g\circ f}\subset B_{X}\times B_{Z},

and we need to show that

(πXZ)[(Γf×BZ)(BX×Γg)]=Γgf.(\pi_{XZ})_{*}\left[(\Gamma_{f}\times B_{Z})\cdot(B_{X}\times\Gamma_{g})\right]=\Gamma_{g\circ f}.

Using the fact that, for the subcycles of Bergman fans which are cut by Weil divisors of rational functions φi\varphi_{i}, the intersections via diagonal construction and via φi\varphi_{i}’s coincide ([FR12], Theorem 4.5(6)), we will check the following:

(πXZ)(πXY(γ(f))(BX×Γg))=γ(gf)(BX×BZ).(\pi_{XZ})_{*}\left(\prod\pi_{XY}^{*}(\gamma(f))\cdot(B_{X}\times\Gamma_{g})\right)=\prod\gamma(g\circ f)\cdot(B_{X}\times B_{Z}).

To control edge weights of πXY(γ(f))(BX×Γg)\prod\pi_{XY}^{*}(\gamma(f))\cdot(B_{X}\times\Gamma_{g}), where BX×ΓgB_{X}\times\Gamma_{g} is not necessarily a Bergman fan anymore, we will need an almost verbatim generalization of Theorem 5.12 which we formulate as a standalone lemma.

Lemma 5.14.

In the notation of Theorem 5.13, if (,ω)(\mathcal{H},\omega_{\mathcal{H}}) is the pair corresponding to the Flag fan πXY(γi(f))(BX×Γg)\prod\pi_{XY}^{*}(\gamma_{i}(f))\cdot(B_{X}\times\Gamma_{g}), then all the edges of \mathcal{H} are as follows:

(X1,Y1,Z1)(X2,Y2,Z2), where (Y1,Z1)𝒢YZ(Y2,Z2),f(X1)M2Y1 and X1M1X2.(X_{1},Y_{1},Z_{1})\lessdot_{\mathcal{H}}(X_{2},Y_{2},Z_{2}),\mbox{ where }\;(Y_{1},Z_{1})\lessdot_{\mathcal{G}_{YZ}}(Y_{2},Z_{2}),\;f(X_{1})\leqslant_{M_{2}}Y_{1}\mbox{ and }X_{1}\leqslant_{M_{1}}X_{2}.

The weight of the edge is given by the formula

ω((X1,Y1,Z1),(X2,Y2,Z2))=ω𝒢YZ((Y1,Z1),(Y2,Z2))η(X1,X2,Y1)(X1).\omega_{\mathcal{H}}((X_{1},Y_{1},Z_{1}),(X_{2},Y_{2},Z_{2}))=\omega_{\mathcal{G}_{YZ}}((Y_{1},Z_{1}),(Y_{2},Z_{2}))\cdot\eta_{\mathcal{F}(X_{1},X_{2},Y_{1})}(X_{1}).
Proof.

By Lemma 3.10 the edges of BX×ΓgB_{X}\times\Gamma_{g} before cutting Weil divisors are of two types. The first type is ((X1,Y,Z),(X2,Y,Z))((X_{1},Y,Z),(X_{2},Y,Z)), where X1M1X2X_{1}\lessdot_{M_{1}}X_{2} and (Y,Z)𝒢YZ(Y,Z)\in\mathcal{G}_{YZ}. The weight of such an edge is 11, as in (M1)\mathcal{F}(M_{1}). The second type is ((X,Y1,Z1),(X,Y2,Z2))((X,Y_{1},Z_{1}),(X,Y_{2},Z_{2})), where (Y1,Z1)𝒢YZ(Y2,Z2)(Y_{1},Z_{1})\lessdot_{\mathcal{G}_{YZ}}(Y_{2},Z_{2}). The weight of such an edge is η(Y1,Y2,Z1)(Y1)\eta_{\mathcal{F}(Y_{1},Y_{2},Z_{1})}(Y_{1}).

