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

Foundations of matroids
Part 1: Matroids without large uniform minors

Matthew Baker Matthew Baker, School of Mathematics, Georgia Institute of Technology, Atlanta, USA [email protected]  and  Oliver Lorscheid Oliver Lorscheid, Instituto Nacional de Matemática Pura e Aplicada, Rio de Janeiro, Brazil [email protected]
Abstract.

The foundation of a matroid is a canonical algebraic invariant which classifies, in a certain precise sense, all representations of the matroid up to rescaling equivalence. Foundations of matroids are pastures, a simultaneous generalization of partial fields and hyperfields which are special cases of both tracts (as defined by the first author and Bowler) and ordered blue fields (as defined by the second author).

Using deep results due to Tutte, Dress–Wenzel, and Gelfand–Rybnikov–Stone, we give a presentation for the foundation of a matroid in terms of generators and relations. The generators are certain “cross-ratios” generalizing the cross-ratio of four points on a projective line, and the relations encode dependencies between cross-ratios in certain low-rank configurations arising in projective geometry.

Although the presentation of the foundation is valid for all matroids, it is simplest to apply in the case of matroids without large uniform minors. i.e., matroids having no minor corresponding to five points on a line or its dual configuration. For such matroids, we obtain a complete classification of all possible foundations.

We then give a number of applications of this classification theorem, for example:

  1. (1)

    We prove the following strengthening of a 1997 theorem of Lee and Scobee: every orientation of a matroid without large uniform minors comes from a dyadic representation, which is unique up to rescaling.

  2. (2)

    For a matroid MM without large uniform minors, we establish the following strengthening of a 2017 theorem of Ardila–Rincón–Williams: if MM is positively oriented then MM is representable over every field with at least 3 elements.

  3. (3)

    Two matroids are said to belong to the same representation class if they are representable over precisely the same pastures. We prove that there are precisely 12 possibilities for the representation class of a matroid without large uniform minors, exactly three of which are not representable over any field.

The first author was supported in part by the NSF Research Grant DMS-1529573. Parts of this project have been carried out while the second author was hosted by the Max Planck Institute in Bonn.

Introduction

Matroids are a combinatorial abstraction of the notion of linear independence in vector spaces. If KK is a field and nn is a positive integer, any linear subspace of KnK^{n} gives rise to a matroid; such matroids are called representable over KK. The task of deciding whether or not certain families of matroids are representable over certain kinds of fields has occupied a plethora of papers in the matroid theory literature.

Dress and Wenzel [13, 14] introduced the Tutte group and the inner Tutte group of a matroid. These are abelian groups which, in a certain precise sense, can be used to understand representations of MM over all so-called fuzzy rings (which, in particular include fields). Dress and Wenzel gave several different presentations for these groups in terms of generators and relations, and Gelfand–Rybnikov–Stone [16] subsequently gave additional presentations for the inner Tutte group of MM. The Dress–Wenzel theory of Tutte groups, inner Tutte groups, and fuzzy rings is powerful but lacks simple definitions and characterizations in terms of universal properties.

In their 1996 paper [28], Semple and Whittle generalized the notion of matroid representations to partial fields (which are special cases of fuzzy rings); this allows one to consider certain families of matroids (e.g. regular or dyadic) as analogous to matroids over a field, and to prove new theorems in the spirit of Tutte’s theorem that a matroid is both binary and ternary if and only if it is regular. Pendavingh and van Zwam [23, 24] subsequently introduced the universal partial field of a matroid MM, which governs the representations of MM over all partial fields. Unfortunately, most matroids (asymptotically 100%, in fact, by a theorem of Nelson [20]) are not representable over any partial field, and in this case the universal partial field gives no information. One can view non-representable matroids as the “dark matter” of matroid theory: they are ubiquitous but somehow mysterious.

Using the theory of matroids over partial hyperstructures presented in [3] (which has been continued in [1], [9] and [22]), we introduced in [5] a generalization of the universal partial field which we call the foundation of a matroid. The foundation is a kind of algebraic object which we call a pasture; pastures include both hyperfields and partial fields and form a natural class of “field-like” objects within the second author’s theory of ordered blueprints in [18]. The category of pastures has various desirable categorical properties (e.g., the existence of products and co-products) which makes it a natural context in which to study algebraic invariants of matroids. Pastures are closely related to fuzzy rings, but they are axiomatically much simpler.

One advantage of the foundation over the universal partial field is that the foundation exists for every matroid MM, not just matroids that are representable over some field. Moreover, unlike the inner Tutte group, the foundation of a matroid is characterized by a universal property which immediately clarifies its importance and establishes its naturality.

More precisely, the foundation of a matroid MM represents the functor taking a pasture FF to the set of rescaling equivalence classes of FF-representations of MM; in particular, MM is representable over a pasture FF if and only if there is a morphism from the foundation of MM to FF.

Our first main result (Theorem 4.20) gives a precise and useful description of the foundation of a matroid in terms of generators and relations. Although this theorem applies to all matroids, it is easiest to apply in the case of matroids without large uniform minors, by which we mean matroids which do not have minors isomorphic to either U52U^{2}_{5} or U53U^{3}_{5}.111Note that if MM has no minor of type U52U^{2}_{5} or U53U^{3}_{5}, then MM also has no uniform minor UnrU^{r}_{n} with n5n\geqslant 5 and 2rn22\leqslant r\leqslant n-2, hence the term “large”. For such matroids, we obtain a complete classification (Theorem 5.9) of all possible foundations, from which one can read off just about any desired representability property. This applies, notably, to the dark matter of matroid theory: we show, for example, that there are precisely three different representation classes of matroids without large uniform minors which are not representable over any field. The applications of Theorem 5.9 which we present in Section 6 are merely a representative sample of the kinds of things one can deduce from this structural result.

We now give a somewhat more precise introduction to the main concepts, definitions, and results in the present paper.

A quick introduction to pastures

A field KK can be thought of as an abelian group G=(K×,,1)G=(K^{\times},\cdot,1), a multiplicatively absorbing element 0, and a binary operation ++ on K=G{0}K=G\cup\{0\} which satisfies certain additional natural axioms (e.g. commutativity, associativity, distributivity, and the existence of additive inverses). Pastures are a generalization of the notion of field in which we still have a multiplicative abelian group GG, an absorbing element 0, and an “additive structure”, but we relax the requirement that the additive structure come from a binary operation. The following two examples are illustrative of the type of relaxations we have in mind.

Example (Krasner hyperfield).

As a pasture, the Krasner hyperfield 𝕂{\mathbb{K}} consists of the multiplicative monoid {0,1}\{0,1\} with 0x=00\cdot x=0 and 11=11\cdot 1=1 and the additive relations 0+x=x0+x=x, 1+1=11+1=1, and 1+1=01+1=0. Note, in particular, that both 1+1=11+1=1 and 1+1=01+1=0 are true, and in particular the additive structure is not derived from a binary operation. The fact that 1+11+1 is equal to two different things may seem counterintuitive at first, but if we think of 1 as a symbol meaning “non-zero”, it is simply a reflection of the fact that the sum of two non-zero elements (in a field, say) can be either non-zero or zero.

Example (Regular partial field).

As a pasture, the regular partial field 𝔽1±{\mathbb{F}}_{1}^{\pm} consists of the multiplicative monoid {0,1,1}\{0,1,-1\} with 0x=00\cdot x=0, 11=11\cdot 1=1, 1(1)=11\cdot(-1)=-1, and (1)(1)=1(-1)\cdot(-1)=1, together with the additive relations 0+x=x0+x=x and 1+(1)=01+(-1)=0. Note, in particular, that there is no additive relation of the form 1+1=x1+1=x or (1)+(1)=x(-1)+(-1)=x, so that once again the additive structure is not derived from a binary operation (but for a different reason: here, 1+11+1 is undefined rather than being multi-valued). We think of 𝔽1±{\mathbb{F}}_{1}^{\pm} as encoding the restriction of addition and multiplication in the ring {\mathbb{Z}} to the multiplicative subset {0,±1}\{0,\pm 1\}.

In general, we will require that a pasture PP has an involution xxx\mapsto-x (which is trivial in the case of 𝕂{\mathbb{K}}), and we can use this involution to rewrite additive relations of the form x+y=zx+y=z as x+yz=0x+y-z=0. It turns out to be more convenient to define pastures using this formalism, and from now on we view the expression x+y=zx+y=z as shorthand for x+y+(z)=0x+y+(-z)=0. For additional notational convenience, we identify relations of the form x+y+z=0x+y+z=0 with triples (x,y,z)(x,y,z); the set of all such triples will be denoted NPN_{P} and called the null set of the pasture.

More formally, a pasture is a multiplicative monoid-with-zero PP such that P×=P\{0}P^{\times}=P\backslash\{0\} is an abelian group, an involution xxx\mapsto-x on PP fixing 0, and a subset NPN_{P} of P3P^{3} such that:

  1. (1)

    (Symmetry) NPN_{P} is invariant under the natural action of S3S_{3} on P3P^{3}.

  2. (2)

    (Weak Distributivity) NPN_{P} is invariant under the diagonal action of P×P^{\times} on P3P^{3}.

  3. (3)

    (Unique Weak Inverses) (0,x,y)NP(0,x,y)\in N_{P} if and only if y=xy=-x.

If we set x[-0]y={zP:x+y=z}x\mathrel{\,\raisebox{-1.1pt}{\larger[-0]{$\boxplus$}}\,}y=\{z\in P\;:\;x+y=z\}, then the pasture PP corresponds to a field if and only if [-0]\mathrel{\,\raisebox{-1.1pt}{\larger[-0]{$\boxplus$}}\,} is an associative binary operation. If x[-0]yx\mathrel{\,\raisebox{-1.1pt}{\larger[-0]{$\boxplus$}}\,}y contains at least one element for all x,yPx,y\in P and [-0]\mathrel{\,\raisebox{-1.1pt}{\larger[-0]{$\boxplus$}}\,} is associative (in the sense of set-wise addition), we call PP a hyperfield. If x[-0]yx\mathrel{\,\raisebox{-1.1pt}{\larger[-0]{$\boxplus$}}\,}y contains at most one element for all x,yPx,y\in P and satisfies a suitable associative law, we call PP a partial field. Pastures generalize (and simplify) both hyperfields and partial fields by imposing no conditions on the size of the sets x[-0]yx\mathrel{\,\raisebox{-1.1pt}{\larger[-0]{$\boxplus$}}\,}y and no associativity conditions.

Example (Hyperfields).

Let KK be a field and let GK×G\leqslant K^{\times} be a multiplicative subgroup. Then the quotient monoid K/G=(K×/G){0}K/G=(K^{\times}/G)\cup\{0\} is naturally a hyperfield: the additive relations are all expressions of the form [x]+[y]=[z][x]+[y]=[z] for which there exist a,b,cGa,b,c\in G such that ax+by=czax+by=cz. For example, /×{\mathbb{R}}/{\mathbb{R}}^{\times} is isomorphic to the Krasner hyperfield 𝕂{\mathbb{K}}, />0{\mathbb{R}}/{\mathbb{R}}_{>0} is isomorphic to the sign hyperfield 𝕊{\mathbb{S}} (cf. [3, Example 2.13]), and if p7p\geqslant 7 is a prime number with p3(mod4)p\equiv 3\pmod{4} then 𝔽p/(𝔽p×)2{\mathbb{F}}_{p}/({\mathbb{F}}_{p}^{\times})^{2} is isomorphic to the weak sign hyperfield 𝕎{\mathbb{W}} (cf. [3, Example 2.13]). However, not every hyperfield arises in this way (cf. [4, 19]).

Example (Partial fields).

Let RR be a commutative ring and let GR×G\leqslant R^{\times} be a subgroup of the unit group of RR containing 1-1. Then P=G{0}P=G\cup\{0\} is naturally a partial field: the additive relations are all expressions of the form x+y=zx+y=z with x,y,zG{0}x,y,z\in G\cup\{0\} such that x+y=zx+y=z in RR. Unlike the situation with hyperfields, every partial field arises from this construction (cf. [24, Theorem 2.16]).

Example (Partial fields, continued).

If RR is a commutative ring, let P(R)P(R) be the partial field corresponding to R×RR^{\times}\subset R. In this paper, we will make extensive use of the following partial fields:

  1. (1)

    𝔽1±=P(){\mathbb{F}}_{1}^{\pm}=P({\mathbb{Z}}). We call this the regular partial field.

  2. (2)

    𝔻=P([12]){\mathbb{D}}=P({\mathbb{Z}}[\frac{1}{2}]). We call this the dyadic partial field.

  3. (3)

    =P([ζ6]){\mathbb{H}}=P({\mathbb{Z}}[\myzeta_{6}]), where ζ6\myzeta_{6}\in{\mathbb{C}} is a primitive sixth root of unity. We call this the hexagonal partial field.222In [24] the partial field {\mathbb{H}} is denoted 𝕊{\mathbb{S}}, but in our context that would conflict with the established terminology for the sign hyperfield, so we re-christen it as {\mathbb{H}}. The partial field 𝕌{\mathbb{U}} is denoted 𝕌1{\mathbb{U}}_{1} in [24].

  4. (4)

    𝕌=P([x,1x,11x]){\mathbb{U}}=P({\mathbb{Z}}[x,\frac{1}{x},\frac{1}{1-x}]), where xx is an indeterminate. We call this the near-regular partial field.

Example (Fields).

It is perhaps worth pointing out explicitly that fields are special cases of both hyperfields and partial fields; in fact, they are precisely the pastures which are both hyperfields and partial fields. Since we will be making extensive use of the finite fields 𝔽2{\mathbb{F}}_{2} and 𝔽3{\mathbb{F}}_{3} in this paper, here is how to explicitly realize these fields as pastures:

  1. (1)

    𝔽2{\mathbb{F}}_{2} has as its underlying monoid {0,1}\{0,1\} with the usual multiplication. The involution xxx\mapsto-x is trivial, and the 33-term additive relations are 0+0+0=00+0+0=0 and 0+1+1=00+1+1=0 (and all permutations thereof).

  2. (2)

    𝔽3{\mathbb{F}}_{3} has as its underlying monoid {0,1,1}\{0,1,-1\} with the usual multiplication. The involution xxx\mapsto-x sends 0 to 0 and 11 to 1-1. The 33-term additive relations are 0+0+0=00+0+0=0, 1+(1)+0=01+(-1)+0=0 (and all permutations thereof), and 1+1+1=01+1+1=0.

A morphism of pastures is a multiplicative map f:PPf:P\to P^{\prime} of monoids such that f(0)=0f(0)=0, f(1)=1f(1)=1 and f(x)+f(y)+f(z)=0f(x)+f(y)+f(z)=0 in PP^{\prime} whenever x+y+z=0x+y+z=0 in PP. Pastures form a category whose initial object is 𝔽1±{\mathbb{F}}_{1}^{\pm} and whose final object is 𝕂{\mathbb{K}}.

Representations of matroids over pastures and the foundation of a matroid

Let MM be a matroid of rank rr on the finite set EE, and let PP be a pasture.

A PP-representation of MM is a function Δ:ErP\Delta:E^{r}\to P such that:

  1. (1)

    Δ(e1,,er)0\Delta(e_{1},\ldots,e_{r})\neq 0 if and only if {e1,,er}\{e_{1},\ldots,e_{r}\} is a basis of MM.

  2. (2)

    Δ(σ(e1),,σ(er))=sign(σ)Δ(e1,,er)\Delta(\mysigma(e_{1}),\ldots,\mysigma(e_{r}))=\operatorname{sign}(\mysigma)\cdot\Delta(e_{1},\ldots,e_{r}) for all permutations σSr\mysigma\in S_{r}.

  3. (3)

    Δ\Delta satisfies the 3-term Plücker relations: for all 𝐉Er2{\mathbf{J}}\in E^{r-2} and all (e1,e2,e3,e4)E4(e_{1},e_{2},e_{3},e_{4})\in E^{4}, the null set NPN_{P} of PP contains the additive relation

    Δ(𝐉e1e2)Δ(𝐉e3e4)Δ(𝐉e1e3)Δ(𝐉e2e4)+Δ(𝐉e1e4)Δ(𝐉e2e3)= 0,\Delta({\mathbf{J}}e_{1}e_{2})\cdot\Delta({\mathbf{J}}e_{3}e_{4})-\Delta({\mathbf{J}}e_{1}e_{3})\cdot\Delta({\mathbf{J}}e_{2}e_{4})+\Delta({\mathbf{J}}e_{1}e_{4})\cdot\Delta({\mathbf{J}}e_{2}e_{3})\ =\ 0,

    where 𝐉eiej:=(j1,,jr2,ei,ej){\mathbf{J}}e_{i}e_{j}:=(j_{1},\dotsc,j_{r-2},e_{i},e_{j}).

Definition.
  1. (1)

    MM is representable over PP if there is at least one PP-representation of MM.

  2. (2)

    Two PP-representations Δ\Delta and Δ\Delta^{\prime} are isomorphic if there exists cP×c\in P^{\times} such that Δ(e1,,er)=cΔ(e1,,er)\Delta^{\prime}(e_{1},\ldots,e_{r})=c\Delta(e_{1},\ldots,e_{r}) for all (e1,,er)Er(e_{1},\ldots,e_{r})\in E^{r}.333An isomorphism class of PP-representations of MM is the same thing as a weak PP-matroid whose support is MM, in the terminology of [3].

  3. (3)

    Δ\Delta and Δ\Delta^{\prime} rescaling equivalent if there exist cP×c\in P^{\times} and a map d:EP×d:E\to P^{\times} such that Δ(e1,,er)=cd(e1)d(er)Δ(e1,,er)\Delta^{\prime}(e_{1},\ldots,e_{r})=c\cdot d(e_{1})\cdots d(e_{r})\cdot\Delta(e_{1},\ldots,e_{r}) for all (e1,,er)Er(e_{1},\ldots,e_{r})\in E^{r}.

  4. (4)

    We denote by 𝒳MI(P){\mathcal{X}}^{I}_{M}(P) (resp. 𝒳MR(P){\mathcal{X}}^{R}_{M}(P)) the set of isomorphism classes (resp. rescaling classes) of PP-representations of MM.444In [5], these sets are denoted 𝒳Mw(P){\mathcal{X}}^{w}_{M}(P) and 𝒳Mf(P){\mathcal{X}}^{f}_{M}(P), respectively.

Example.

By the results in [3] and [5], we have:

  1. (1)

    If KK is a field, the isomorphism classes of KK-representations of MM are naturally in bijection with rr-dimensional subspaces of KEK^{E} (the KK-vector space of functions from EE to KK) whose underlying matroid is MM.

  2. (2)

    Every matroid has a unique representation over the Krasner hyperfield 𝕂{\mathbb{K}}.

  3. (3)

    If PP is a partial field, MM is representable over PP if and only if it is representable by a PP-matrix in the sense of [24]. In particular, a matroid is regular (i.e., representable over {\mathbb{Z}} by a totally unimodular matrix) if and only if it is representable over the partial field 𝔽1±{\mathbb{F}}_{1}^{\pm}. A regular matroid will in general have many different (non-isomorphic) representations over 𝔽1±{\mathbb{F}}_{1}^{\pm}, but there is a unique rescaling class of such representations.

  4. (4)

    A matroid is orientable if and only if it is representable over the sign hyperfield 𝕊{\mathbb{S}}. An orientation of MM is the same thing as an 𝕊{\mathbb{S}}-representation, and in this case rescaling equivalence is usually called reorientation equivalence.

For fixed MM the map taking a pasture PP to the set 𝒳MI(P){\mathcal{X}}^{I}_{M}(P) (resp. 𝒳MR(P){\mathcal{X}}^{R}_{M}(P)) is a functor. In particular, if f:P1P2f:P_{1}\to P_{2} is a morphism of pastures, there are natural maps 𝒳MI(P1)𝒳MI(P2){\mathcal{X}}^{I}_{M}(P_{1})\to{\mathcal{X}}^{I}_{M}(P_{2}) and 𝒳MR(P1)𝒳MR(P2){\mathcal{X}}^{R}_{M}(P_{1})\to{\mathcal{X}}^{R}_{M}(P_{2}).

We now come to the key result from [5] motivating the present paper:

Theorem.

Given a matroid MM, the functor taking a pasture PP to the set 𝒳MI(P){\mathcal{X}}^{I}_{M}(P) is representable by a pasture PMP_{M} which we call the universal pasture of MM. In other words, we have a natural isomorphism

(1) Hom(PM,)𝒳MI.\operatorname{Hom}(P_{M},-)\simeq{\mathcal{X}}^{I}_{M}.

The functor taking a pasture PP to the set 𝒳MR(P){\mathcal{X}}^{R}_{M}(P) is representable by a subpasture FMF_{M} of PMP_{M} which we call the foundation of MM, i.e. there is a natural isomorphism

(2) Hom(FM,)𝒳MR.\operatorname{Hom}(F_{M},-)\simeq{\mathcal{X}}^{R}_{M}.

For various reasons, including the fact that the foundation can be presented by generators and relations “induced from small minors”, we will mainly focus in this paper on studying the foundation of MM rather than the universal pasture. Note that both PMP_{M} and FMF_{M} have the property that MM is representable over a pasture PP if and only if there is a morphism from PMP_{M} (resp. FMF_{M}) to PP.

Remark.
  1. (1)

    The universal partial field and foundation behave nicely with respect to various matroid operations. For example, the universal partial fields (resp. foundations) of MM and its dual matroid MM^{*} are canonically isomorphic. And there is a natural morphism from the universal partial field (resp. foundation) of a minor N=M\I/JN=M\backslash I/J of MM to the universal partial field (resp. foundation) of MM.

  2. (2)

    The multiplicative group PM×P_{M}^{\times} (resp. FM×F_{M}^{\times}) of the universal partial field (resp. foundation) of MM is isomorphic to the Tutte group (resp. inner Tutte group) of Dress and Wenzel [13, Definition 1.6].

If we take P=PMP=P_{M} in (1), the identity map is a distinguished element of Hom(PM,PM)\operatorname{Hom}(P_{M},P_{M}). It therefore corresponds to a distinguished element Δ^M𝒳MI(PM)\hat{\Delta}_{M}\in{\mathcal{X}}^{I}_{M}(P_{M}), which (by abuse of terminology) we call the universal representation of MM. (Technically speaking, the universal representation is actually an isomorphism class of representations.)

Remark.

When FMF_{M} is a partial field, the foundation coincides with the universal partial field of [23]. However, when MM is not representable over any field, the universal partial field does not exist. On the other hand, the foundation of MM is always well-defined; this is one sense in which the theory of pastures helps us explore the “dark matter” of the matroid universe.

Products and coproducts

The category of pastures admits finite products and co-products (a.k.a. tensor products). This is a key advantage of pastures over the categories of fields, partial fields, and hyperfields, none of which admit both products and co-products. The relevance of such considerations to matroid theory is illustrated by the following observations:

  1. (1)

    MM is representable over both P1P_{1} and P2P_{2} if and only if MM is representable over the product pasture P1×P2P_{1}\times P_{2}. (This is immediate from the universal property of the foundation and of categorical products.)

  2. (2)

    If M1M_{1} and M2M_{2} are matroids, the foundation of the direct sum M1M2M_{1}\oplus M_{2} is canonically isomorphic to the tensor product FM1FM2F_{M_{1}}\otimes F_{M_{2}}, and similarly for the 22-sum of M1M_{1} and M2M_{2}. (These facts, along with some applications, will be discussed in detail a follow-up paper.)

  3. (3)

    Tensor products of pastures are needed in order to state and apply the main theorem of this paper, the classification theorem for foundations of matroids without large uniform minors (Theorem 5.9 below).

In order to illustrate the utility of categorical considerations for studying matroid representations, we briefly discuss a couple of key examples.

Example.

The product of the fields 𝔽2{\mathbb{F}}_{2} and 𝔽3{\mathbb{F}}_{3}, considered as pastures, is isomorphic to the regular partial field 𝔽1±{\mathbb{F}}_{1}^{\pm}. As an immediate consequence, we obtain Tutte’s celebrated result that a matroid MM is representable over every field if and only if MM is regular. (Proof: If MM is regular then since 𝔽1±{\mathbb{F}}_{1}^{\pm} is an initial object in the category of pastures, MM is representable over every pasture, and in particular over every field. Conversely, if MM is representable over every field, then it is in particular representable over both 𝔽2{\mathbb{F}}_{2} and 𝔽3{\mathbb{F}}_{3}, hence over their product 𝔽1±{\mathbb{F}}_{1}^{\pm}, and thus MM is regular.)

One can, in the same way, establish Whittle’s theorem that a matroid is representable over both 𝔽3{\mathbb{F}}_{3} and 𝔽4{\mathbb{F}}_{4} if and only if it is hexagonal, i.e., representable over the partial field {\mathbb{H}}.

These kind of arguments are well-known in the theory of partial fields; however, the theory of pastures is more flexible. For example, the product of the field 𝔽2{\mathbb{F}}_{2} and the hyperfield 𝕊{\mathbb{S}} is also isomorphic to the partial field 𝔽1±{\mathbb{F}}_{1}^{\pm}. In this way, we obtain a unified proof of the result of Tutte just mentioned and the theorem of Bland and Las Vergnas that a matroid is regular if and only if it is both binary and orientable [8].

Example.

If we try to extend this type of argument to more general pastures, we run into some intriguing complications. As an illuminating example, consider the theorem of Lee and Scobee [17] that a matroid is both ternary and orientable if and only if it is dyadic, i.e., representable over the partial field 𝔻{\mathbb{D}}. In this case, the product of 𝔽3{\mathbb{F}}_{3} and 𝕊{\mathbb{S}} is not isomorphic to 𝔻{\mathbb{D}}; there is merely a morphism from 𝔻{\mathbb{D}} to 𝔽3×𝕊{\mathbb{F}}_{3}\times{\mathbb{S}}. The theorem of Lee and Scobee therefore lies deeper than the theorems mentioned in the previous example; proving it requires establishing, in particular, that 𝔽3×𝕊{\mathbb{F}}_{3}\times{\mathbb{S}} is not the foundation of any matroid.

To do this, one needs a structural understanding of foundations, which we obtain by utilizing highly non-trivial results of Tutte, Dress–Wenzel, and Gelfand–Rybnikov–Stone. The result of our analysis, in the context of matroids which are both ternary and orientable, is that every morphism from the foundation of some matroid to 𝔽3×𝕊{\mathbb{F}}_{3}\times{\mathbb{S}} lifts uniquely to 𝔻{\mathbb{D}}. More precisely, we prove that if MM is a matroid without large uniform minors (e.g. if MM is ternary), then the morphism 𝔻𝕊{\mathbb{D}}\to{\mathbb{S}} induces a canonical bijection Hom(FM,𝔻)Hom(FM,𝕊)\operatorname{Hom}(F_{M},{\mathbb{D}})\to\operatorname{Hom}(F_{M},{\mathbb{S}}). This gives a new and non-trivial strengthening of the Lee–Scobee theorem. The proof goes roughly as follows: by Theorem B we have FMF1FsF_{M}\cong F_{1}\otimes\cdots\otimes F_{s}, where each FiF_{i} belongs to an explicit finite set 𝒫{\mathcal{P}} of pastures. By categorical considerations, the statement that a morphism f:FM𝕊f:F_{M}\to{\mathbb{S}} lifts uniquely to 𝔻{\mathbb{D}} is equivalent to the statement that Hom(P,𝕊)=Hom(P,𝔻)\operatorname{Hom}(P,{\mathbb{S}})=\operatorname{Hom}(P,{\mathbb{D}}) for all P𝒫P\in{\mathcal{P}}, and this can be checked by concrete elementary computations.

Universal cross ratios

In order to explain why the “large” uniform minors U52U^{2}_{5} and U53U^{3}_{5} play a special role in the theory of foundations, we need to first explain the concept of a universal cross ratio, which is intimately related to U42U^{2}_{4}-minors.

Let MM be a matroid of rank rr, let PP be a pasture, and let Δ\Delta be a PP-representation of MM. Let 𝐉Er2{\mathbf{J}}\in E^{r-2} have distinct coordinates and let JJ be the corresponding unordered subset of EE of size r2r-2. If Δ(𝐉e1e4)\Delta({\mathbf{J}}e_{1}e_{4}) and Δ(𝐉e2e3)\Delta({\mathbf{J}}e_{2}e_{3}) are both non-zero (i.e., if J{e1,e4}J\cup\{e_{1},e_{4}\} and J{e2,e3}J\cup\{e_{2},e_{3}\} are both bases of MM), then we can rewrite the 3-term Plücker relation

Δ(𝐉e1e2)Δ(𝐉e3e4)Δ(𝐉e1e3)Δ(𝐉e2e4)+Δ(𝐉e1e4)Δ(𝐉e2e3)=0\Delta({\mathbf{J}}e_{1}e_{2})\Delta({\mathbf{J}}e_{3}e_{4})-\Delta({\mathbf{J}}e_{1}e_{3})\Delta({\mathbf{J}}e_{2}e_{4})+\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{2}e_{3})=0

as

Δ(𝐉e1e3)Δ(𝐉e2e4)Δ(𝐉e1e4)Δ(𝐉e2e3)+Δ(𝐉e1e2)Δ(𝐉e3e4)Δ(𝐉e1e4)Δ(𝐉e3e2)=1.\frac{\Delta({\mathbf{J}}e_{1}e_{3})\Delta({\mathbf{J}}e_{2}e_{4})}{\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{2}e_{3})}+\frac{\Delta({\mathbf{J}}e_{1}e_{2})\Delta({\mathbf{J}}e_{3}e_{4})}{\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{3}e_{2})}=1.

Moreover, as one easily checks, the quantities Δ(𝐉e1e3)Δ(𝐉e2e4)Δ(𝐉e1e4)Δ(𝐉e2e3)\frac{\Delta({\mathbf{J}}e_{1}e_{3})\Delta({\mathbf{J}}e_{2}e_{4})}{\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{2}e_{3})} and Δ(𝐉e1e2)Δ(𝐉e3e4)Δ(𝐉e1e4)Δ(𝐉e3e2)\frac{\Delta({\mathbf{J}}e_{1}e_{2})\Delta({\mathbf{J}}e_{3}e_{4})}{\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{3}e_{2})} are invariant under rescaling equivalence and do not depend on the choice of ordering of elements of JJ. In particular,

[e1e2e3e4]Δ,𝐉:=Δ(𝐉e1e3)Δ(𝐉e2e4)Δ(𝐉e1e4)Δ(𝐉e2e3)\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\Delta,{\mathbf{J}}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\Delta,{\mathbf{J}}}}{}{}:=\frac{\Delta({\mathbf{J}}e_{1}e_{3})\Delta({\mathbf{J}}e_{2}e_{4})}{\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{2}e_{3})}

depends only on JJ and on the rescaling class [Δ][\Delta] of Δ\Delta in 𝒳MR(P){\mathcal{X}}^{R}_{M}(P).

The cross ratio associated to the universal representation Δ^M:ErPM\hat{\Delta}_{M}:E^{r}\to P_{M} plays an especially important role in our theory. For notational convenience, we set

[e1e2e3e4]M,J:=[e1e2e3e4]Δ^M,𝐉.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}}{}{}:=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\hat{\Delta}_{M},{\mathbf{J}}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\hat{\Delta}_{M},{\mathbf{J}}}}{}{}.

We will write [e1e2e3e4]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{} instead of [e1e2e3e4]M,J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}}{}{} when MM is understood.

Using the fact that [e1e2e3e4]Δ^M,J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\hat{\Delta}_{M},J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\hat{\Delta}_{M},J}}{}{} depends only on the rescaling class of Δ^M\widehat{\Delta}_{M}, one sees easily that [e1e2e3e4]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}, which a priori is an element of the universal pasture PMP_{M}, in fact belongs to the foundation FMF_{M}.

We call elements of FMF_{M} of the form [e1e2e3e4]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{} universal cross ratios of MM. When J=J=\varnothing we omit the subscript entirely. By [5, Lemma 7.7], we have:

Lemma.

The foundation of MM is generated by its universal cross ratios.

Remark.
  1. (1)

    When J=J=\varnothing and M=U42M=U^{2}_{4} is the uniform matroid of rank 2 on the 4-element set {1,2,3,4}\{1,2,3,4\}, the quantity [1234]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{} can be viewed as a “universal” version of the usual cross-ratio of four points on a projective line. The fact that the cross-ratio is the only projective invariant of four points on a line corresponds to the fact that the foundation of U42U^{2}_{4} is isomorphic to the partial field 𝕌=P([x,1x,11x]){\mathbb{U}}=P({\mathbb{Z}}[x,\frac{1}{x},\frac{1}{1-x}]) described above. The six different values of [σ(1)σ(2)σ(3)σ(4)]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$\mysigma(1)$}}&{\scalebox{0.9}{$\mysigma(2)$}}\\[-2.0pt] {\scalebox{0.9}{$\mysigma(3)$}}&{{\scalebox{0.9}{$\mysigma(4)$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{\mysigma(1)}&{\mysigma(2)}\\ {\mysigma(3)}&{\mysigma(4)}\end{smallmatrix}\big{]}}{}{} for σS4\mysigma\in S_{4} correspond to the elements x,1x,1x,11x,11xx,1-x,\frac{1}{x},1-\frac{1}{x},\frac{1}{1-x}, and 111x1-\frac{1}{1-x} of 𝕌{\mathbb{U}}.

  2. (2)

    More generally, we can associate a universal cross ratio to each U42U^{2}_{4}-minor N=M\I/JN=M\backslash I/J of MM (together with an ordering of the ground set of NN) via the natural map from FNF_{N} to FMF_{M}, and every universal cross ratio arises from this construction.

The structure theorem for foundations of matroids without large uniform minors

In order to calculate and classify foundations of matroids, in addition to knowing that the universal cross ratios generate FMF_{M}, we need to understand the relations between these generators.

Example.

The universal cross ratios of the uniform matroid U52U^{2}_{5} on {1,2,3,4,5}\{1,2,3,4,5\} satisfy certain tip relations of the form

[1234][1245][1253]= 1.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {4}&{5}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$5$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {5}&{3}\end{smallmatrix}\big{]}}{}{}\ =\ 1.

By duality, the universal cross ratios of U53U^{3}_{5} satisfy similar identities which we call the cotip relations.

The theoretical tool which allows one to understand all relations between universal cross ratios is Tutte’s Homotopy Theorem [31, 32, 33] (or, more specifically, [16, Theorem 4], whose proof is based on Tutte’s Homotopy Theorem). We give an informal description here; a more precise version is given in Theorem 4.20 below. To state the result, we say that a relation between universal cross-ratios of MM is inherited from a minor N=M\I/JN=M\backslash I/J if it is the image (with respect to the natural inclusion FNFMF_{N}\subseteq F_{M}) of a relation between universal cross ratios in FNF_{N}.

Theorem A.

Every relation between universal cross ratios of a matroid MM is inherited from a minor on a 66-element set. The foundation of MM is generated as an 𝔽1±{\mathbb{F}}_{1}^{\pm}-algebra by such generators and relations, together with the relation 1=1-1=1 if MM has a minor isomorphic to either the Fano matroid F7F_{7} or its dual.

The most complicated relations between universal cross ratios come from the tip and cotip relations in U52U^{2}_{5} and U53U^{3}_{5}, respectively (six-element minors and non-uniform five-element minors only contribute additional relations identifying certain cross ratios with one another). In the absence of such minors, we can completely classify all possible foundations. Roughly speaking, the conclusion is that the foundation of a matroid is the tensor product of copies of 𝔽2{\mathbb{F}}_{2} and quotients of 𝕌{\mathbb{U}} (the foundation of U42U^{2}_{4}) by groups of automorphisms. By calculating all possible quotients of 𝕌{\mathbb{U}} by automorphisms, we obtain the following result (Theorem 5.9):

Theorem B.

Let MM be a matroid without large uniform minors and FMF_{M} its foundation. Then

FMF1FrF_{M}\ \simeq\ F_{1}\otimes\dotsb\otimes F_{r}

for some r0r\geqslant 0 and pastures F1,,Fr{𝕌,𝔻,,𝔽3,𝔽2}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}},{\mathbb{F}}_{3},{\mathbb{F}}_{2}\}.

Remark.

In a sequel paper, we will show that every pasture of the form F1FrF_{1}\otimes\dotsb\otimes F_{r} with F1,,Fr{𝕌,𝔻,,𝔽3,𝔽2}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}},{\mathbb{F}}_{3},{\mathbb{F}}_{2}\} is the foundation of some matroid.