The claim is then verified analogously to Theorem 5.12. First, we show that only vertices with f(X1)M2Y1f(X_{1})\leqslant_{M_{2}}Y_{1} survive in \mathcal{H}. It follows that for any edge ((X1,Y1,Z1),(X2,Y2,Z2))((X_{1},Y_{1},Z_{1}),(X_{2},Y_{2},Z_{2})) of \mathcal{H} the pair (Y2,Z2)(Y_{2},Z_{2}) must cover (Y1,Z1)(Y_{1},Z_{1}) in 𝒢YZ\mathcal{G}_{YZ}. To show this, as before, assume the contrary, and take any maximal surviving vertex (X1,Y1,Z1)(X_{1},Y_{1},Z_{1})\in\mathcal{H} with f(X1)Y1Y1f(X_{1})\vee Y_{1}\neq Y_{1}. Consider the step ii after which the ranks in i\mathcal{H}_{i} of all the vertices (X1,f(X1)Y1,Z)(X_{1},f(X_{1})\vee Y_{1},Z^{\prime}) become equal to the rank of (X1,Y1,Z1)(X_{1},Y_{1},Z_{1}). See that at this step, localized function πXY(γi(f))e(X1,Y1,Z1)\pi_{XY}^{*}(\gamma_{i}(f))^{e_{(X_{1},Y_{1},Z_{1})}} on the star of e(X1,Y1,Z1)e_{(X_{1},Y_{1},Z_{1})} is equal to minus the characteristic function of any element of (f(X1)Y1)Y1(f(X_{1})\vee Y_{1})\setminus Y_{1}.

Then, we verify that the weights are as claimed using induction on applying πXY(γi(f))\pi_{XY}^{*}(\gamma_{i}(f)). The necessary observations are that

  • all the initial covering relations coming from M1M_{1} have weights 11;

  • all the newly-created edges of the type ((X1,Y,Z),(X2,Y,Z))((X_{1},Y,Z),(X_{2},Y,Z)) between the vertices present in \mathcal{H} also have weights 11, because πXY(γ(f))\pi_{XY}^{*}(\gamma(f)) coincides with rkM1-rk_{M_{1}} on the segment;

  • all the initial covering relations coming from 𝒢YZ\mathcal{G}_{YZ} inherit their weights from 𝒢YZ\mathcal{G}_{YZ}, thus, in induction handling equation 5.2, everything is multiplied by η(Y1,Y2,Z1)(Y1)\eta_{\mathcal{F}(Y_{1},Y_{2},Z_{1})}(Y_{1}).

According to Lemma 5.14 and Definition 2.20, the weight of the maximal cone generated by the chain (X1,Z1)(E(M1),E(M3))\varnothing\lessdot(X_{1},Z_{1})\lessdot\ldots\lessdot(E(M_{1}),E(M_{3})) of length r=rkM3r=\operatorname{rk}{M_{3}} in the push-forward is given by

Y0Yr(i=0r1ω𝒢YZ((Yi,Yi+1),(Zi,Zi+1))i=0r1ω𝒢XY((Xi,Xi+1),(Yi,Yi+1))),\sum_{Y_{0}\leqslant\ldots\leqslant Y_{r}}\left(\prod_{i=0}^{r-1}\omega_{\mathcal{G}_{YZ}}((Y_{i},Y_{i+1}),(Z_{i},Z_{i+1}))\cdot\prod_{i=0}^{r-1}\omega_{\mathcal{G}_{XY}}((X_{i},X_{i+1}),(Y_{i},Y_{i+1}))\right), (5.3)

where we assume summation over only suitable sequences YiY_{i} (mind that, unlike ZiZ_{i}, they do not have to be strictly increasing in (M2)\mathcal{F}(M_{2})). Next, we rearrange

i=0r1ω𝒢YZ((Yi,Yi+1),(Zi,Zi+1))=i=0r1(Yi(Yi,Yi+1,Zi)μ(Yi,Yi))==Y0Yr(i=0r1μ(Yi,Yi+1,Zi)(Yi,Yi)),\begin{split}\prod_{i=0}^{r-1}\omega_{\mathcal{G}_{YZ}}((Y_{i},Y_{i+1}),(Z_{i},Z_{i+1}))=\prod_{i=0}^{r-1}\left(\sum_{Y^{\prime}_{i}\in\mathcal{F}(Y_{i},Y_{i+1},Z_{i})}\mu(Y_{i},Y^{\prime}_{i})\right)=\\ =\sum_{Y^{\prime}_{0}\ldots Y^{\prime}_{r}}\left(\prod_{i=0}^{r-1}\mu_{\mathcal{F}(Y_{i},Y_{i+1},Z_{i})}(Y_{i},Y^{\prime}_{i})\right),\end{split}

and, likewise,

i=0r1ω𝒢XY((Xi,Xi+1),(Yi,Yi+1))=X0Xr(i=0r1μ(Xi,Xi+1,Yi)(Xi,Xi)).\prod_{i=0}^{r-1}\omega_{\mathcal{G}_{XY}}((X_{i},X_{i+1}),(Y_{i},Y_{i+1}))=\sum_{X^{\prime}_{0}\ldots X^{\prime}_{r}}\left(\prod_{i=0}^{r-1}\mu_{\mathcal{F}(X_{i},X_{i+1},Y_{i})}(X_{i},X^{\prime}_{i})\right).

Substituting into equation 5.3 and rearranging, we get (dropping the subscripts of μ\mu’s)

Y0YrY0YrX0Xr(i=0r1μ(Yi,Yi)μ(Xi,Xi)).\sum_{Y_{0}\ldots Y_{r}}\sum_{\begin{subarray}{c}Y^{\prime}_{0}\ldots Y^{\prime}_{r}\\ X^{\prime}_{0}\ldots X^{\prime}_{r}\end{subarray}}\left(\prod_{i=0}^{r-1}\mu(Y_{i},Y^{\prime}_{i})\mu(X_{i},X^{\prime}_{i})\right). (5.4)

The system of sets Xi,Yi,YiX^{\prime}_{i},Y_{i},Y^{\prime}_{i} satisfies the following conditions:

XiXiXi+1YiYi+1YiYiYi+1f(Xi)Yig(Yi)Zig(Yi)Zi\begin{matrix}X_{i}\leqslant X^{\prime}_{i}\leqslant X_{i+1}&Y_{i}\leqslant Y_{i+1}&Y_{i}\leqslant Y^{\prime}_{i}\leqslant Y_{i+1}\\ f(X^{\prime}_{i})\leqslant Y_{i}&g(Y_{i})\leqslant Z_{i}&g(Y_{i}^{\prime})\leqslant Z_{i}\end{matrix}

In the initial sum 5.3 we’ve chosen YiY_{i}’s satisfying conditions from the middle column and then XiX^{\prime}_{i}’s and YiY^{\prime}_{i}’s satisfying the rest of conditions. We obtain the same system of sets by first choosing XiX^{\prime}_{i} such that

XiXiXi+1 and g(f(Xi))Zi,X_{i}\leqslant X^{\prime}_{i}\leqslant X_{i+1}\;\mbox{ and }\;g(f(X^{\prime}_{i}))\leqslant Z_{i}, (5.5)

and then choosing

f(Xi)f(Xi+1)YiYiYi+1Yi+1g(Yi)Zig(Yi+1)Zi+1\setcounter{MaxMatrixCols}{11}\begin{matrix}&&f(X^{\prime}_{i})&&&&f(X^{\prime}_{i+1})&&&&\\ &&\rotatebox[origin={c}]{90.0}{$\geqslant$}&&&&\rotatebox[origin={c}]{90.0}{$\geqslant$}&&&&\\ \ldots&\leqslant&Y_{i}&\leqslant&Y^{\prime}_{i}&\leqslant&Y_{i+1}&\leqslant&Y^{\prime}_{i+1}&\leqslant&\ldots\\ &&&&\rotatebox[origin={c}]{270.0}{$\mapsto$}&&&&\rotatebox[origin={c}]{270.0}{$\mapsto$}&&\\ &&&&g(Y^{\prime}_{i})&\leqslant&Z_{i}&&g(Y^{\prime}_{i+1})&\leqslant&Z^{\prime}_{i+1}\end{matrix} (5.6)