Consequences of the structure theorem

A matroid MM is representable over a pasture PP if and only if there is a morphism from the foundation FMF_{M} of MM to PP. If MM is without large uniform minors (which is automatic if MM is binary or ternary), then by Theorem 5.9 its foundation is isomorphic to a tensor product of copies FiF_{i} of 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}}, 𝔽3{\mathbb{F}}_{3} and 𝔽2{\mathbb{F}}_{2}. There is a morphism from FMF_{M} to PP if and only if there is a morphism from each FiF_{i} to PP, so one readily obtains various theorems about representability of such matroids.

We mention just a selection of sample applications from the more complete list of results in section 6. For instance, our method yields short proofs of the excluded minor characterizations of regular, binary and ternary matroids (Theorems 6.3 and 6.4). We find a similarly short proof for Brylawski-Lucas’s result that every matroid has at most one rescaling class over 𝔽3{\mathbb{F}}_{3} (Theorem 6.5 and Remark 6.6).

As already mentioned, we derive a strengthening of a theorem by Lee and Scobee ([17]) on lifts of oriented matroids. The lifting result assumes a particularly strong form in the case of positively oriented matroids, improving on a result by Ardila, Rincón and Williams ([2]). The following summarizes Theorems 6.9 and 6.15:

Theorem C.

Let MM be an oriented matroid whose underlying matroid is without large uniform minors. Then MM is uniquely dyadic up to rescaling. If MM is positively oriented, then MM is near-regular.

In Theorem 6.7, we derive similar statements for the weak hyperfield of signs 𝕎{\mathbb{W}} and the phase hyperfield {\mathbb{P}}; cf. section 2.1.2 for definitions. Namely, a matroid MM without large uniform minors is ternary if it is representable over 𝕎{\mathbb{W}}, and is representable over 𝔻{\mathbb{D}}\otimes{\mathbb{H}} if it is representable over {\mathbb{P}}.

We define the representation class of a matroid MM as the class 𝒫M{\mathcal{P}}_{M} of all pastures PP over which MM is representable. Two matroids MM and MM^{\prime} are representation equivalent if 𝒫M=𝒫M{\mathcal{P}}_{M}={\mathcal{P}}_{M^{\prime}}. The following is Theorem 6.20.

Theorem D.

Let MM be a matroid without large uniform minors. Then there are precisely 12 possibilities for the representation class of MM. Nine of these classes are representable over some field, and the other three are not.

The structure theorem also provides short proofs of various characterizations (some new, some previously known by other methods) of certain classes of matroids. The following summarizes Theorems 6.266.34:

Theorem E.

Let MM be a matroid without large uniform minors and FMF_{M} its foundation. Then all conditions in a given row in Table 1 are equivalent, where the conditions should be read as follows:

  1. (1)

    The first column describes the class by name (cf. Definition 2.14 for any unfamiliar terms).

  2. (2)

    The second column characterizes the class in terms of the factors FiF_{i} that may appear in a decomposition FMFiF_{M}\simeq\bigotimes F_{i}, as in Theorem B.

  3. (3)

    The third column lists various classifying pastures PP, separated by slashes, which means that MM is contained in the class in question if and only if it is representable over PP.

Table 1. Characterizations of classes of matroids without large uniform minors
class possible factors of FMF_{M} representable over
regular 𝕌{\mathbb{U}} /\Big{/} 𝔽2×P{\mathbb{F}}_{2}\times P with 11-1\neq 1 in PP
binary 𝔽2{\mathbb{F}}_{2} 𝔽2{\mathbb{F}}_{2}
ternary 𝕌,𝔻,,𝔽3{\mathbb{U}},{\mathbb{D}},{\mathbb{H}},{\mathbb{F}}_{3} any field extension kk of 𝔽3{\mathbb{F}}_{3} /\Big{/} 𝕎{\mathbb{W}}
quaternary 𝕌,,𝔽2{\mathbb{U}},{\mathbb{H}},{\mathbb{F}}_{2} any field extension kk of 𝔽4{\mathbb{F}}_{4}
near-regular 𝕌{\mathbb{U}} 𝕌/𝔽3×𝔽8/𝔽4×𝔽5/𝔽4×𝕊/𝔽8×𝕎/𝔻×\begin{array}[]{c}{\mathbb{U}}\ \Big{/}\ {\mathbb{F}}_{3}\times{\mathbb{F}}_{8}\ \Big{/}\ {\mathbb{F}}_{4}\times{\mathbb{F}}_{5}\ \Big{/}\ {\mathbb{F}}_{4}\times{\mathbb{S}}\ \Big{/}\\ {\mathbb{F}}_{8}\times{\mathbb{W}}\ \Big{/}\ {\mathbb{D}}\times{\mathbb{H}}\end{array}
dyadic 𝕌,𝔻{\mathbb{U}},{\mathbb{D}} 𝔻/𝔽3×/𝔽3×𝕊/𝔽3×𝔽q with 2q and 3q1\begin{array}[]{c}{\mathbb{D}}\ \Big{/}\ {\mathbb{F}}_{3}\times{\mathbb{Q}}\ \Big{/}\ {\mathbb{F}}_{3}\times{\mathbb{S}}\ \Big{/}\\ {\mathbb{F}}_{3}\times{\mathbb{F}}_{q}\text{ with }2\nmid q\text{ and }3\nmid q-1\end{array}
hexagonal 𝕌,{\mathbb{U}},{\mathbb{H}} {\mathbb{H}} /\Big{/} 𝔽3×𝔽4{\mathbb{F}}_{3}\times{\mathbb{F}}_{4} /\Big{/} 𝔽4×𝕎{\mathbb{F}}_{4}\times{\mathbb{W}}
𝔻{\mathbb{D}}\otimes{\mathbb{H}}-representable 𝕌,𝔻,{\mathbb{U}},{\mathbb{D}},{\mathbb{H}} 𝔽3×/𝔽3×/𝔽3×𝔽q with 2q and 3q1\begin{array}[]{c}{\mathbb{F}}_{3}\times{\mathbb{C}}\ \Big{/}\ {\mathbb{F}}_{3}\times{\mathbb{P}}\ \Big{/}\\ {\mathbb{F}}_{3}\times{\mathbb{F}}_{q}\text{ with }2\nmid q\text{ and }3\mid q-1\end{array}
representable 𝕌,𝔻,,𝔽3or 𝕌,,𝔽2\begin{array}[]{c}{\mathbb{U}},{\mathbb{D}},{\mathbb{H}},{\mathbb{F}}_{3}\\ \text{or }{\mathbb{U}},{\mathbb{H}},{\mathbb{F}}_{2}\end{array} either 𝔽3{\mathbb{F}}_{3} or 𝔽4{\mathbb{F}}_{4}

Another consequence of the structure theorem for foundations of matroids without large uniform minors is the following result, which will be the theme of a forthcoming paper.

Theorem F.

Let MM be a ternary matroid. Then up to rescaling equivalence,

  1. (1)

    every quarternary representation of MM lifts uniquely to {\mathbb{H}};

  2. (2)

    every quinternary representation of MM lifts uniquely to 𝔻{\mathbb{D}};

  3. (3)

    every septernary representation of MM lifts uniquely to 𝔻{\mathbb{D}}\otimes{\mathbb{H}};

  4. (4)

    every octernary representation of MM lifts uniquely to 𝕌{\mathbb{U}}.

Content overview

In section 1, we introduce embedded minors and review basic facts concerning the Tutte group of a matroid. In section 2, we discuss matroid representations over pastures and explain the concept of the universal pasture of a matroid. In section 3, we extend the concept of cross ratios to matroid representations over pastures and define universal cross ratios. In section 4, we introduce the foundation of a matroid and exhibit a complete set of relations between cross ratios, which culminates in Theorem A. In section 5, we focus on matroids without large uniform minors and prove Theorem B. In section 6, we explain several consequences of Theorem B, such as Theorems C, D and E.

Acknowledgements

The authors thank Nathan Bowler and Rudi Pendavingh for helpful discussions; in particular, we thank Rudi Pendavingh for suggesting that a result along the lines of Theorem 4.20 should follow from [16]. The authors also thank their respective muses Camille and Cecília for inspiring the name of the matroid C5C_{5}.

1. Background

1.1. Notation

In this paper, we assume that the reader is familiar with basic concepts from matroid theory.

Typically, MM denotes a matroid of rank rr on the ground set E={1,,n}E=\{1,\dotsc,n\}. We denote its set of bases by =M{\mathcal{B}}={\mathcal{B}}_{M} and its set of hyperplanes by =M{\mathcal{H}}={\mathcal{H}}_{M}. We denote the closure of a subset JJ of EE by J\langle J\rangle. We denote the dual matroid of MM by MM^{\ast}.

Given two subsets II and JJ of EE, we denote by IJ={iIiJ}I-J=\{i\in I\mid i\notin J\} the complement of JJ in II. For an ordered tuple 𝐉=(j1,,js){\mathbf{J}}=(j_{1},\dotsc,j_{s}) in EsE^{s}, we denote by |𝐉||{\mathbf{J}}| the subset {j1,,js}\{j_{1},\dotsc,j_{s}\} of EE. Given kk elements e1,,ekEe_{1},\dotsc,e_{k}\in E, we denote by 𝐉e1ek{\mathbf{J}}e_{1}\dotsb e_{k} the s+ks+k-tuple (j1,,js,e1,,ek)Es+k(j_{1},\dotsc,j_{s},e_{1},\dotsc,e_{k})\in E^{s+k}. If JJ is a subset of EE, then we denote by Je1ekJe_{1}\dotsb e_{k} the subset J{e1,,ek}J\cup\{e_{1},\dotsc,e_{k}\} of EE. In particular, we have |𝐉e1ek|=|𝐉|e1ek|{\mathbf{J}}e_{1}\dotsb e_{k}|=|{\mathbf{J}}|e_{1}\dotsb e_{k} for 𝐉Es{\mathbf{J}}\in E^{s}. inline]Reminder: we should point out discrepancies to the notation and terminology of [5].

1.2. The Tutte group

The Tutte group is an invariant of a matroid that was introduced and studied by Dress and Wenzel in [13]. We will review the parts of this theory that are relevant for the present text in the following.

Definition 1.1.

Let MM be a matroid of rank rr on EE with Grassmann-Plücker function Δ:Er𝕂\Delta:E^{r}\to{\mathbb{K}}. The multiplicatively written abelian group 𝕋M{\mathbb{T}}_{M}^{\mathcal{B}} is generated by symbols 1-1 and X𝐈X_{\mathbf{I}} for every 𝐈supp(Δ){\mathbf{I}}\in\operatorname{supp}(\Delta) modulo the relations

(1)2= 1;(-1)^{2}\ =\ 1;
X(eσ(1),,eσ(r))=sign(σ)X(e1,,er)X_{(e_{\mysigma(1)},\dotsc,e_{\mysigma(r)})}=\operatorname{sign}(\mysigma)X_{(e_{1},\dotsc,e_{r})}

for every permutation σSr\mysigma\in S_{r}, where we consider sign(σ)\operatorname{sign}(\mysigma) as an element of {±1}𝕋M\{\pm 1\}\subset{\mathbb{T}}_{M}^{\mathcal{B}};

X𝐉e1e3X𝐉e2e4=X𝐉e1e4X𝐉e2e3X_{{\mathbf{J}}e_{1}e_{3}}X_{{\mathbf{J}}e_{2}e_{4}}=X_{{\mathbf{J}}e_{1}e_{4}}X_{{\mathbf{J}}e_{2}e_{3}}

for 𝐉=(j1,,jr2)Er2{\mathbf{J}}=(j_{1},\dotsc,j_{r-2})\in E^{r-2} and e1,,e4Ee_{1},\dotsc,e_{4}\in E such that 𝐉eiejsupp(Δ){\mathbf{J}}e_{i}e_{j}\in\operatorname{supp}(\Delta) for all i=1,2i=1,2 and j=3,4j=3,4 but Δ(𝐉e1e2)Δ(𝐉e3e4)=0\Delta({\mathbf{J}}e_{1}e_{2})\Delta({\mathbf{J}}e_{3}e_{4})=0.

The group 𝕋M{\mathbb{T}}_{M}^{\mathcal{B}} comes with a morphism deg:𝕋M\deg:{\mathbb{T}}_{M}^{\mathcal{B}}\to{\mathbb{Z}} that sends X𝐈X_{\mathbf{I}} to 11 for every 𝐈supp(Δ){\mathbf{I}}\in\operatorname{supp}(\Delta). The Tutte group of MM is the kernel 𝕋M=ker(deg){\mathbb{T}}_{M}=\ker(\deg) of this map.

By definition, the Tutte group 𝕋M{\mathbb{T}}_{M} is generated by ratios X𝐈/X𝐉X_{\mathbf{I}}/X_{\mathbf{J}} of generators of X𝐈X_{\mathbf{I}}, X𝐉X_{\mathbf{J}} of 𝕋M{\mathbb{T}}_{M}^{\mathcal{B}}. Since the basis exchange graph of a matroid is connected, it follows that 𝕋M{\mathbb{T}}_{M} is generated by elements of the form X𝐉e/X𝐉eX_{{\mathbf{J}}e}/X_{{\mathbf{J}}e^{\prime}}, where 𝐉Er1{\mathbf{J}}\in E^{r-1} and e,eEe,e^{\prime}\in E are such that both 𝐉e{\mathbf{J}}e and 𝐉e{\mathbf{J}}e^{\prime} are in the support of Δ\Delta.

The Tutte group can equivalently be defined in terms of hyperplanes, as explained in the following.

Definition 1.2.

Let MM be a matroid and {\mathcal{H}} its set of hyperplanes. We define 𝕋M{\mathbb{T}}_{M}^{\mathcal{H}} as the abelian group generated by symbols 1-1 and XH,eX_{H,e} for all HH\in{\mathcal{H}} and eEHe\in E-H modulo the relations

(1)2= 1;(-1)^{2}\ =\ 1;
XH1,e2XH2,e3XH3,e1XH1,e3XH2,e1XH3,e2=1,\frac{X_{H_{1},e_{2}}X_{H_{2},e_{3}}X_{H_{3},e_{1}}}{X_{H_{1},e_{3}}X_{H_{2},e_{1}}X_{H_{3},e_{2}}}\ =\ -1,

where H1,H2,H3H_{1},H_{2},H_{3}\in{\mathcal{H}} are pairwise distinct such that F=H1H2H3F=H_{1}\cap H_{2}\cap H_{3} is a flat of rank r2r-2 and eiHiFe_{i}\in H_{i}-F for i=1,2,3i=1,2,3.

This group comes with a map deg:𝕋M\deg_{\mathcal{H}}:{\mathbb{T}}_{M}^{\mathcal{H}}\to{\mathbb{Z}}^{\mathcal{H}} that sends an element XH,eX_{H,e} to the characteristic function χH:\mychi_{H}:{\mathcal{H}}\to{\mathbb{Z}} of {H}\{H\}\subset{\mathcal{H}}, i.e. χH(H)=δH,H\mychi_{H}(H^{\prime})=\mydelta_{H,H^{\prime}} for HH^{\prime}\in{\mathcal{H}}.

The relation between 𝕋M{\mathbb{T}}_{M} and 𝕋M{\mathbb{T}}_{M}^{\mathcal{H}} is explained in [13, Thms. 1.1 and 1.2], which is as follows.

Theorem 1.3 (Dress-Wenzel ’89).

Let MM be a matroid and {\mathcal{B}} its set of bases. Then the association 11-1\mapsto-1 and X𝐉e/X𝐉eXH,e/XH,eX_{{\mathbf{J}}e}/X_{{\mathbf{J}}e^{\prime}}\mapsto X_{H,e}/X_{H,e^{\prime}}, where 𝐉Er1{\mathbf{J}}\in E^{r-1}, e,eEe,e^{\prime}\in E with |𝐉e|,|𝐉e||{\mathbf{J}}e|,|{\mathbf{J}}e^{\prime}|\in{\mathcal{B}} and H=|𝐉|H=\langle|{\mathbf{J}}|\rangle, defines an injective group homomorphism 𝕋M𝕋M{\mathbb{T}}_{M}\to{\mathbb{T}}_{M}^{\mathcal{H}} whose image is ker(deg)\ker(\deg_{\mathcal{H}}).

1.3. Embedded minors

In this section, we review some basic facts about minors of a matroid and introduce the concept of an embedded minor.

Let MM and NN be matroids with respective ground sets EME_{M} and ENE_{N}. An isomorphism φ:NM\myvarphi:N\to M of matroids is a bijection φ:ENEM\myvarphi:E_{N}\to E_{M} such that BENB\subset E_{N} is a basis of NN if and only if φ(B)\myvarphi(B) is a basis of MM.

Definition 1.4.

Let MM be a matroid on EE. A minor of MM is a matroid isomorphic to M\I/JM\backslash I/J, where II and JJ are disjoint subsets of EE, M\IM\backslash I denotes the deletion of II in MM and M\I/JM\backslash I/J denotes the contraction of JJ in M\IM\backslash I.

Note that there are in general different pairs of subsets (I,J)(I,J) and (I,J)(I^{\prime},J^{\prime}) as above that give rise to isomorphic minors M\I/JM\I/JM\backslash I/J\simeq M\backslash I^{\prime}/J^{\prime}. In particular, [21, Prop. 3.3.6] shows that there is a co-independent subset JJ and an independent subset II of EE for every minor NN of MM such that IJ=I\cap J=\varnothing and NM\I/JN\simeq M\backslash I/J. Still, such II and JJ are in general not uniquely determined by NN, cf. Example 1.8.

If we fix II and JJ as above, then we can identify the ground set ENE_{N} of NN with E(IJ)E-(I\cup J), which yields an inclusion ι:ENE\myiota:E_{N}\to E. Since II is co-independent and JJ is independent, the set of bases of NN is

N={BJ|BM such that JBEI},{\mathcal{B}}_{N}\ =\ \big{\{}\,B-J\,\big{|}\,B\in{\mathcal{B}}_{M}\text{ such that }J\subset B\subset E-I\,\big{\}},

where M{\mathcal{B}}_{M} is the set of bases of MM. Consequently, the difference between the rank rr of MM and the rank rNr_{N} of NN is rrN=#Jr-r_{N}=\#J. Moreover, the inclusion ENEE_{N}\to E induces an inclusion

ι:NMBBJ.\begin{array}[]{cccc}\myiota:&{\mathcal{B}}_{N}&\longrightarrow&{\mathcal{B}}_{M}\\ &B&\longmapsto&B\cup J.\end{array}
Definition 1.5.