Thus, the summation in equation 5.4 can be rewritten as

X0Xr(i=0r1μ(Xi,Xi)Y0YrY0Yri=0r1μ(Yi,Yi)),\sum_{X^{\prime}_{0}\ldots X^{\prime}_{r}}\left(\prod_{i=0}^{r-1}\mu(X_{i},X^{\prime}_{i})\cdot\sum_{\begin{subarray}{c}Y_{0}\ldots Y_{r}\\ Y^{\prime}_{0}\ldots Y^{\prime}_{r}\end{subarray}}\prod_{i=0}^{r-1}\mu(Y_{i},Y^{\prime}_{i})\right),

while the desired coefficient of the same maximal cone generated by the chain (X1,Z1)(E(M1),E(M3))\varnothing\lessdot(X_{1},Z_{1})\lessdot\ldots\lessdot(E(M_{1}),E(M_{3})) in Γgf\Gamma_{g\circ f} is just

i=0r1ω𝒢XZ((Xi,Xi+1),(Zi,Zi+1))=X0Xr(i=0r1μ(Xi,Xi)),\prod_{i=0}^{r-1}\omega_{\mathcal{G}_{XZ}}((X_{i},X_{i+1}),(Z_{i},Z_{i+1}))=\sum_{X^{\prime}_{0}\ldots X^{\prime}_{r}}\left(\prod_{i=0}^{r-1}\mu(X_{i},X^{\prime}_{i})\right),

where the summation is taken over XiX^{\prime}_{i} with same conditions as in 5.5, so we need to show that

Y0YrY0Yri=0r1μ(Yi,Yi)=1.\sum_{\begin{subarray}{c}Y_{0}\ldots Y_{r}\\ Y^{\prime}_{0}\ldots Y^{\prime}_{r}\end{subarray}}\prod_{i=0}^{r-1}\mu(Y_{i},Y^{\prime}_{i})=1.

By Lemma 5.10, for a poset 𝒫\mathcal{P} with bottom element, a𝒫bμ(a,b)=1\sum_{a\leqslant_{\mathcal{P}}b}\mu(a,b)=1. Making choices of pairs (Yi,Yi)(Y_{i},Y^{\prime}_{i}) one by one, we obtain

Y0YrY0Yri=0r1μ(Yi,Yi)=Y0,Y0μ(Y0,Y0)Y1YrY1Yri=1r1μ(Yi,Yi)==Y0,Y0μ(Y0,Y0)Y1,Y1μ(Y1,Y1)Y2YrY2Yri=2r1μ(Yi,Yi)\begin{split}\sum_{\begin{subarray}{c}Y_{0}\ldots Y_{r}\\ Y^{\prime}_{0}\ldots Y^{\prime}_{r}\end{subarray}}\prod_{i=0}^{r-1}\mu(Y_{i},Y^{\prime}_{i})=\sum_{Y_{0},Y^{\prime}_{0}}\mu(Y_{0},Y^{\prime}_{0})\sum_{\begin{subarray}{c}Y_{1}\ldots Y_{r}\\ Y^{\prime}_{1}\ldots Y^{\prime}_{r}\end{subarray}}\prod_{i=1}^{r-1}\mu(Y_{i},Y^{\prime}_{i})=\\ =\sum_{Y_{0},Y^{\prime}_{0}}\mu(Y_{0},Y^{\prime}_{0})\sum_{Y_{1},Y^{\prime}_{1}}\mu(Y_{1},Y^{\prime}_{1})\sum_{\begin{subarray}{c}Y_{2}\ldots Y_{r}\\ Y^{\prime}_{2}\ldots Y^{\prime}_{r}\end{subarray}}\prod_{i=2}^{r-1}\mu(Y_{i},Y^{\prime}_{i})\ldots\end{split}

where the summation (Yj,Yj)(Y_{j},Y^{\prime}_{j}) goes over pairs that satisfy all the conditions of 5.6 concerning XiX_{i} and ZiZ_{i}, and also YjYj1Y_{j}\geqslant Y^{\prime}_{j-1}. Each of those modified subsets has the minimal element Yj1f(Xj)Y^{\prime}_{j-1}\vee f(X^{\prime}_{j}), therefore, the whole sum turns into 11 from the end to the beginning. ∎

Remark 5.15.