An embedded minor of MM is a minor N=M\I/JN=M\backslash I/J together with the pair (I,J)(I,J), where II is a co-independent subset and JJ is an independent subset JJ of EE such that IJ=I\cap J=\varnothing. By abuse of notation, we say that ι:NM\myiota:N\hookrightarrow M is an embedded minor, where N=M\I/JN=M\backslash I/J for fixed subsets II and JJ as above and where ι:NM\myiota:{\mathcal{B}}_{N}\to{\mathcal{B}}_{M} is the induced inclusion of the respective set of bases.

Let NN^{\prime} be a matroid. Then we say that an embedded minor ι:NM\myiota:N\hookrightarrow M is of type NN^{\prime}, or is an embedded NN^{\prime}-minor, if NN is isomorphic to NN^{\prime}.

Let NN and MM be matroids. A minor embedding of NN into MM is an isomorphism NM\I/JN\simeq M\backslash I/J of NN together with an embedded minor M\I/JMM\backslash I/J\hookrightarrow M of MM.

Given two minor embeddings ι:N=M\J/IM\myiota:N=M\backslash J/I\hookrightarrow M and ι:N=N\I/JN\myiota^{\prime}:N^{\prime}=N\backslash I^{\prime}/J^{\prime}\to N, we define the composition ιι\myiota\circ\myiota^{\prime} of ι\myiota^{\prime} with ι\myiota as the minor embedding N=M\(II)/(JJ)MN^{\prime}=M\backslash(I\cup I^{\prime})/(J\cup J^{\prime})\hookrightarrow M.

Example 1.6 (Embedded minors of type U42U^{2}_{4}).

Let MM be a matroid and ι:NM\myiota:N\to M an embedded minor of type U42U^{2}_{4}. Let II and JJ be as above. Then #J=r2\#J=r-2 since the rank of NN is 22, and EN=E(IJ)E_{N}=E-(I\cup J) has 44 elements e1,,e4e_{1},\dotsc,e_{4}. The set of bases N{\mathcal{B}}_{N} of NN consists of all 22-subsets of ENE_{N}, and thus

ι(N)={Jeiej|{i,j}{1,,4} and ij}.\myiota({\mathcal{B}}_{N})\ =\ \Big{\{}\,Je_{i}e_{j}\,\Big{|}\,\{i,j\}\subset\{1,\dotsc,4\}\text{ and }i\neq j\,\Big{\}}.
Remark 1.7.

Note that a composition N=N\I/JN=M\J/IMN^{\prime}=N\backslash I^{\prime}/J^{\prime}\hookrightarrow N=M\backslash J/I\hookrightarrow M of minor embeddings induces a composition NNM{\mathcal{B}}_{N^{\prime}}\to{\mathcal{B}}_{N}\to{\mathcal{B}}_{M} of inclusions of sets of bases. On the other hand, a minor embedding ι:N=M\J/IM\myiota:N=M\backslash J/I\to M decomposes into ι=ι1ι2\myiota=\myiota_{1}\circ\myiota_{2} with ι1:N=M\I1/J1M\myiota_{1}:N^{\prime}=M\backslash I_{1}/J_{1}\to M and ι2:N=N\I2/J2N\myiota_{2}:N=N^{\prime}\backslash I_{2}/J_{2}\to N^{\prime} for every pair of partitions I=I1I2I=I_{1}\cup I_{2} and J=J1J2J=J_{1}\cup J_{2}.

Note further that it is slightly inaccurate to suppress the subsets II and JJ from the notation of an embedded minor ι:NM\myiota:N\to M since they are in general not uniquely determined by the isomorphism type of NN and the injection ι:NM\myiota:{\mathcal{B}}_{N}\to{\mathcal{B}}_{M}, cf. Example 1.9. However, there is always a maximal choice for II and JJ for a given injection ι:NM\myiota:{\mathcal{B}}_{N}\to{\mathcal{B}}_{M}.

More precisely, for two disjoint subsets II and JJ of EE and =M{\mathcal{B}}={\mathcal{B}}_{M}, let \I/J={BJBEI}{\mathcal{B}}\backslash I/J=\{B\in{\mathcal{B}}\mid J\subset B\subset E-I\}. If \I/J{\mathcal{B}}\backslash I/J is not empty, then II is co-independent and JJ is independent and \I/J{\mathcal{B}}\backslash I/J is the image ι(M\I/J)\myiota({\mathcal{B}}_{M\backslash I/J})\subset{\mathcal{B}} for the embedded minor M\I/JM\backslash I/J of MM. Tautologically,

Imax=EB\I/JBandJmax=B\I/JBI_{\textup{max}}\ =\ E-\bigcup_{B\in{\mathcal{B}}\backslash I/J}B\qquad\text{and}\qquad J_{\textup{max}}\ =\ \bigcap_{B\in{\mathcal{B}}\backslash I/J}B

are the maximal co-independent and independent subsets of EE such that \I/J=\Imax/Jmax=ι(M\Imax/Jmax){\mathcal{B}}\backslash I/J={\mathcal{B}}\backslash I_{\textup{max}}/J_{\textup{max}}=\myiota({\mathcal{B}}_{M\backslash I_{\textup{max}}/J_{\textup{max}}}).

Example 1.8.

In the following, we illustrate how different choices of disjoint subsets II and JJ of EE lead to different injections ι:M\I/JM\myiota:{\mathcal{B}}_{M\backslash I/J}\to{\mathcal{B}}_{M}.

Let MM be the matroid on E={1,2,3}E=\{1,2,3\} whose set of bases is M={{1,2},{1,3}}{\mathcal{B}}_{M}=\big{\{}\{1,2\},\{1,3\}\big{\}}. Let N=M\{23}N=M\backslash\{23\} be the restriction of MM to {1}\{1\}, whose set of bases is N={{1}}{\mathcal{B}}_{N}=\big{\{}\{1\}\big{\}}. Since there is no canonical map NM{\mathcal{B}}_{N}\to{\mathcal{B}}_{M}, it is clear that not every pair of disjoint subsets II and JJ leads to an embedding M\I/JM{\mathcal{B}}_{M\backslash I/J}\to{\mathcal{B}}_{M}.

The minor NN is isomorphic to both N2=M\{2}/{3}N_{2}=M\backslash\{2\}/\{3\} and N3=M\{3}/{2}N_{3}=M\backslash\{3\}/\{2\}, which are embedded minors with respect to the inclusions ι2:N2M\myiota_{2}:{\mathcal{B}}_{N_{2}}\to{\mathcal{B}}_{M} with ι2({1})={1,2}\myiota_{2}(\{1\})=\{1,2\} and ι3:N3M\myiota_{3}:{\mathcal{B}}_{N_{3}}\to{\mathcal{B}}_{M} with ι3({1})={1,3}\myiota_{3}(\{1\})=\{1,3\}, respectively.

Example 1.9.

The contrary effect to that illustrated in Example 1.8 can also happen: different embedded minors can give rise to the same inclusions of sets of bases.

For instance, consider the matroid MM on E={1,2}E=\{1,2\} with M={{1,2}}{\mathcal{B}}_{M}=\big{\{}\{1,2\}\big{\}} and the embedded minor N=M\{2}N=M\backslash\{2\}. Then N={{1}}{\mathcal{B}}_{N}=\big{\{}\{1\}\big{\}} and the induced embedding ι:NM\myiota:{\mathcal{B}}_{N}\to{\mathcal{B}}_{M} is a bijection. This is obviously also the case for the trivial minor N=M=M\/N^{\prime}=M=M\backslash\varnothing/\varnothing. This shows that NN is not determined by ι:NM\myiota:{\mathcal{B}}_{N}\to{\mathcal{B}}_{M}.

2. Pastures

2.1. Definition and first properties

By a monoid with zero we mean a multiplicatively written commutative monoid PP with an element 0 that satisfies 0a=00\cdot a=0 for all aPa\in P. We denote the unit of PP by 11 and write P×P^{\times} for the group of invertible elements in PP. We denote by Sym3(P)\operatorname{Sym}_{3}(P) all elements of the form a+b+ca+b+c in the monoid semiring [P]{\mathbb{N}}[P], where a,b,cPa,b,c\in P.

Definition 2.1.

A pasture is a monoid PP with zero such that P×=P{0}P^{\times}=P-\{0\}, together with a subset NPN_{P} of Sym3(P)\operatorname{Sym}_{3}(P) such that for all a,b,c,dPa,b,c,d\in P

  1. (P1)

    a+0+0NPa+0+0\in N_{P} if and only if a=0a=0,

  2. (P2)

    if a+b+cNPa+b+c\in N_{P}, then ad+bd+cdad+bd+cd is in NPN_{P},

  3. (P3)

    there is a unique element ϵP×\myepsilon\in P^{\times} such that 1+ϵ+0NP1+\myepsilon+0\in N_{P}.

We call NPN_{P} the nullset of PP, and say that a+b+ca+b+c is null, and write symbolically a+b+c=0a+b+c=0, if a+b+cNPa+b+c\in N_{P}. For aPa\in P, we call ϵa\myepsilon a the weak inverse of aa.

The element ϵ\myepsilon plays the role of an additive inverse of 11, and the relations a+b+c=0a+b+c=0 express that certain sums of elements are zero, even though the multiplicative monoid PP does not carry an addition. For this reason, we will write frequently a-a for ϵa\myepsilon a and aba-b for a+ϵba+\myepsilon b. In particular, we have ϵ=1\myepsilon=-1. Moreover, we shall write a+b=ca+b=c or c=a+bc=a+b for a+b+ϵc=0a+b+\myepsilon c=0.

Remark 2.2.

As a word of warning, note that 1-1 is not an additive inverse of 11 if considered as elements in the semiring [P]{\mathbb{N}}[P], i.e. 11=1+ϵ01-1=1+\myepsilon\neq 0 as elements of [P]{\mathbb{N}}[P]. Psychologically, it is better to think of “-” as an involution on PP.

Definition 2.3.

A morphism of pastures is a multiplicative map f:P1P2f:P_{1}\to P_{2} with f(0)=0f(0)=0 and f(1)=1f(1)=1 such that f(a)+f(b)+f(c)=0f(a)+f(b)+f(c)=0 in NP2N_{P_{2}} whenever a+b+c=0a+b+c=0 in NP1N_{P_{1}}. This defines the category Pastures\operatorname{Pastures}.

Definition 2.4.

A subpasture of a pasture PP is a submonoid PP^{\prime} of PP together with a subset NPSym3(P)N_{P}^{\prime}\subset\operatorname{Sym}_{3}(P^{\prime}) such that a1Pa^{-1}\in P^{\prime} for every nonzero aPa\in P^{\prime} and a+b+cNPa+b+c\in N_{P^{\prime}} for all a+b+cNPa+b+c\in N_{P} with a,b,cPa,b,c\in P^{\prime}.

Given a subset SS of P×P^{\times}, the subpasture generated by SS is the submonoid P={0}SP^{\prime}=\{0\}\cup\langle S\rangle, where S\langle S\rangle denotes the subgroup of P×P^{\times} generated by SS, together with the nullset NP=NPSym3(P)N_{P^{\prime}}=N_{P}\cap\operatorname{Sym}_{3}(P^{\prime}).

Lemma 2.5.

Let PP be a pasture. Then a+b=0a+b=0 if and only if b=ϵab=\myepsilon a. In particular, we have ϵ2=1\myepsilon^{2}=1. Let f:P1P2f:P_{1}\to P_{2} be a morphism of pastures. Then f(ϵ)=ϵf(\myepsilon)=\myepsilon.

Proof.

Note that ϵ\myepsilon is uniquely determined by the relation 1+ϵ+0=01+\myepsilon+0=0. By (P2), this implies that ϵ1+1+0=0\myepsilon^{-1}+1+0=0 and thus by (P3), we conclude that ϵ1=ϵ\myepsilon^{-1}=\myepsilon, or equivalently, ϵ2=1\myepsilon^{2}=1.

Given a morphism f:P1P2f:P_{1}\to P_{2} be a morphism of pastures, the null relation 1+ϵ+0=01+\myepsilon+0=0 in P1P_{1} yields the relation f(1)+f(ϵ)+0=0f(1)+f(\myepsilon)+0=0 in P2P_{2}. Thus f(ϵ)f(\myepsilon) is the weak inverse of f(1)=1f(1)=1, which is ϵ\myepsilon. ∎

2.1.1. Free algebras and quotients

Let PP be a pasture with null set NPN_{P}. We define the free PP-algebra in x1,,xsx_{1},\dotsc,x_{s} as the pasture Px1,,xsP\langle x_{1},\dotsc,x_{s}\rangle whose unit group is Px1,,xs×=P××x1,,xsP\langle x_{1},\dotsc,x_{s}\rangle^{\times}=P^{\times}\times\langle x_{1},\dotsc,x_{s}\rangle, where x1,,xs\langle x_{1},\dotsc,x_{s}\rangle is the free abelian group generated by the symbols x1,,xsx_{1},\dotsc,x_{s}, and whose null set is

NPx1,,xs={da+db+dc|dx1,,xs,a+b+cNP},N_{P\langle x_{1},\dotsc,x_{s}\rangle}\ =\ \big{\{}da+db+dc\,\big{|}\,d\in\langle x_{1},\dotsc,x_{s}\rangle,\,a+b+c\in N_{P}\big{\}},

where dada stands for (a,d)Px1,,xs×(a,d)\in P\langle x_{1},\dotsc,x_{s}\rangle^{\times} if a0a\neq 0 and for 0 if a=0a=0. This pasture comes with a canonical morphism PPx1,,xsP\to P\langle x_{1},\dotsc,x_{s}\rangle of pastures that sends aa to 1a1a.

Let SSym3(P)S\subset\operatorname{Sym}_{3}(P) be a set of relations of the form a+b+ca+b+c with ab0ab\neq 0. We define the quotient PSP\!\sslash\!S of PP by SS as the following pasture. Let N~PS\tilde{N}_{P\!\sslash\!S} be the smallest subset of Sym3(P)\operatorname{Sym}_{3}(P) that is closed under property (P2) and that contains NPN_{P} and SS. Since all elements a+b+ca+b+c in SS have at least two nonzero terms by assumption, N~PS\tilde{N}_{P\!\sslash\!S} also satisfies (P1). But it might fail to satisfy (P3), necessitating the following quotient construction for P×P^{\times}.

We define the unit group (PS)×(P\!\sslash\!S)^{\times} of PSP\!\sslash\!S as the quotient of the group P×P^{\times} by the subgroup generated by all elements aa for which a1+0N~PSa-1+0\in\tilde{N}_{P\!\sslash\!S}. The underlying monoid of PSP\!\sslash\!S is, by definition, {0}(PS)×\{0\}\cup(P\!\sslash\!S)^{\times}, and it comes with a surjection π:PPS\mypi:P\to P\!\sslash\!S of monoids. We denote the image of aPa\in P by a¯=π(a)\bar{a}=\mypi(a), and define the null set of PSP\!\sslash\!S as the subset

NPS={a¯+b¯+c¯|a+b+cN~PS}N_{P\!\sslash\!S}\ =\ \big{\{}\bar{a}+\bar{b}+\bar{c}\,\big{|}\,a+b+c\in\tilde{N}_{P\!\sslash\!S}\big{\}}

of Sym3(PS)\operatorname{Sym}_{3}(P\!\sslash\!S). The quotient PSP\!\sslash\!S of PP by SS comes with a canonical map PPSP\to P\!\sslash\!S that sends aa to a¯\bar{a} and is a morphism of pastures.

If SSym3(Px1,,xs)S\subset\operatorname{Sym}_{3}(P\langle x_{1},\dots,x_{s}\rangle) is a subset of relations of the form a+b+ca+b+c with ab0ab\neq 0, then the composition of the canonical morphisms for the free algebra and for the quotient yields a canonical morphism

π:PPx1,,xsPx1,,xsS.\mypi:\ P\ \longrightarrow\ P\langle x_{1},\dots,x_{s}\rangle\ \longrightarrow\ P\langle x_{1},\dotsc,x_{s}\rangle\!\sslash\!S.

We denote by π0:{x1,,xs}Px1,,xsS\mypi_{0}:\{x_{1},\dotsc,x_{s}\}\to P\langle x_{1},\dotsc,x_{s}\rangle\!\sslash\!S the map that sends xix_{i} to x¯i\bar{x}_{i}.

The following result describes the universal property of Px1,,xsSP\langle x_{1},\dotsc,x_{s}\rangle\!\sslash\!S, which is analogous to the universal property of the quotient k[T1±1,,Tr±]/(S)k[T_{1}^{\pm 1},\dotsc,T_{r}^{\pm}]/(S) of the algebra of Laurent polynomials over a field kk by the ideal (S)(S) generated by a set SS of Laurent polynomials (each with only two or three terms). Note that the special case S=S=\varnothing yields the universal property of the free algebra Px1,,xsP\langle x_{1},\dotsc,x_{s}\rangle and the special case s=0s=0 yields the universal property of the quotient PSP\!\sslash\!S.

Proposition 2.6.

Let PP be a pasture, s0s\geqslant 0 and SSym3(Px1,,xs)S\subset\operatorname{Sym}_{3}(P\langle x_{1},\dots,x_{s}\rangle) a subset of relations of the form a+b+ca+b+c with ab0ab\neq 0. Let f:PQf:P\to Q be a morphism of pastures and f0:{x1,,xs}Q×f_{0}:\{x_{1},\dotsc,x_{s}\}\to Q^{\times} a map with the property that axiαi+bxiβi+cxiγiSa\prod x_{i}^{\myalpha_{i}}+b\prod x_{i}^{\mybeta_{i}}+c\prod x_{i}^{\mygamma_{i}}\in S with a,b,cPa,b,c\in P and αi,βi,γi\myalpha_{i},\mybeta_{i},\mygamma_{i}\in{\mathbb{Z}} for i=1,,ri=1,\dotsc,r implies that

f(a)f0(xi)αi+f(b)f0(xi)βi+f(c)f0(xi)γiNQ.\textstyle f(a)\prod f_{0}(x_{i})^{\myalpha_{i}}+f(b)\prod f_{0}(x_{i})^{\mybeta_{i}}+f(c)\prod f_{0}(x_{i})^{\mygamma_{i}}\in N_{Q}.