It is possible to define Γf\Gamma_{f} without ff itself, using the description from Theorem 5.12, thus avoiding its proof. There are several reasons, though, for which we find our approach preferable. Firstly, the fact that Γf\Gamma_{f} is a Flag tropical fan is now an instance of Lemma 3.7, and we do not need to verify balancing conditions of Γf\Gamma_{f} combinatorially. Secondly, using ff to construct Γf\Gamma_{f} turns graph correspondence into a rigorous generalization of diagonal ΔB(N)×B(N)\Delta\subset B(N)\times B(N) and Bergman subfan B(N)B(M)B(N)\subset B(M) from [FR12] (Example 5.7 (2)). Thirdly, Γf\Gamma_{f} is defined even if ff is not a covering lattice map, while it is not a Flag fan and does not admit this kind of description — this may turn out to be valuable to us (see Question 5.19 below). Finally, ff clearly stores information about Γf\Gamma_{f} in a more compact way. The next statement hints that, maybe, we can even work with γ(f)\gamma(f) instead of ff when trying to generalize claims like Theorem 5.13.

Lemma 5.16.

If f,gf,g are covering lattice maps, then γ(gf)\gamma(g\circ f) can be defined as

γ(gf)(FX,FZ)=minFY(M2)(γ(g)(FY,FZ)+γ(f)(FX,FY)+rkM2FY).\gamma(g\circ f)(F_{X},F_{Z})=\min_{F_{Y}\in\mathcal{F}(M_{2})}{(\gamma(g)(F_{Y},F_{Z})+\gamma(f)(F_{X},F_{Y})+\operatorname{rk}_{M_{2}}{F_{Y}})}.
Proof.

By Definition 5.6 we have

γ(f)((FX,FY))=rkM2(f(FX)FY)rkM1FXrkM2FY,γ(g)((FY,FZ))=rkM3(g(FY)FZ)rkM2FYrkM3FZ.\begin{split}\gamma(f)((F_{X},F_{Y}))=\operatorname{rk}_{M_{2}}(f(F_{X})\cup F_{Y})-\operatorname{rk}_{M_{1}}F_{X}-\operatorname{rk}_{M_{2}}F_{Y},\\ \gamma(g)((F_{Y},F_{Z}))=\operatorname{rk}_{M_{3}}(g(F_{Y})\cup F_{Z})-\operatorname{rk}_{M_{2}}F_{Y}-\operatorname{rk}_{M_{3}}F_{Z}.\end{split}

Substituting FY=f(FX)F_{Y}=f(F_{X}), we get

γ(g)(f(FX),FZ)+γ(f)(FX,f(FX))+rkM2f(FX)==rkM3(g(f(FX))FZ)rkM1FXrkM3FZ=γ(gf)(FX,FZ).\begin{split}\gamma(g)(f(F_{X}),F_{Z})+\gamma(f)(F_{X},f(F_{X}))+\operatorname{rk}_{M_{2}}{f(F_{X})}=\\ =\operatorname{rk}_{M_{3}}(g(f(F_{X}))\cup F_{Z})-\operatorname{rk}_{M_{1}}F_{X}-\operatorname{rk}_{M_{3}}F_{Z}=\gamma(g\circ f)(F_{X},F_{Z}).\end{split}

Thus, it remains to prove that for any other F(M2)F^{\prime}\in\mathcal{F}(M_{2}) the sum is not smaller, which is equivalent to

rkM2(f(FX)F)rkM2FrkM3(g(f(FX))FZ)rkM3(g(F)FZ).\operatorname{rk}_{M_{2}}(f(F_{X})\cup F^{\prime})-\operatorname{rk}_{M_{2}}{F^{\prime}}\geqslant\operatorname{rk}_{M_{3}}(g(f(F_{X}))\cup F_{Z})-\operatorname{rk}_{M_{3}}(g(F^{\prime})\cup F_{Z}).