Then there is a unique morphism f^:Px1,,xsSQ\hat{f}:P\langle x_{1},\dotsc,x_{s}\rangle\!\sslash\!S\to Q such that the diagrams

P{P}Q{Q}Px1,,xsS{P\langle x_{1},\dotsc,x_{s}\rangle\!\sslash\!S}f\scriptstyle{f}π\scriptstyle{\mypi}f^\scriptstyle{\hat{f}}      and      {x1,,xs}{\{x_{1},\dotsc,x_{s}\}}Q{Q}Px1,,xsS{P\langle x_{1},\dotsc,x_{s}\rangle\!\sslash\!S}f0\scriptstyle{f_{0}}π0\scriptstyle{\mypi_{0}}f^\scriptstyle{\hat{f}}

commute.

Proof.

We claim that the association

f^:Px1,,xsSQaxiαif(a)f0(xi)αi\begin{array}[]{cccc}\hat{f}:&P\langle x_{1},\dotsc,x_{s}\rangle\!\sslash\!S&\longrightarrow&Q\\ &a\prod x_{i}^{\myalpha_{i}}&\longmapsto&f(a)\prod f_{0}(x_{i})^{\myalpha_{i}}\end{array}

is a morphism of pastures. Once we have proven this, it is clear that f=f^πf=\hat{f}\circ\mypi and f0=f^π0f_{0}=\hat{f}\circ\mypi_{0}. Since the unit group of P^=Px1,,xsS\widehat{P}=P\langle x_{1},\dotsc,x_{s}\rangle\!\sslash\!S is generated by {axiaP×,i=1,,s}\{ax_{i}\mid a\in P^{\times},i=1,\dotsc,s\}, it follows that f^\hat{f} is uniquely determined by the conditions f=f^πf=\hat{f}\circ\mypi and f0=f^π0f_{0}=\hat{f}\circ\mypi_{0}.

We are left with the verification that f^\hat{f} is a morphism. As a first step, we show that the restriction f^×:P^×Q×\hat{f}^{\times}:\widehat{P}^{\times}\to Q^{\times} defines a group homomorphism. Note that NP^={yz+yz+yz′′yP^×,z+z+z′′S}N_{\widehat{P}}=\{yz+yz^{\prime}+yz^{\prime\prime}\mid y\in\widehat{P}^{\times},z+z^{\prime}+z^{\prime\prime}\in S\}. Thus we have an equality axiαi=bxiβia\prod x_{i}^{\myalpha_{i}}=b\prod x_{i}^{\mybeta_{i}} in P^×\widehat{P}^{\times} if and only if daxiαi+δidbxiβi+δiSda\prod x_{i}^{\myalpha_{i}+\mydelta_{i}}-db\prod x_{i}^{\mybeta_{i}+\mydelta_{i}}\in S for some dxiδiP^×d\prod x_{i}^{\mydelta_{i}}\in\widehat{P}^{\times}. By our assumptions, we have f(da)f0(xi)αi+δif(db)f0(xi)βi+δiNQf(da)\prod f_{0}(x_{i})^{\myalpha_{i}+\mydelta_{i}}-f(db)\prod f_{0}(x_{i})^{\mybeta_{i}+\mydelta_{i}}\in N_{Q}, and thus multiplying with f(d1)f0(xi)δif(d^{-1})\prod f_{0}(x_{i})^{-\mydelta_{i}} yields f^(axiαi)=f^(bxiβi)\hat{f}(a\prod x_{i}^{\myalpha_{i}})=\hat{f}(b\prod x_{i}^{\mybeta_{i}}). This verifies that f^×:P^×Q×\hat{f}^{\times}:\widehat{P}^{\times}\to Q^{\times} is well-defined as a map. It is clear from the definition that it is a group homomorphism.

For showing that f^:P^Q\hat{f}:\widehat{P}\to Q is a morphism of pastures, we need to verify that for every element z+z+z′′z+z^{\prime}+z^{\prime\prime} in NP^N_{\widehat{P}}, the element f^(z)+f^(z)+f^(z′′)\hat{f}(z)+\hat{f}(z^{\prime})+\hat{f}(z^{\prime\prime}) is in NQN_{Q}. This can be done by a similar argument as before. We omit the details. ∎

2.1.2. Examples

The regular partial field is the pasture 𝔽1±={0,1,1}{11}{{\mathbb{F}}_{1}^{\pm}}=\{0,1,-1\}\!\sslash\!\{1-1\} whose multiplication is determined by (1)2=1(-1)^{2}=1.

Let KK be a field and KK^{\bullet} its multiplicative monoid. Then we can associate with KK the pasture K{a+b+c|a+b+c=0 in K}K^{\bullet}\!\sslash\!\{a+b+c\;|\;a+b+c=0\text{ in }K\}. We can recover the addition of KK by the rule c=a+b-c=a+b if a+b+c=0a+b+c=0. In particular, we can identify the finite field with 22 elements with the pasture 𝔽2=𝔽1±{1+1}{\mathbb{F}}_{2}={{\mathbb{F}}_{1}^{\pm}}\!\sslash\!\{1+1\}, which implies that 1=1-1=1, and the finite field with 33 elements with the pasture 𝔽3=𝔽1±{1+1+1}{\mathbb{F}}_{3}={{\mathbb{F}}_{1}^{\pm}}\!\sslash\!\{1+1+1\}.

Let FF be a hyperfield and FF^{\bullet} its multiplicative monoid. Then we can associate with FF the pasture F{a+b+c| 0a[-0]b[-0]c in F}F^{\bullet}\!\sslash\!\{a+b+c\;|\;0\in a\mathrel{\,\raisebox{-1.1pt}{\larger[-0]{$\boxplus$}}\,}b\mathrel{\,\raisebox{-1.1pt}{\larger[-0]{$\boxplus$}}\,}c\text{ in }F\}. In particular, we can realize the Krasner hyperfield as 𝕂=𝔽1±{1+1,1+1+1}{\mathbb{K}}={{\mathbb{F}}_{1}^{\pm}}\!\sslash\!\{1+1,1+1+1\}, and the sign hyperfield as 𝕊=𝔽1±{1+11}{\mathbb{S}}={{\mathbb{F}}_{1}^{\pm}}\!\sslash\!\{1+1-1\}.

The near-regular partial field is

𝕌=𝔽1±x,y{x+y1}.{\mathbb{U}}={{\mathbb{F}}_{1}^{\pm}}\langle x,y\rangle\!\sslash\!\{x+y-1\}.

The dyadic partial field is

𝔻=𝔽1±z{z+z1}.{\mathbb{D}}\ =\ {{\mathbb{F}}_{1}^{\pm}}\langle z\rangle\!\sslash\!\{z+z-1\}.

The hexagonal partial field is

=𝔽1±z{z3+1,zz21}.{\mathbb{H}}\ =\ {{\mathbb{F}}_{1}^{\pm}}\langle z\rangle\!\sslash\!\{z^{3}+1,z-z^{2}-1\}.

It is a straightforward exercise to verify that these descriptions of 𝕌,𝔻,{\mathbb{U}},{\mathbb{D}},{\mathbb{H}} agree with the definitions given in the introduction.

As final examples, the weak sign hyperfield is the pasture

𝕎=𝔽1±1+1+1,1+11{\mathbb{W}}\ =\ {{\mathbb{F}}_{1}^{\pm}}\!\sslash\!\langle 1+1+1,1+1-1\rangle

and the phase hyperfield is the pasture {\mathbb{P}} whose unit group ×{\mathbb{P}}^{\times} is the subgroup of norm 11-elements in ×{\mathbb{C}}^{\times} and whose null set is

N={a+b+cSym3(P)|a,b,c>0 is an -linear subspace of }N_{\mathbb{P}}\ =\ \Big{\{}a+b+c\in\operatorname{Sym}_{3}(P)\,\Big{|}\,\langle a,b,c\rangle_{>0}\text{ is an ${\mathbb{R}}$-linear subspace of ${\mathbb{C}}$}\Big{\}}

where a,b,c>0\langle a,b,c\rangle_{>0} is the smallest cone in {\mathbb{C}} that contains aa, bb and cc. In fact, {\mathbb{P}} is isomorphic to the quotient of the pasture associated with {\mathbb{C}} by the action of >0{\mathbb{R}}_{>0} by multiplication.

2.1.3. Initial and final objects

The category Pastures\operatorname{Pastures} admits both initial and final objects. The initial object of Pastures\operatorname{Pastures} is the regular partial field 𝔽1±{{\mathbb{F}}_{1}^{\pm}}. Given a pasture PP, we denote by iPi_{P} the unique initial morphism iP:𝔽1±Pi_{P}:{\mathbb{F}}_{1}^{\pm}\to P.

The final object of Pastures\operatorname{Pastures} is the Krasner hyperfield 𝕂{\mathbb{K}}. Given a pasture PP, we denote by tPt_{P} the unique terminal morphism tP:P𝕂t_{P}:P\to{\mathbb{K}} sending 0 to 0 and every nonzero element of PP to 11.

2.1.4. Products and coproducts

The category Pastures\operatorname{Pastures} admits both a product and coproduct.

Let P1,P2P_{1},P_{2} be pastures. The (categorical) product P1×P2P_{1}\times P_{2} can be constructed explicitly as follows. As sets, we have P1×P2=(P1×P2×){0}P_{1}\times P_{2}=(P_{1}^{\times}\oplus P_{2}^{\times})\cup\{0\}, endowed with the coordinatewise multiplication on P1×P2×P_{1}^{\times}\oplus P_{2}^{\times}, extended by the rule (a1,a2)0=0(a1,a2)=0(a_{1},a_{2})\cdot 0=0\cdot(a_{1},a_{2})=0, and the nullset is the subset

NP1×P2={(a1,a2)+(b1,b2)+(c1,c2)|ai+bi+ciNPi for i=1,2}N_{P_{1}\times P_{2}}\ =\ \Big{\{}(a_{1},a_{2})+(b_{1},b_{2})+(c_{1},c_{2})\,\Big{|}\,a_{i}+b_{i}+c_{i}\in N_{P_{i}}\text{ for }i=1,2\Big{\}}

of Sym3(P1×P2)\operatorname{Sym}^{3}(P_{1}\times P_{2}).

The categorical coproduct is given by the tensor product P1P2P_{1}\otimes P_{2} defined as follows. As sets, we have P1P2=(P1×P2)/P_{1}\otimes P_{2}=(P_{1}\times P_{2})/\sim, where P1×P2P_{1}\times P_{2} denotes the Cartesian product (not the underlying set of the product in the category of pastures) and (x1,x2)(y1,y2)(x_{1},x_{2})\sim(y_{1},y_{2}) if and only if either:

  • At least one of x1,x2x_{1},x_{2} is zero and at least one of y1,y2y_{1},y_{2} is zero; or

  • x1=y1x_{1}=y_{1} and x2=y2x_{2}=y_{2}; or

  • x1=y1x_{1}=-y_{1} and x2=y2x_{2}=-y_{2}.

Denoting the equivalence class of (x1,x2)(x_{1},x_{2}) by x1x2x_{1}\otimes x_{2}, the additive relations are given by:

  • ay+by+cyNP1P2a\otimes y+b\otimes y+c\otimes y\in N_{P_{1}\otimes P_{2}} for yP2y\in P_{2} and a,b,cP1a,b,c\in P_{1} with a+b+cNP1a+b+c\in N_{P_{1}}.

  • xa+xb+xcNP1P2x\otimes a+x\otimes b+x\otimes c\in N_{P_{1}\otimes P_{2}} for xP1x\in P_{1} and a,b,cP2a,b,c\in P_{2} with a+b+cNP2a+b+c\in N_{P_{2}}.

Lemma 2.7.

The tensor product of pastures satisfies the universal property of a coproduct with respect to the morphisms f1:P1P1P2f_{1}:P_{1}\to P_{1}\otimes P_{2} and f2:P2P1P2f_{2}:P_{2}\to P_{1}\otimes P_{2} given by xx1x\mapsto x\otimes 1 and y1yy\mapsto 1\otimes y, respectively.

Proof.

Given a pasture PP and morphisms gi:PiPg_{i}:P_{i}\to P for i=1,2i=1,2, we must show that there is a unique morphism g:P1P2Pg:P_{1}\otimes P_{2}\to P such that gi=gfig_{i}=g\circ f_{i} for i=1,2i=1,2.

Define gg by the formula g(x1x2)=g1(x1)g2(x2)g(x_{1}\otimes x_{2})=g_{1}(x_{1})\cdot g_{2}(x_{2}). To see that this is well-defined, suppose (x1,x2)(y1,y2)(x_{1},x_{2})\sim(y_{1},y_{2}). If x1x2=0x_{1}x_{2}=0 and y1y2=0y_{1}y_{2}=0, then g(x1x2)=g(y1y2)=0g(x_{1}\otimes x_{2})=g(y_{1}\otimes y_{2})=0. Otherwise xi=(1)kyix_{i}=(-1)^{k}y_{i} for i=1,2i=1,2 with k{0,1}k\in\{0,1\}, and we have

g(x1x2)=(1)kg1(x1)(1)kg2(x2)=g1(y1)g2(y2)=g(y1y2).g(x_{1}\otimes x_{2})=(-1)^{k}g_{1}(x_{1})(-1)^{k}g_{2}(x_{2})=g_{1}(y_{1})g_{2}(y_{2})=g(y_{1}\otimes y_{2}).

Hence gg is well-defined.

It is straightforward to verify that gfi=gig\circ f_{i}=g_{i} for i=1,2i=1,2 and that gg is a morphism.

To see that gg is unique, suppose gg^{\prime} is another such morphism. Then g(x11)=g1(x1)g^{\prime}(x_{1}\otimes 1)=g_{1}(x_{1}) and g(1x2)=g2(x2)g^{\prime}(1\otimes x_{2})=g_{2}(x_{2}), and since gg^{\prime} is a morphism we have

g(x1x2)=g((x11)(1x2))=g(x11)g(1x2)=g1(x1)g2(x2)g^{\prime}(x_{1}\otimes x_{2})=g^{\prime}((x_{1}\otimes 1)(1\otimes x_{2}))=g^{\prime}(x_{1}\otimes 1)g^{\prime}(1\otimes x_{2})=g_{1}(x_{1})g_{2}(x_{2})

for all x1P1x_{1}\in P_{1} and x2P2x_{2}\in P_{2}. Thus g=gg^{\prime}=g. ∎

By comparison, the category of fields (which is a full subcategory of Pastures\operatorname{Pastures}) does not have an initial object, a final object, products, or coproducts.

Example 2.8.

We have 𝔽2×𝔽3𝔽1±{\mathbb{F}}_{2}\times{\mathbb{F}}_{3}\cong{\mathbb{F}}_{1}^{\pm} and 𝔽2𝔽3𝕂{\mathbb{F}}_{2}\otimes{\mathbb{F}}_{3}\cong{\mathbb{K}}. The first isomorphism follows easily from our formula for the product of two pastures, and the second is an immediate consequence of the following lemma, which in turn follows easily from the universal property of the coproduct.

Lemma 2.9.

If P2𝔽1±SP_{2}\cong{{\mathbb{F}}_{1}^{\pm}}\!\sslash\!S, where SSym3(𝔽1±)S\subseteq\operatorname{Sym}_{3}({{\mathbb{F}}_{1}^{\pm}}), then P1P2P1SP_{1}\otimes P_{2}\cong P_{1}\!\sslash\!S.

Example 2.10.

We have 𝔽3×𝕊𝔻{z2}{\mathbb{F}}_{3}\times{\mathbb{S}}\simeq{\mathbb{D}}\!\sslash\!\{z^{2}\} and 𝔽3𝕊𝔽1±{1+1+1,1+11}{\mathbb{F}}_{3}\otimes{\mathbb{S}}\simeq{{\mathbb{F}}_{1}^{\pm}}\!\sslash\!\{1+1+1,1+1-1\}. For the first isomorphism, note that the underlying set of 𝔽3×𝕊{\mathbb{F}}_{3}\times{\mathbb{S}} is ({±1}×{±1}){0}\left(\{\pm 1\}\times\{\pm 1\}\right)\cup\{0\} while the underlying set of 𝔻{z2}{\mathbb{D}}\!\sslash\!\{z^{2}\} is ({±1}×{±z}){0}\left(\{\pm 1\}\times\{\pm z\}\right)\cup\{0\}. One checks easily that the map sending (1,1)(1,1) to 11 and (1,1)(-1,1) to zz is an isomorphism of pastures. The second isomorphism is a consequence of Lemma 2.9.

Example 2.11.

Here (without proof) are a few more examples of products and coproducts:

  • 𝔽1±=𝔽2×𝕊=𝔽2×𝕎{{\mathbb{F}}_{1}^{\pm}}={\mathbb{F}}_{2}\times{\mathbb{S}}={\mathbb{F}}_{2}\times{\mathbb{W}}.

  • 𝕂=𝔽2𝕊=𝔽2𝕎{\mathbb{K}}={\mathbb{F}}_{2}\otimes{\mathbb{S}}={\mathbb{F}}_{2}\otimes{\mathbb{W}}.

  • =𝔽3×𝔽4{\mathbb{H}}={\mathbb{F}}_{3}\times{\mathbb{F}}_{4}.

Remark 2.12.

More generally, one can show that the category Pastures\operatorname{Pastures} is complete and co-complete, i.e., it admits all small limits and colimits. In particular, one can form arbitrary fiber products and fiber coproducts in Pastures\operatorname{Pastures}. We omit the details since we will not need these more general statements in the present paper.

2.1.5. Comparison with partial fields, hyperfields, fuzzy rings, tracts and ordered blueprints

The definitions of partial fields, hyperfields, fuzzy rings, tracts and ordered blueprints, and a comparison thereof, can be found in [5]. We are not aiming at repeating all definitions, but we will explain how the category of pastures fits within the landscape of these types of algebraic objects.

We have already explained how partial fields and hyperfields give rise to pastures. The tract associated with a pasture PP is defined as F=(P×,NF)F=(P^{\times},N_{F}), where NFN_{F} is the ideal generated by NPN_{P} in [P×]{\mathbb{N}}[P^{\times}]. The ordered blueprint associated to a pasture PP is defined as B=P{0u+v+wu+v+wNP}B=P\!\sslash\!\{0\leqslant u+v+w\mid u+v+w\in N_{P}\}.