Since gg is a covering map, applying it does not increase the difference of ranks, therefore,

rkM2(f(FX)F)rkM2FrkM3(g(f(FX)F))rkM3g(F)rkM3(g(f(FX)))rkM3g(F),\operatorname{rk}_{M_{2}}(f(F_{X})\cup F^{\prime})-\operatorname{rk}_{M_{2}}{F^{\prime}}\geqslant\operatorname{rk}_{M_{3}}(g(f(F_{X})\cup F^{\prime}))-\operatorname{rk}_{M_{3}}{g(F^{\prime})}\geqslant\operatorname{rk}_{M_{3}}(g(f(F_{X})))-\operatorname{rk}_{M_{3}}{g(F^{\prime})},

and after joining both terms with FZF_{Z} the difference of ranks cannot increase by Lemma 2.2. ∎

Remark 5.17.

As we can see from Theorem 5.13 and Example 5.3, a reasonable subcategory of 𝐒𝐌𝐚𝐭𝐫𝐂𝐨𝐫\mathbf{SMatrCor} will not have tropical fibre product as pullback. More precisely, if M1M_{1} is a uniform matroid of rank 33 on the groundset {1,2,3,4}\{1,2,3,4\}, M2M_{2} is a uniform matroid of rank 33 on the groundset {1,2,3,5}\{1,2,3,5\} and NN is a uniform matroid of rank 33 on the groundset {1,2,3}\{1,2,3\}, then the tropical fibre product is B(M)B(M), where MM is a uniform matroid of rank 33 on the groundset {1,2,3,4,5}\{1,2,3,4,5\}. Consider Bergman fan B(M)B(M^{\prime}) of another amalgam MM^{\prime} on the groundset {1,2,3,4,5}\{1,2,3,4,5\} with 4,54,5 being parallel elements. Then, there are numerous correspondences MMM\vdash M^{\prime} making 44 and 55 coincide, exactly as in Example 5.3. All of them commute with the projection graph correspondences M1,2M,MM_{1,2}\vdash M,M^{\prime}, though.

Remark 5.18.

Note that in the image of the 𝐒𝐌𝐚𝐭𝐫𝐖𝐋\mathbf{SMatrWL_{\gtrdot}} under the forgetful functor Pt\mathrm{Pt} the tropical fibre product does not satisfy the universal property 2.1. We show this by constructing an example of the proper amalgam and another amalgam such that the unique weak map between their groundset can only come from the non-covering lattice map — see Figure 9. One verifies that MM is the tropical fibre product, but the identity map of the groundsets MMM\to M^{\prime} cannot be obtained from covering lattice map. Indeed, the flat {7,8}\{7,8\} of MM cannot map to anything other than {7,8}\{7,8\} — if it maps to, say, {1,7,8}\{1,7,8\}, then {4,5,6,7,8}\{4,5,6,7,8\} covers {7,8}\{7,8\} in MM, but the closure of {1,4,5,6,7,8}\{1,4,5,6,7,8\} in MM^{\prime} is the whole groundset, a flat that does not cover {7,8}\{7,8\}. But then {1,2,3,7,8}\{1,2,3,7,8\} covers {7,8}\{7,8\} in MM and not in MM^{\prime}.

Refer to caption
Figure 9. NN is the restriction of M1M_{1} or M2M_{2} onto {1,2,3,4,5,6}\{1,2,3,4,5,6\}.

This insufficiency leads to the following question:

Question 5.19.

Can it be shown that fΓff\mapsto\Gamma_{f} is a functor from the whole 𝐒𝐌𝐚𝐭𝐫𝐖𝐋\mathbf{SMatrWL} to 𝐒𝐌𝐚𝐭𝐫𝐂𝐨𝐫\mathbf{SMatrCor}?

References

  • [AHK15] Karim Adiprasito, June Huh and Eric Katz “Hodge Theory for Combinatorial Geometries” In Annals of Mathematics 188, 2015 DOI: 10.4007/annals.2018.188.2.1
  • [AR09] Lars Allermann and Johannes Rau “First steps in tropical intersection theory” In Mathematische Zeitschrift 264.3 Springer ScienceBusiness Media LLC, 2009, pp. 633–670 DOI: 10.1007/s00209-009-0483-1
  • [BEST] Andrew Berget, Christopher Eur, Hunter Spink and Dennis Tseng “Tautological classes of matroids” In Inventiones mathematicae 233.2, 2023, pp. 951–1039 DOI: 10.1007/s00222-023-01194-5
  • [BK88] Achim Bachem and Walter Kern “On sticky matroids” In Discrete Mathematics 69.1, 1988, pp. 11–18 DOI: https://doi.org/10.1016/0012-365X(88)90173-2
  • [Bon09] Joseph Bonin “A Note on the Sticky Matroid Conjecture” In Annals of Combinatorics 15, 2009 DOI: 10.1007/s00026-011-0112-7
  • [Cav+16] Renzo Cavalieri, Simon Hampe, Hannah Markwig and Dhruv Ranganathan “Moduli spaces of rational weighted stable curves and tropical geometry”, 2016 arXiv:1404.7426 [math.AG]
  • [Est11] Alexander Esterov “Tropical varieties with polynomial weights and corner loci of piecewise polynomials”, 2011 arXiv:1012.5800 [math.AG]
  • [FH13] Georges Francois and Simon Hampe “Universal Families of Rational Tropical Curves” In Canadian Journal of Mathematics 65.1 Canadian Mathematical Society, 2013, pp. 120–148 DOI: 10.4153/cjm-2011-097-0
  • [Fin13] Alex Fink “Tropical cycles and Chow polytopes” In Beiträge zur Algebra und Geometrie / Contributions to Algebra and Geometry 54.1, 2013, pp. 13–40 DOI: 10.1007/s13366-012-0122-6
  • [FR12] Georges François and Johannes Rau “The diagonal of tropical matroid varieties and cycle intersections” In Collectanea Mathematica 64.2 Springer ScienceBusiness Media LLC, 2012, pp. 185–210 DOI: 10.1007/s13348-012-0072-1
  • [Ful98] William Fulton “Correspondences” In Intersection Theory New York, NY: Springer New York, 1998, pp. 305–318 DOI: 10.1007/978-1-4612-1700-8˙17
  • [MRS19] Lucía López Medrano, Felipe Rincón and Kristin Shaw “Chern–Schwartz–MacPherson cycles of matroids” In Proceedings of the London Mathematical Society 120.1 Wiley, 2019, pp. 1–27 DOI: 10.1112/plms.12278
  • [Oxl11] James Oxley “Matroid Theory” Oxford University Press, 2011 DOI: 10.1093/acprof:oso/9780198566946.001.0001
  • [PT82] Svatopluk Poljak and Daniel Turzik “A note on sticky matroids” In Discrete Mathematics 42.1, 1982, pp. 119–123 DOI: https://doi.org/10.1016/0012-365X(82)90060-7
  • [Sha13] Kristin M. Shaw “A Tropical Intersection Product in Matroidal Fans” In SIAM Journal on Discrete Mathematics 27.1, 2013, pp. 459–491 DOI: 10.1137/110850141
  • [Shi21] Jaeho Shin “The Sticky Matroid Conjecture”, 2021 arXiv:2108.04757 [math.CO]
  • [Sta86] Richard P. Stanley “Partially Ordered Sets” In Enumerative Combinatorics Boston, MA: Springer US, 1986, pp. 96–201 DOI: 10.1007/978-1-4615-9763-6˙3
  • [SW23] Kris Shaw and Annette Werner “On the birational geometry of matroids” In Combinatorial Theory 3.2, 2023 DOI: https://doi.org/10.5070/C63261996