These associations yield fully faithful embeddings of the category PartFields\operatorname{{PartFields}} of partial fields and the category HypFields\operatorname{{HypFields}} of hyperfields into Pastures\operatorname{Pastures}, and of Pastures\operatorname{Pastures} into the category Tracts\operatorname{{Tracts}} of tracts and into the category OBlpr±\operatorname{{OBlpr}}^{\pm} of ordered blueprints with unique weak inverses. This completes the diagram of [5, Thm. 2.21] to

PartFields{\operatorname{{PartFields}}}Pastures{\operatorname{Pastures}}Tracts{\operatorname{{Tracts}}}Fields{\operatorname{{Fields}}}HypFields{\operatorname{{HypFields}}}FuzzRings{\operatorname{{FuzzRings}}}OBlpr±{\operatorname{{OBlpr}}^{\pm}}\scriptstyle{\vdash}

where FuzzRings\operatorname{{FuzzRings}} is the category of fuzzy rings. This diagram commutes and all functors are fully faithful, with exception of the adjunction between Tracts\operatorname{{Tracts}} and OBlpr±\operatorname{{OBlpr}}^{\pm}. We omit the details of these claims.

Note that fuzzy rings, seen as objects in either Tracts\operatorname{{Tracts}} or OBlpr±\operatorname{{OBlpr}}^{\pm}, are not pastures in general since the ideal II of the fuzzy ring might not be generated by 33-term elements of [P×]{\mathbb{N}}[P^{\times}]. Conversely, not every pasture, seen as a tract or as an ordered blueprint, gives rise to a fuzzy ring since the axiom (FR2) (in the notation of [5, Section 2.4]) might not be satisfied. An example of a pasture for which (FR2) fails to hold is the pasture 𝔽1±z{z2+1,1+1+z}{{\mathbb{F}}_{1}^{\pm}}\langle z\rangle\!\sslash\!\{z^{2}+1,1+1+z\}; cf. [5, Ex. 2.11] for more details on this example.

2.2. Matroid representations

We recall the notion of weak matroids over pastures from [3]. Let PP be a pasture. A weak Grassmann–Plücker function of rank rr on EE with values in PP is a function Δ:ErP\Delta:E^{r}\to P such that:

  1. (1)

    The set of rr-element subsets {e1,,er}E\{e_{1},\ldots,e_{r}\}\subseteq E such that Δ(e1,,er)0\Delta(e_{1},\ldots,e_{r})\neq 0 is the set of bases of a matroid M¯\underline{M}.

  2. (2)

    Δ(σ(e1),,σ(er))=sign(σ)Δ(e1,,er)\Delta(\mysigma(e_{1}),\ldots,\mysigma(e_{r}))=\operatorname{sign}(\mysigma)\cdot\Delta(e_{1},\ldots,e_{r}) for all permutations σSr\mysigma\in S_{r}.

  3. (3)

    Δ\Delta satisfies the 3-term Plücker relations: for all 𝐉Er2{\mathbf{J}}\in E^{r-2} and all (e1,e2,e3,e4)E4(e_{1},e_{2},e_{3},e_{4})\in E^{4},

    Δ(𝐉e1e2)Δ(𝐉e3e4)Δ(𝐉e1e3)Δ(𝐉e2e4)+Δ(𝐉e1e4)Δ(𝐉e2e3)= 0.\Delta({\mathbf{J}}e_{1}e_{2})\cdot\Delta({\mathbf{J}}e_{3}e_{4})-\Delta({\mathbf{J}}e_{1}e_{3})\cdot\Delta({\mathbf{J}}e_{2}e_{4})+\Delta({\mathbf{J}}e_{1}e_{4})\cdot\Delta({\mathbf{J}}e_{2}e_{3})\ =\ 0.

Two weak Grassmann–Plücker functions Δ,Δ\Delta,\Delta^{\prime} are isomorphic if there is a cP×c\in P^{\times} such that Δ(e1,,er)=cΔ(e1,,er)\Delta^{\prime}(e_{1},\ldots,e_{r})=c\Delta(e_{1},\ldots,e_{r}) for all (e1,,er)Er(e_{1},\ldots,e_{r})\in E^{r}.

A weak PP-matroid MM of rank rr on EE is an isomorphism class of weak Grassmann–Plücker functions Δ:ErP\Delta:E^{r}\to P.

We call M¯\underline{M} the underlying matroid of MM, and we refer to Δ\Delta as a PP-representation of M¯\underline{M}.

We say that a matroid M¯\underline{M} is representable over a pasture PP if there is at least one PP-representation of M¯\underline{M}.

Remark 2.13.

In [3] one also finds a definition of strong PP-matroids, but this will not play a role in the present paper. We therefore omit the adjective “weak” when talking about PP-representations.

With this terminology, we introduce the following subclasses of matroids:

Definition 2.14.

A matroid MM is

  • regular if it is representable over 𝔽1±{{\mathbb{F}}_{1}^{\pm}};

  • binary if it is representable over 𝔽2{\mathbb{F}}_{2};

  • ternary if it is representable over 𝔽3{\mathbb{F}}_{3};

  • quaternary if it is representable over 𝔽4{\mathbb{F}}_{4};

  • near-regular if it is representable over 𝕌{\mathbb{U}};

  • dyadic if it is representable over 𝔻{\mathbb{D}};

  • hexagonal if it is representable over {\mathbb{H}};

  • 𝔻{\mathbb{D}}\otimes{\mathbb{H}}-representable555In [24, p. 55], the partial field 𝔻{\mathbb{D}}\otimes{\mathbb{H}} is denoted 𝕐{\mathbb{Y}}. if it is representable over 𝔻{\mathbb{D}}\otimes{\mathbb{H}};

  • representable if it representable over some field;

  • orientable if it is representable over 𝕊{\mathbb{S}}.

  • weakly orientable if it is representable over 𝕎{\mathbb{W}}.

Note that hexagonal matroids are also called 16\sqrt[6]{1}-matroids or sixth-root-of-unity-matroids in the literature, cf. [24] and [27].

2.3. Matroid representations via hyperplane functions

There are various “cryptomorphic” descriptions of weak PP-matroids, for example in terms of “weak PP-circuits”, cf. [3]. For the purposes of the present paper, it will be more convenient to reformulate things in terms of hyperplanes rather than circuits.

Definition 2.15.

Let PP be a pasture and let M¯\underline{M} be a matroid on the finite set EE. Let ¯\underline{{\mathcal{H}}} be the set of hyperplanes of M¯\underline{M}.

  1. (1)

    Given H¯H\in\underline{{\mathcal{H}}}, we say that fH:EPf_{H}:E\to P is a PP-hyperplane function for HH if fH(e)=0f_{H}(e)=0 if and only if eHe\in H.

  2. (2)

    Two PP-hyperplane functions fH,fHf_{H},f^{\prime}_{H} for HH are projectively equivalent if there exists cP×c\in P^{\times} such that fH(e)=cfH(e)f_{H}^{\prime}(e)=cf_{H}(e) for all eEe\in E.

  3. (3)

    A triple of hyperplanes (H1,H2,H3)¯3(H_{1},H_{2},H_{3})\in\underline{{\mathcal{H}}}^{3} is modular if F=H1H2H3F=H_{1}\cap H_{2}\cap H_{3} is a flat of corank 22 such that F=HiHjF=H_{i}\cap H_{j} for all distinct i,j{1,2,3}i,j\in\{1,2,3\}.

  4. (4)

    A modular system of PP-hyperplane functions for M¯\underline{M} is a collection of PP-hyperplane functions fH:EPf_{H}:E\to P, one for each H¯H\in\underline{{\mathcal{H}}}, such that whenever H1,H2,H3H_{1},H_{2},H_{3} is a modular triple of hyperplanes in ¯\underline{{\mathcal{H}}}, the corresponding functions HiH_{i} are linearly dependent, i.e., there exist constants c1,c2,c3c_{1},c_{2},c_{3} in PP, not all zero, such that

    c1fH1(e)+c2fH2(e)+c3fH3(e)=0c_{1}f_{H_{1}}(e)+c_{2}f_{H_{2}}(e)+c_{3}f_{H_{3}}(e)=0

    for all eEe\in E.

  5. (5)

    Two modular systems of PP-hyperplane functions {fH}\{f_{H}\} and {fH}\{f^{\prime}_{H}\} are equivalent if fHf_{H} and fHf^{\prime}_{H} are projectively equivalent for all H¯H\in\underline{{\mathcal{H}}}.

The following result can be viewed as a generalization of “Tutte’s representation theorem” [33, Theorem 5.1] (compare with  [15, Theorem 3.5]). One can also view it as adding to the collection of cryptomorphisms for weak matroids established in [3].

Theorem 2.16.

Let PP be a pasture and let M¯\underline{M} be a matroid of rank rr on EE. Let ¯\underline{{\mathcal{H}}} be the set of hyperplanes of M¯\underline{M}. There is a canonical bijection

Ξ:{P-representations of M¯}{modular systems of P-hyperplanes for M¯}.\begin{array}[]{cccc}\Xi:&\Big{\{}\text{$P$-representations of $\underline{M}$}\Big{\}}&\longrightarrow&\Big{\{}\text{modular systems of $P$-hyperplanes for $\underline{M}$}\Big{\}}.\end{array}

If Δ:ErP\Delta:E^{r}\to P is a PP-representation of M¯\underline{M} and =Ξ(Δ){\mathcal{H}}=\Xi(\Delta), then

fH(e)fH(e)=Δ(𝐈e)Δ(𝐈e)\frac{f_{H}(e)}{f_{H}(e^{\prime})}\ =\ \frac{\Delta({\mathbf{I}}e)}{\Delta({\mathbf{I}}e^{\prime})}

for every fHf_{H}\in{\mathcal{H}}, elements e,eEHe,e^{\prime}\in E-H and 𝐈Er1{\mathbf{I}}\in E^{r-1} such that |𝐈||{\mathbf{I}}| is an independent set which spans HH.

Proof.

Let MM be a weak PP-matroid with underlying matroid M¯\underline{M}. Let HH be a hyperplane of M¯\underline{M}. The complement of HH in EE is a cocircuit D¯\underline{D} of M¯\underline{M}; choose a PP-cocircuit DD of MM whose support is D¯\underline{D}. Now define fH:EPf_{H}:E\to P by fH(e)=D(e)f_{H}(e)=D(e). Then fH(e)=0f_{H}(e)=0 iff D(e)=0D(e)=0 iff eD¯e\not\in\underline{D} iff eHe\in H, so fHf_{H} is a PP-hyperplane function for HH.

Suppose H1,H2,H3H_{1},H_{2},H_{3} is a modular triple of hyperplanes of M¯\underline{M} with intersection FF, a flat of corank 2. Let ee be an element of H3FH_{3}-F. Then eH3(H1H2)e\in H_{3}-(H_{1}\cup H_{2}) by the covering axiom for flats [21, Exercise 1.4.11, Axiom (F3)]. Let D1D_{1} and D2D_{2} be the PP-cocircuits of MM corresponding to H1H_{1} and H2H_{2}, respectively, and let α1=D2(e),α2=D1(e)P\myalpha_{1}=D_{2}(e),\myalpha_{2}=-D_{1}(e)\in P. Then α1D1(e)=α2D2(e)\myalpha_{1}D_{1}(e)=-\myalpha_{2}D_{2}(e), so by [3, Axiom (C3){\rm(C3)}^{\prime}], there is a PP-cocircuit D3D_{3} of MM such that D3(e)=0D_{3}(e)=0 and α1D1(f)+α2D2(f)D3(f)=0\myalpha_{1}D_{1}(f)+\myalpha_{2}D_{2}(f)-D_{3}(f)=0 for all fEf\in E. By [3, Lemma 3.7], the support of D3D_{3} is EH3E-H_{3}. By [3, Axiom (C2)], D3D_{3} is a scalar multiple of fH3f_{H_{3}}, say D3=α3fH3D_{3}=-\myalpha_{3}f_{H_{3}}. Then α1fH1+α2fH2+α3fH3=0\myalpha_{1}f_{H_{1}}+\myalpha_{2}f_{H_{2}}+\myalpha_{3}f_{H_{3}}=0, so {fH}\{f_{H}\} is a modular system of PP-hyperplane functions for M¯\underline{M}.

Conversely, a similar argument shows that given a modular system of PP-hyperplane functions {fH}\{f_{H}\} for M¯\underline{M}, there is a corresponding family of PP-cocircuits 𝒟{\mathcal{D}} defining a weak PP-matroid MM. These operations are inverse to one another by construction, and this establishes the desired bijection.

We turn to the second claim, which is obvious for e=ee=e^{\prime}, so we may assume that eee\neq e^{\prime}. Let n=#En=\#E and choose 𝐈Enr1{\mathbf{I}}^{\prime}\in E^{n-r-1} such that E=|𝐈||𝐈|{e,e}E=|{\mathbf{I}}|\cup|{\mathbf{I}}^{\prime}|\cup\{e,e^{\prime}\}. Note that since |𝐈e||{\mathbf{I}}e^{\prime}| is a basis of M¯\underline{M}, the complement |𝐈e||{\mathbf{I}}^{\prime}e| is a basis for M¯\underline{M}^{\ast}. If 𝐈=(i1,,ir1){\mathbf{I}}=(i_{1},\dotsc,i_{r-1}) and 𝐈=(i1,,inr1){\mathbf{I}}^{\prime}=(i^{\prime}_{1},\dotsc,i^{\prime}_{n-r-1}), we define a total order on EE by

i1<<inr1<e<i1<<ir1<e.i^{\prime}_{1}<\dotsb<i^{\prime}_{n-r-1}<e<i_{1}<\dotsb<i_{r-1}<e^{\prime}.

By [3, Lemma 4.1], there is a dual Grassmann-Plücker function Δ:EnrP\Delta^{\ast}:E^{n-r}\to P to Δ\Delta that satisfies

Δ(𝐈e)=sign(idE)Δ(𝐈e)=Δ(𝐈e)\Delta^{\ast}({\mathbf{I}}^{\prime}e)\ =\ \operatorname{sign}(\textup{id}_{E})\cdot\Delta({\mathbf{I}}e^{\prime})\ =\ \Delta({\mathbf{I}}e^{\prime})

and

Δ(𝐈e)=sign(τe,e)Δ(𝐈e)=Δ(𝐈e),\Delta^{\ast}({\mathbf{I}}^{\prime}e^{\prime})\ =\ \operatorname{sign}(\mytau_{e,e^{\prime}})\cdot\Delta({\mathbf{I}}e)\ =\ -\Delta({\mathbf{I}}e),

where idE:EE\textup{id}_{E}:E\to E is the identity and τe,e:EE\mytau_{e,e^{\prime}}:E\to E is the transposition that exchanges ee with ee^{\prime}. This implies that

fH(e)fH(e)=Δ(𝐈e)Δ(𝐈e)=Δ(𝐈e)Δ(𝐈e)\frac{f_{H}(e)}{f_{H}(e^{\prime})}\ =\ -\frac{\Delta^{\ast}({\mathbf{I}}^{\prime}e^{\prime})}{\Delta^{\ast}({\mathbf{I}}^{\prime}e)}\ =\ \frac{\Delta({\mathbf{I}}e)}{\Delta({\mathbf{I}}e^{\prime})}

as desired, where we use [3, Def. 4.6 and Lemma 4.7] for the first equality. ∎

2.4. The universal pasture

The universal pasture of a matroid was introduced in [5] as a tool to control the representations of a matroid MM over other pastures. We review this in the following.

The symmetric group SrS_{r} on rr elements acts by permutation of coefficients on ErE^{r}. In the following, we understand the sign sign(σ)\operatorname{sign}(\mysigma) of a permutation σSr\mysigma\in S_{r} as an element of (𝔽1±)×={±1}({{\mathbb{F}}_{1}^{\pm}})^{\times}=\{\pm 1\}.

Definition 2.17.

Let MM be a matroid with Grassmann-Plücker function Δ:Er𝕂\Delta:E^{r}\to{\mathbb{K}}. The extended universal pasture of MM is the pasture PM+=𝔽1±T𝐈|Δ(𝐈)0{S}P_{M}^{+}={{\mathbb{F}}_{1}^{\pm}}\langle T_{\mathbf{I}}|\Delta({\mathbf{I}})\neq 0\rangle\!\sslash\!\{S\}, where SS is the set of the relations Tσ(𝐈)=sign(σ)T𝐈T_{\mysigma({\mathbf{I}})}=\operatorname{sign}(\mysigma)T_{\mathbf{I}} for all 𝐈Er{\mathbf{I}}\in E^{r} and σSr\mysigma\in S_{r}, together with the 33-term Plücker relations

T𝐉e1e2T𝐉e3e4T𝐉e1e3T𝐉e2e4+T𝐉e1e4T𝐉e2e3= 0T_{{\mathbf{J}}e_{1}e_{2}}T_{{\mathbf{J}}e_{3}e_{4}}-T_{{\mathbf{J}}e_{1}e_{3}}T_{{\mathbf{J}}e_{2}e_{4}}+T_{{\mathbf{J}}e_{1}e_{4}}T_{{\mathbf{J}}e_{2}e_{3}}\ =\ 0

for all 𝐉Er2{\mathbf{J}}\in E^{r-2} and e1,,e4Ee_{1},\dotsc,e_{4}\in E.

The pasture PM+P_{M}^{+} is naturally graded by the rule that T𝐈T_{\mathbf{I}} has degree 11 for every 𝐈supp(Δ){\mathbf{I}}\in\operatorname{supp}(\Delta). The universal pasture of MM is the subpasture PMP_{M} of degree 0-elements of PM+P_{M}^{+}.

The relevance of the universal pasture is that it represents the set of isomorphism classes of PP-representations of MM. This is derived in [5] by means of the algebraic geometry of the moduli space of matroids. We include an independent, and more elementary, proof in the following.

Theorem 2.18 ([5, Prop. 6.22]).

Let MM be a matroid of rank rr on EE and PP a pasture. Then there is a functorial bijection between the set of isomorphism classes of PP-representations of MM and Hom(PM,P)\operatorname{Hom}(P_{M},P). In particular, MM is representable over PP if and only if there is a morphism PMPP_{M}\to P.

Proof.

Let Δ:ErP\Delta:E^{r}\to P be a PP-representation of MM and PM+P_{M}^{+} the extended universal pasture of MM. Define the map χΔ,0+:T𝐈Δ(𝐈)\mychi^{+}_{\Delta,0}:T_{\mathbf{I}}\mapsto\Delta({\mathbf{I}}) from the set {T𝐈𝐈supp(Δ)}\{T_{\mathbf{I}}\mid{\mathbf{I}}\in\operatorname{supp}(\Delta)\} of generators of PM+P_{M}^{+} to PP. Let SS be the set of 33-term Plücker relations

T𝐉e1e2T𝐉e3e4T𝐉e1e3T𝐉e2e4+T𝐉e1e4T𝐉e2e3,T_{{\mathbf{J}}e_{1}e_{2}}T_{{\mathbf{J}}e_{3}e_{4}}-T_{{\mathbf{J}}e_{1}e_{3}}T_{{\mathbf{J}}e_{2}e_{4}}+T_{{\mathbf{J}}e_{1}e_{4}}T_{{\mathbf{J}}e_{2}e_{3}},

where 𝐉Er2{\mathbf{J}}\in E^{r-2} and e1,,e4Ee_{1},\dotsc,e_{4}\in E such that |𝐉e1e4||{\mathbf{J}}e_{1}\dotsc e_{4}| has r+2r+2 elements. Applying χΔ,0+\mychi^{+}_{\Delta,0} to this relation, with the convention that χΔ,0+(T𝐈)=0\mychi^{+}_{\Delta,0}(T_{\mathbf{I}})=0 if Δ(𝐈)=0\Delta({\mathbf{I}})=0, yields

χΔ,0+(T𝐉e1e2)χΔ,0+(T𝐉e3e4)χΔ,0+(T𝐉e1e3)χΔ,0+(T𝐉e2e4)+χΔ,0+(T𝐉e1e4)χΔ,0+(T𝐉e2e3)=Δ(𝐉e1e2)Δ(𝐉e3e4)Δ(𝐉e1e3)Δ(𝐉e2e4)+Δ(𝐉e1e4)Δ(𝐉e2e3),\mychi^{+}_{\Delta,0}(T_{{\mathbf{J}}e_{1}e_{2}})\mychi^{+}_{\Delta,0}(T_{{\mathbf{J}}e_{3}e_{4}})-\mychi^{+}_{\Delta,0}(T_{{\mathbf{J}}e_{1}e_{3}})\mychi^{+}_{\Delta,0}(T_{{\mathbf{J}}e_{2}e_{4}})+\mychi^{+}_{\Delta,0}(T_{{\mathbf{J}}e_{1}e_{4}})\mychi^{+}_{\Delta,0}(T_{{\mathbf{J}}e_{2}e_{3}})\\ =\ \Delta({\mathbf{J}}e_{1}e_{2})\Delta({\mathbf{J}}e_{3}e_{4})-\Delta({\mathbf{J}}e_{1}e_{3})\Delta({\mathbf{J}}e_{2}e_{4})+\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{2}e_{3}),

which is an element of NPN_{P} since Δ\Delta is a Grassmann-Plücker function. Thus, by Proposition 2.6, the map χM,0+\mychi^{+}_{M,0} together with the unique morphism 𝔽1±P{{\mathbb{F}}_{1}^{\pm}}\to P define a morphism

χΔ+:PM+=𝔽1±T𝐈𝐈supp(Δ)SP\mychi^{+}_{\Delta}:\ P_{M}^{+}={{\mathbb{F}}_{1}^{\pm}}\langle T_{\mathbf{I}}\mid{\mathbf{I}}\in\operatorname{supp}(\Delta)\rangle\!\sslash\!S\ \longrightarrow\ P

with χΔ+(T𝐈)=Δ(𝐈)\mychi^{+}_{\Delta}(T_{\mathbf{I}})=\Delta({\mathbf{I}}) for 𝐈supp(Δ){\mathbf{I}}\in\operatorname{supp}(\Delta). We define χΔ:PMP\mychi_{\Delta}:P_{M}\to P as the composition of the inclusion PMPM+P_{M}\to P_{M}^{+} with χΔ+\mychi^{+}_{\Delta}. Since every element of PMP_{M} has degree 0, we have χΔ=χaΔ\mychi_{\Delta}=\mychi_{a\Delta} for every aP×a\in P^{\times}, which shows that χΔ\mychi_{\Delta} depends only on the isomorphism class of Δ\Delta.

This yields a canonical map

{isomorphism classes of P-representations of M}Hom(PM,P),[Δ]χΔ\begin{array}[]{ccc}\Big{\{}\text{isomorphism classes of $P$-representations of $M$}\Big{\}}&\longrightarrow&\operatorname{Hom}(P_{M},P),\\ {}[\Delta]&\longmapsto&\mychi_{\Delta}\end{array}

which turns out to be a bijection whose inverse can be described as follows. Let χ:PMP\mychi:P_{M}\to P be a morphism. Choose an 𝐈0Er{\mathbf{I}}_{0}\in E^{r} such that |𝐈0||{\mathbf{I}}_{0}| is a basis of MM and define the map

Δχ:ErP,𝐈{χ(T𝐈/T𝐈0)if |𝐈| is a basis of M;0otherwise.\begin{array}[]{cccl}\Delta_{\mychi}:&E^{r}&\longrightarrow&P,\\ &{\mathbf{I}}&\longmapsto&\bigg{\{}\begin{array}[]{ll}\text{\footnotesize$\mychi(T_{\mathbf{I}}/T_{{\mathbf{I}}_{0}})$}&\text{\footnotesize if $|{\mathbf{I}}|$ is a basis of $M$;}\\ \text{\footnotesize 0}&\text{\footnotesize otherwise.}\end{array}\end{array}

This is a Grassmann-Plücker function, since

Δχ(𝐉e1e2)Δχ(𝐉e3e4)Δχ(𝐉e1e3)Δχ(𝐉e2e4)+Δχ(𝐉e1e4)Δχ(𝐉e2e3)=χ(T𝐉e1e2T𝐈0)χ(T𝐉e3e4T𝐈0)χ(T𝐉e1e3T𝐈0)χ(T𝐉e2e4T𝐈0)+χ(T𝐉e1e4T𝐈0)χ(T𝐉e2e3T𝐈0)\Delta_{\mychi}({\mathbf{J}}e_{1}e_{2})\Delta_{\mychi}({\mathbf{J}}e_{3}e_{4})-\Delta_{\mychi}({\mathbf{J}}e_{1}e_{3})\Delta_{\mychi}({\mathbf{J}}e_{2}e_{4})+\Delta_{\mychi}({\mathbf{J}}e_{1}e_{4})\Delta_{\mychi}({\mathbf{J}}e_{2}e_{3})\\ \textstyle=\ \mychi\bigg{(}\frac{T_{{\mathbf{J}}e_{1}e_{2}}}{T_{{\mathbf{I}}_{0}}}\bigg{)}\mychi\bigg{(}\frac{T_{{\mathbf{J}}e_{3}e_{4}}}{T_{{\mathbf{I}}_{0}}}\bigg{)}-\mychi\bigg{(}\frac{T_{{\mathbf{J}}e_{1}e_{3}}}{T_{{\mathbf{I}}_{0}}}\bigg{)}\mychi\bigg{(}\frac{T_{{\mathbf{J}}e_{2}e_{4}}}{T_{{\mathbf{I}}_{0}}}\bigg{)}+\mychi\bigg{(}\frac{T_{{\mathbf{J}}e_{1}e_{4}}}{T_{{\mathbf{I}}_{0}}}\bigg{)}\mychi\bigg{(}\frac{T_{{\mathbf{J}}e_{2}e_{3}}}{T_{{\mathbf{I}}_{0}}}\bigg{)}

is in the nullset of PMP_{M}. Note that the isomorphism class of Δχ\Delta_{\mychi} is independent of the choice of 𝐈0{\mathbf{I}}_{0}, since any two such choices yield Grassmann-Plücker functions that are constant multiples of each other.

It is straightforward to verify that the associations χ[Δχ]\mychi\mapsto[\Delta_{\mychi}] and [Δ]χΔ[\Delta]\mapsto\mychi_{\Delta} are mutually inverse, and that both maps are functorial in PP; we omit the details. ∎

Remark 2.19.

We call the morphism χΔ:PMP\mychi_{\Delta}:P_{M}\to P associated with the (isomorphism class of a) PP-representation Δ\Delta the characteristic morphism.

The proof of Theorem 2.18 also shows that the set of PP-representations of MM are in functorial bijection with Hom(PM+,P)\operatorname{Hom}(P_{M}^{+},P). Under this identification, the identity morphism PM+PM+P_{M}^{+}\to P_{M}^{+} defines a PM+P_{M}^{+}-representation Δ^:ErPM+\widehat{\Delta}:E^{r}\to P_{M}^{+} of MM, which we call the universal Grassmann-Plücker function of MM. It satisfies Δ^(𝐈)=T𝐈\widehat{\Delta}({\mathbf{I}})=T_{\mathbf{I}} if |𝐈||{\mathbf{I}}| is a basis of MM and Δ^(𝐈)=0\widehat{\Delta}({\mathbf{I}})=0 otherwise, and tPM+Δ^:Er𝕂t_{P_{M}^{+}}\circ\widehat{\Delta}:E^{r}\to{\mathbb{K}} is a Grassmann-Plücker function for MM where tPM+:PM+𝕂t_{P_{M}^{+}}:P_{M}^{+}\to{\mathbb{K}} is the terminal morphism, cf. section 2.1.3.

2.5. The Tutte group and the universal pasture

The connection between the Tutte group and the universal pasture is explained in Theorem 6.26 of [5], which is as follows:

Theorem 2.20.

Let MM be a matroid with Grassmann-Plücker function Δ:Er𝕂\Delta:E^{r}\to{\mathbb{K}}. The association 11-1\mapsto-1 and T𝐈X𝐈T_{\mathbf{I}}\mapsto X_{\mathbf{I}} for 𝐈supp(Δ){\mathbf{I}}\in\operatorname{supp}(\Delta) defines an isomorphism of groups (PM+)×𝕋M(P_{M}^{+})^{\times}\to{\mathbb{T}}_{M}^{\mathcal{B}} that restricts to an isomorphism PM×𝕋MP_{M}^{\times}\to{\mathbb{T}}_{M}.

Remark 2.21.

Dress and Wenzel show in [15, Thm. 3.7] that a matroid MM is representable over a fuzzy ring RR if and only if there is a group homomorphism 𝕋MR×{\mathbb{T}}_{M}\to R^{\times} that preserves the Plücker relations. This can be seen as an analogue of Theorem 2.18 in the formalism of Dress and Wenzel, but it also lets us explain the advantage of our formulation.

Namely, the foundation of a matroid is an object in the same category Pastures\operatorname{Pastures} as the coefficient domains for matroid representations. We can thus use standard arguments from category theory to deduce results about the representability of a matroid. For example, if the foundation of a matroid MM is the tensor product F1F2F_{1}\otimes F_{2} of two pastures F1F_{1} and F2F_{2}, then MM is representable over a third pasture PP if and only if there exist morphisms F1PF_{1}\to P and F2PF_{2}\to P. We will make a frequent use of this observation in section 6.

3. Cross ratios

In this section, we review the theory of cross ratios for matroids from different angles, and explain the connection between these viewpoints, which are derived from cryptomorphic descriptions of a matroid in terms of bases and hyperplanes. There are two principally different types of cross ratios: cross ratios for PP-matroids, which are elements of PP, and universal cross ratios of a matroid MM, which are elements of the universal pasture PMP_{M} of MM. It turns out that there is a close relation between these two types of cross ratios and their different incarnations in terms of bases and hyperplanes. In particular, we identify in a concluding subsection the set of universal cross ratios with the set of fundamental elements in PMP_{M}.

3.1. Cross ratios of PP-matroids

Let E={1,,n}E=\{1,\dotsc,n\} and 0rn0\leqslant r\leqslant n. Let PP be a pasture and MM a PP-matroid with Grassmann-Plücker function Δ:ErP\Delta:E^{r}\to P.

Define ΩM\Omega_{M} to be the set of tuples (J;e1,,e4)(J;e_{1},\dotsc,e_{4}) for which there exists a 𝐉Er2{\mathbf{J}}\in E^{r-2} with underlying set |𝐉|=J|{\mathbf{J}}|=J such that

Δ(𝐉e1e4)Δ(𝐉e2e3)Δ(𝐉e1e3)Δ(𝐉e2e4) 0,\Delta({\mathbf{J}}e_{1}e_{4})\ \Delta({\mathbf{J}}e_{2}e_{3})\ \Delta({\mathbf{J}}e_{1}e_{3})\ \Delta({\mathbf{J}}e_{2}e_{4})\ \neq\ 0,

where 𝐉ekel=(j1,,jr2,ek,el){\mathbf{J}}e_{k}e_{l}=(j_{1},\dotsc,j_{r-2},e_{k},e_{l}).

Definition 3.1.

Let MM be a PP-matroid with Grassmann-Plücker function Δ:ErP\Delta:E^{r}\to P and (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M}. The cross ratio of (J;e1,,e4)(J;e_{1},\dotsc,e_{4}) in MM is the element

[e1e2e3e4]M,J=[e1e2e3e4]Δ,𝐉=Δ(𝐉e1e3)Δ(𝐉e2e4)Δ(𝐉e1e4)Δ(𝐉e2e3)\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\Delta,{\mathbf{J}}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\Delta,{\mathbf{J}}}}{}{}\ =\ \frac{\Delta({\mathbf{J}}e_{1}e_{3})\Delta({\mathbf{J}}e_{2}e_{4})}{\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{2}e_{3})}

of PP for any 𝐉Er2{\mathbf{J}}\in E^{r-2} with |𝐉|=J|{\mathbf{J}}|=J .

Note that the value of the cross ratio [e1e2e3e4]M,J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}}{}{} does not depend on the ordering of 𝐉{\mathbf{J}}, nor on the choice of Grassmann-Plücker function Δ\Delta for MM, which justifies our notation.

We find the following relations between cross ratios with permuted arguments. Let (J;e1,,e4)ΩM(J;e_{1},\dots,e_{4})\in\Omega_{M} and 𝐉Er2{\mathbf{J}}\in E^{r-2} be such that J=|𝐉|J=|{\mathbf{J}}|. We say that (J;e1,,e4)(J;e_{1},\dots,e_{4}) is non-degenerate if

Δ(𝐉e1e2)Δ(𝐉e3e4) 0,\Delta({\mathbf{J}}e_{1}e_{2})\Delta({\mathbf{J}}e_{3}e_{4})\ \neq\ 0,

or equivalently, if [eσ(1)eσ(2)eσ(3)eσ(4)]M,J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{\mysigma(1)}$}}&{\scalebox{0.9}{$e_{\mysigma(2)}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{\mysigma(3)}$}}&{{\scalebox{0.9}{$e_{\mysigma(4)}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{\mysigma(1)}}&{e_{\mysigma(2)}}\\ {e_{\mysigma(3)}}&{e_{\mysigma(4)}}\end{smallmatrix}\big{]}_{M,J}}{}{} is defined and nonzero for every permutation σ\mysigma of {1,,4}\{1,\dotsc,4\}. We define ΩM\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}} to be the subset of ΩM\Omega_{M} consisting of all non-degenerate (J;e1,,e4)(J;e_{1},\dots,e_{4}). We call a cross ratio [e1e2e3e4]M,J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}}{}{} non-degenerate if (J;e1,,e4)(J;e_{1},\dots,e_{4}) is non-degenerate. We call (J;e1,,e4)ΩM(J;e_{1},\dots,e_{4})\in\Omega_{M} degenerate if it is not in ΩM\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}}.

One finds some relations that follow immediately from the definition, such as the fact that permuting rows and columns has no effect on the value of the cross ratio, i.e.

[e1e2e3e4]M,J=[e2e1e4e3]M,J=[e3e4e1e2]M,J=[e4e3e2e1]M,J;\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{2}$}}&{\scalebox{0.9}{$e_{1}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{4}$}}&{{\scalebox{0.9}{$e_{3}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{2}}&{e_{1}}\\ {e_{4}}&{e_{3}}\end{smallmatrix}\big{]}_{M,J}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{3}$}}&{\scalebox{0.9}{$e_{4}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{1}$}}&{{\scalebox{0.9}{$e_{2}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{3}}&{e_{4}}\\ {e_{1}}&{e_{2}}\end{smallmatrix}\big{]}_{M,J}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{4}$}}&{\scalebox{0.9}{$e_{3}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{2}$}}&{{\scalebox{0.9}{$e_{1}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{4}}&{e_{3}}\\ {e_{2}}&{e_{1}}\end{smallmatrix}\big{]}_{M,J}}{}{};

that permuting the last two entries inverts the cross ratio, i.e.

[e1e2e4e3]M,J=[e1e2e3e4]M,J1;\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{4}$}}&{{\scalebox{0.9}{$e_{3}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{4}}&{e_{3}}\end{smallmatrix}\big{]}_{M,J}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}^{-1}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}^{-1}}{}{};

and that a cyclic rotation of the last three entries yields the relation

[e1e2e3e4]M,J[e1e3e4e2]M,J[e1e4e2e3]M,J=1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{3}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{4}$}}&{{\scalebox{0.9}{$e_{2}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{3}}\\ {e_{4}}&{e_{2}}\end{smallmatrix}\big{]}_{M,J}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{4}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{2}$}}&{{\scalebox{0.9}{$e_{3}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{4}}\\ {e_{2}}&{e_{3}}\end{smallmatrix}\big{]}_{M,J}}{}{}\ =\ -1

if (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}} is non-degenerate. We will discuss these relations and others in detail in Theorem 4.20.

The cross ratios keep track of the Plücker relations

(2) Δ(𝐉e1e2)Δ(𝐉e3e4)Δ(𝐉e1e3)Δ(𝐉e2e4)+Δ(𝐉e1e4)Δ(𝐉e2e3)= 0\Delta({\mathbf{J}}e_{1}e_{2})\Delta({\mathbf{J}}e_{3}e_{4})-\Delta({\mathbf{J}}e_{1}e_{3})\Delta({\mathbf{J}}e_{2}e_{4})+\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{2}e_{3})\ =\ 0

satisfied by the Grassmann-Plücker function Δ:ErP\Delta:E^{r}\to P. Namely, if (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}} and 𝐉Er2{\mathbf{J}}\in E^{r-2} are such that J=|𝐉|J=|{\mathbf{J}}|, then dividing both sides of the Plücker relation (2) by Δ(𝐉e1e4)Δ(𝐉e2e3)-\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{2}e_{3}) yields the Plücker relation for cross ratios

[e1e2e3e4]M,J+[e1e3e2e4]M,J= 1,\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}}{}{}+\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{3}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{2}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{3}}\\ {e_{2}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}}{}{}\ =\ 1,

where the notation a+b=ca+b=c in a pasture PP is short-hand for a+bcNPa+b-c\in N_{P}.

If (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M} is degenerate, then Δ(𝐉e1e2)Δ(𝐉e3e4)=0\Delta({\mathbf{J}}e_{1}e_{2})\Delta({\mathbf{J}}e_{3}e_{4})=0 and dividing the Plücker relation by Δ(𝐉e1e4)Δ(𝐉e2e3)-\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{2}e_{3}) yields [e1e2e3e4]M,J1=0\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}}{}{}-1=0, and thus

[e1e2e3e4]M,J= 1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}}{}{}\ =\ 1

by the uniqueness of additive inverses in PP.

Lemma 3.2.

Let PP be a pasture and MM a PP-matroid of rank rr on EE with dual MM^{\ast}. Let (J;e1,,e4)ΩM¯(J;e_{1},\dots,e_{4})\in\Omega_{\underline{M}} and I=EJe1e4I=E-Je_{1}\dotsc e_{4}. Then

[e1e2e3e4]M,I=[e1e2e3e4]M,J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M^{\ast},I}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M^{\ast},I}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}}{}{}

as elements of PP.

Proof.

Let n=#En=\#E. Choose 𝐉=(j1,,jr2){\mathbf{J}}=(j_{1},\dotsc,j_{r-2}) with |𝐉|=J|{\mathbf{J}}|=J and 𝐈=(i1,,inr2){\mathbf{I}}=(i_{1},\dotsc,i_{n-r-2}) with |𝐈|=I|{\mathbf{I}}|=I. Choose a total order on EE. Let Δ:ErP\Delta:E^{r}\to P be a Grassmann-Plücker function for MM. Then by [3, Lemma 4.2], there is a Grassmann-Plücker function Δ:EnrP\Delta^{\ast}:E^{n-r}\to P for MM^{\ast} such that for all identifications {i,j,k,l}={1,2,3,4}\{i,j,k,l\}=\{1,2,3,4\}, we have

Δ(𝐈eiek)=sign(πi,j,k,l)Δ(𝐉ejel),\Delta^{\ast}({\mathbf{I}}e_{i}e_{k})\ =\ \operatorname{sign}(\mypi_{i,j,k,l})\cdot\Delta({\mathbf{J}}e_{j}e_{l}),

where π=πi,j,k,l\mypi=\mypi_{i,j,k,l} is the permutation of EE such that

π(i1)<<π(inr2)<π(ei)<π(ek)<π(j1)<<π(jr2)<π(ej)<π(el)\mypi(i_{1})<\dotsc<\mypi(i_{n-r-2})<\mypi(e_{i})<\mypi(e_{k})<\mypi(j_{1})<\dotsc<\mypi(j_{r-2})<\mypi(e_{j})<\mypi(e_{l})

in the chosen total order of EE. Since πi,j,l,k=πi,j,k,lτk,l\mypi_{i,j,l,k}=\mypi_{i,j,k,l}\circ\mytau_{k,l} for the transposition τk,l\mytau_{k,l} that exchanges eke_{k} and ele_{l}, we have sign(πi,j,k,l)/sign(πi,j,l,k)=1\operatorname{sign}(\mypi_{i,j,k,l})/\operatorname{sign}(\mypi_{i,j,l,k})=-1. Thus we obtain

[e1e2e3e4]M,I=Δ(𝐈e1e3)Δ(𝐈e2e4)Δ(𝐈e1e4)Δ(𝐈e2e3)=sign(π1,2,3,4)sign(π1,2,4,3)sign(π2,1,4,3)sign(π2,1,3,4)Δ(𝐉e2e4)Δ(𝐉e1e3)Δ(𝐉e2e3)Δ(𝐉e1e4)=[e1e2e3e4]M,J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M^{\ast},I}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M^{\ast},I}}{}{}\ =\ \frac{\Delta^{\ast}({\mathbf{I}}e_{1}e_{3})\Delta^{\ast}({\mathbf{I}}e_{2}e_{4})}{\Delta^{\ast}({\mathbf{I}}e_{1}e_{4})\Delta^{\ast}({\mathbf{I}}e_{2}e_{3})}\\ =\ \frac{\operatorname{sign}(\mypi_{1,2,3,4})}{\operatorname{sign}(\mypi_{1,2,4,3})}\cdot\frac{\operatorname{sign}(\mypi_{2,1,4,3})}{\operatorname{sign}(\mypi_{2,1,3,4})}\cdot\frac{\Delta({\mathbf{J}}e_{2}e_{4})\Delta({\mathbf{J}}e_{1}e_{3})}{\Delta({\mathbf{J}}e_{2}e_{3})\Delta({\mathbf{J}}e_{1}e_{4})}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}}{}{}

as claimed. ∎

3.2. Cross ratios for hyperplanes

There is a different, but closely related, notion of cross ratios associated to certain quadruples of hyperplanes.

Definition 3.3.

Let MM be a matroid of rank rr on EE and {\mathcal{H}} be its set of hyperplanes. A quadruple of hyperplanes (H1,,H4)4(H_{1},\dotsc,H_{4})\in{\mathcal{H}}^{4} is modular if F=H1H2H3H4F=H_{1}\cap H_{2}\cap H_{3}\cap H_{4} is a flat of corank 22 such that F=HiHjF=H_{i}\cap H_{j} for all i{1,2}i\in\{1,2\} and j{3,4}j\in\{3,4\}. A modular quadruple (H1,,H4)(H_{1},\dotsc,H_{4}) is non-degenerate if F=HiHjF=H_{i}\cap H_{j} for all distinct i,j{1,,4}i,j\in\{1,\dotsc,4\}. Otherwise it is called degenerate.666Note that in some papers the term “modular quadruple” is used for what we call a non-degenerate quadruple; e.g. see [3], [7, Def. 5.1] and [25, Def. 3.18]. We denote the set of all modular quadruples of hyperplanes by ΘM\Theta_{M} and the subset of all non-degenerate modular quadruples by ΘM\Theta_{M}^{\scalebox{0.7}{$\diamondsuit$}}.

Definition 3.4.

Let PP be a pasture and MM a PP-matroid with underlying matroid M¯\underline{M}. Let (H1,,H4)ΘM¯(H_{1},\dotsc,H_{4})\in\Theta_{\underline{M}}. The cross ratio of (H1,,H4)(H_{1},\dotsc,H_{4}) in MM is the element

[H1H2H3H4]M=f1(e3)f2(e4)f1(e4)f2(e3)\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}_{M}}{}{}\ =\ \frac{f_{1}(e_{3})f_{2}(e_{4})}{f_{1}(e_{4})f_{2}(e_{3})}

of PP, where fi:EPf_{i}:E\to P is a PP-hyperplane function for HiH_{i} for i=1,2i=1,2 (cf. Definition 2.15), and where ekHkFe_{k}\in H_{k}-F for k=3,4k=3,4 with F=H1H4F=H_{1}\cap\dotsb\cap H_{4}.

Since f1f_{1} and f2f_{2} are determined by H1H_{1} and H2H_{2} up to a factor in P×P^{\times}, the definition of [H1H2H3H4]M\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}_{M}}{}{} is independent of the choices of f1f_{1} and f2f_{2}. It follows from [3, Theorem 3.21, Lemma 4.5, and Definition 4.6] that it is also independent of the choices of e3e_{3} and e4e_{4}.

We continue with a comparison of the two notions of cross ratios.

Lemma 3.5.

Let MM be a matroid of rank rr on EE. The association (J;e1,,e4)(H1,,H4)(J;e_{1},\dotsc,e_{4})\mapsto(H_{1},\dotsc,H_{4}) with Hi=JeiH_{i}=\langle Je_{i}\rangle for i=1,,4i=1,\dotsc,4 defines a surjective map Ψ:ΩMΘM\Psi:\Omega_{M}\to\Theta_{M}, which restricts to a surjective map Ψ:ΩMΘM\Psi^{\scalebox{0.7}{$\diamondsuit$}}:\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}}\to\Theta_{M}^{\scalebox{0.7}{$\diamondsuit$}}.

Proof.

The flat F=H1H4=JF=H_{1}\cap\dotsb\cap H_{4}=\langle J\rangle is of rank r2r-2 since JJ is an independent set of rank r2r-2. We have HiHj=FH_{i}\cap H_{j}=F for all i=1,2i=1,2 and j=3,4j=3,4 since Δ(Jeiej)0\Delta(Je_{i}e_{j})\neq 0 and thus HiHj=E\langle H_{i}\cup H_{j}\rangle=E. This shows that (H1,,H4)(H_{1},\dotsc,H_{4}) is indeed a modular quadruple. By the same reasoning applied to arbitrary distinct i,j{1,,4}i,j\in\{1,\dotsc,4\}, we conclude that Ψ\Psi restricts to a map Ψ:ΩMΘ\Psi^{\scalebox{0.7}{$\diamondsuit$}}:\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}}\to\Theta_{\mathcal{H}}^{\scalebox{0.7}{$\diamondsuit$}}.

Given (H1,,H4)ΘM(H_{1},\dotsc,H_{4})\in\Theta_{M} and F=H1H4F=H_{1}\cap\dotsb\cap H_{4}, choose an independent subset JFJ\subset F with r2r-2 elements and eiHiFe_{i}\in H_{i}-F for i=1,,4i=1,\dotsc,4. Since HiHk=FH_{i}\cap H_{k}=F for i{1,2}i\in\{1,2\} and k{3,4}k\in\{3,4\}, the closure of JeiekJe_{i}e_{k} is EE, i.e. JeiekJe_{i}e_{k} is a basis of MM. Thus (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M} and Ψ(J;e1,,e4)=(H1,,H4)\Psi(J;e_{1},\dotsc,e_{4})=(H_{1},\dotsc,H_{4}), which establishes the surjectivity of Ψ\Psi. If (H1,,H4)ΘM(H_{1},\dotsc,H_{4})\in\Theta_{M}^{\scalebox{0.7}{$\diamondsuit$}}, then HiHk=FH_{i}\cap H_{k}=F and thus JeiekJe_{i}e_{k} is a basis of MM for all distinct i,k{1,,4}i,k\in\{1,\dotsc,4\}. Thus (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}} and Ψ(J;e1,,e4)=(H1,,H4)\Psi^{\scalebox{0.7}{$\diamondsuit$}}(J;e_{1},\dotsc,e_{4})=(H_{1},\dotsc,H_{4}), which establishes the surjectivity of Ψ\Psi^{\scalebox{0.7}{$\diamondsuit$}}. ∎

Proposition 3.6.

Let PP be a pasture and MM a PP-matroid with underlying matroid M¯\underline{M}. Let (J;e1,,e4)ΩM¯(J;e_{1},\dotsc,e_{4})\in\Omega_{\underline{M}} and (H1,,H4)=Ψ(J;e1,,e4)(H_{1},\dotsc,H_{4})=\Psi(J;e_{1},\dotsc,e_{4}). Then we have

[H1H2H3H4]M=[e1e2e3e4]M,J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}_{M}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}}{}{}

as elements of PP.

Proof.

Since |𝐉ei||{\mathbf{J}}e_{i}| is an (r1)(r-1)-set that generates HiH_{i} and ejHie_{j}\notin H_{i} for i{1,2}i\in\{1,2\} and j{3,4}j\in\{3,4\}, we can apply Theorem 2.16 to conclude that

[H1H2H3H4]M=f1(e3)f2(e4)f1(e4)f2(e3)=Δ(𝐉e1e3)Δ(𝐉e2e4)Δ(𝐉e1e4)Δ(𝐉e2e3)=[e1e2e3e4]M,J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}_{M}}{}{}\ =\ \frac{f_{1}(e_{3})f_{2}(e_{4})}{f_{1}(e_{4})f_{2}(e_{3})}\ =\ \frac{\Delta({\mathbf{J}}e_{1}e_{3})\Delta({\mathbf{J}}e_{2}e_{4})}{\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{2}e_{3})}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}}{}{}

as claimed. ∎

Our comparison of different notions of cross ratios has the following immediate consequence.

Corollary 3.7.

Let MM be a matroid and (J;e1,,e4),(J;f1,,f4)ΩM(J;e_{1},\dotsc,e_{4}),(J^{\prime};f_{1},\dotsc,f_{4})\in\Omega_{M}. If Jei=Jfi\langle Je_{i}\rangle=\langle J^{\prime}f_{i}\rangle for i=1,,4i=1,\dotsc,4, then [e1e2e3e4]J=[f1f2f3f4]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$f_{1}$}}&{\scalebox{0.9}{$f_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$f_{3}$}}&{{\scalebox{0.9}{$f_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{f_{1}}&{f_{2}}\\ {f_{3}}&{f_{4}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{}.

Proof.

By Proposition 3.6, we have [e1e2e3e4]J=[H1H2H3H4]=[f1f2f3f4]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{}=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$f_{1}$}}&{\scalebox{0.9}{$f_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$f_{3}$}}&{{\scalebox{0.9}{$f_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{f_{1}}&{f_{2}}\\ {f_{3}}&{f_{4}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{} if Hi=Jei=JfiH_{i}=\langle Je_{i}\rangle=\langle J^{\prime}f_{i}\rangle for i=1,,4i=1,\dotsc,4. ∎

3.3. Universal cross ratios

Let MM be a matroid of rank rr on E={1,,n}E=\{1,\dotsc,n\} with Grassmann-Plücker function Δ:Er𝕂\Delta:E^{r}\to{\mathbb{K}}.

Recall from section 2.4 the definition of the extended universal pasture

PM+=𝔽1±T𝐈|Δ(𝐈)0{S}P_{M}^{+}\ =\ {{\mathbb{F}}_{1}^{\pm}}\langle T_{\mathbf{I}}|\Delta({\mathbf{I}})\neq 0\rangle\!\sslash\!\{S\}

of MM, where SS contains the relations Tσ(𝐈)=sign(σ)T𝐈T_{\mysigma({\mathbf{I}})}=\operatorname{sign}(\mysigma)T_{\mathbf{I}} and the 33-term Plücker relations

T𝐉e1e2T𝐉e3e4T𝐉e1e3T𝐉e2e4+T𝐉e1e4T𝐉e2e3= 0T_{{\mathbf{J}}e_{1}e_{2}}T_{{\mathbf{J}}e_{3}e_{4}}-T_{{\mathbf{J}}e_{1}e_{3}}T_{{\mathbf{J}}e_{2}e_{4}}+T_{{\mathbf{J}}e_{1}e_{4}}T_{{\mathbf{J}}e_{2}e_{3}}\ =\ 0

for all 𝐉Er2{\mathbf{J}}\in E^{r-2} and e1,,e4Ee_{1},\dotsc,e_{4}\in E, where we use the convention T𝐈=0T_{{\mathbf{I}}}=0 if Δ(𝐈)=0\Delta({\mathbf{I}})=0. The universal Grassmann-Plücker function Δ^:ErPM+\widehat{\Delta}:E^{r}\to P_{M}^{+} for MM sends 𝐈Er{\mathbf{I}}\in E^{r} to T𝐈T_{\mathbf{I}} if |𝐈||{\mathbf{I}}| is a basis of MM, and to 0 otherwise. The universal PMP_{M}-matroid M^\widehat{M} for MM is defined by the Grassmann-Plücker function T𝐈1Δ^:ErPMT_{\mathbf{I}}^{-1}\widehat{\Delta}:E^{r}\to P_{M}, where 𝐈Er{\mathbf{I}}\in E^{r} is any rr-tuple with Δ(𝐈)0\Delta({\mathbf{I}})\neq 0.

Definition 3.8.

Let MM be a matroid with universal PMP_{M}-matroid M^\widehat{M}. Let (J;e1,,e4)ΩM(J;e_{1},\dots,e_{4})\in\Omega_{M} and (H1,,H4)ΘM(H_{1},\dots,H_{4})\in\Theta_{M}. The universal cross ratio of (J;e1,,e4)(J;e_{1},\dots,e_{4}) is the element

[e1e2e3e4]J:=[e1e2e3e4]M^,J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}\ :=\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\widehat{M},J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\widehat{M},J}}{}{}

of PMP_{M}, and the universal cross ratio of (H1,,H4)(H_{1},\dotsc,H_{4}) is the element

[H1H2H3H4]:=[H1H2H3H4]M^\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{}\ :=\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\widehat{M}}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}_{\widehat{M}}}{}{}

of PMP_{M}.

The relation between cross ratios of a PP-matroid and the universal cross ratio of the underlying matroid M¯\underline{M} is explained in the following statement.

Proposition 3.9.

Let PP be a pasture and MM a PP-matroid with Grassmann Plücker function Δ:ErP\Delta:E^{r}\to P. Let M¯\underline{M} be the underlying matroid and PM¯P_{\underline{M}} its universal pasture. Let χM:PM¯P\mychi_{M}:P_{\underline{M}}\to P be the universal morphism associated with MM, which maps T𝐈/T𝐈T_{\mathbf{I}}/T_{{\mathbf{I}}^{\prime}} to Δ(𝐈)/Δ(𝐈)\Delta({\mathbf{I}})/\Delta({\mathbf{I}}^{\prime}). Then

χM([e1e2e3e4]J)=[e1e2e3e4]M,J\mychi_{M}\bigg{(}\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}\bigg{)}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{M,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{M,J}}{}{}

as elements of PP for every (J;e1,,e4)ΩM¯(J;e_{1},\dotsc,e_{4})\in\Omega_{\underline{M}}.

Proof.

This follows directly from the definitions of χM\mychi_{M}, Δ^\widehat{\Delta} and the (universal) cross ratios. ∎

3.4. Fundamental elements

Universal cross ratios can be characterized intrinsically as the fundamental elements of the universal pasture of a matroid. To the best of our knowledge, the importance of fundamental elements in the study of matroid representations goes back to Semple’s paper [26], where this concept was introduced in the context of partial fields. We extend the notion of fundamental elements to pastures and explain its relation to universal cross ratios in the following.

The property of cross ratios that lead to the definition of fundamental elements are the 33-term Plücker relations

Δ(𝐉e1e2)Δ(𝐉e3e4)Δ(𝐉e1e3)Δ(𝐉e2e4)+Δ(𝐉e1e4)Δ(𝐉e2e3)= 0\Delta({\mathbf{J}}e_{1}e_{2})\Delta({\mathbf{J}}e_{3}e_{4})-\Delta({\mathbf{J}}e_{1}e_{3})\Delta({\mathbf{J}}e_{2}e_{4})+\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{2}e_{3})\ =\ 0

for a Grassmann-Plücker function Δ:ErP\Delta:E^{r}\to P, where 𝐉Er2{\mathbf{J}}\in E^{r-2} and e1,,e4Ee_{1},\dotsc,e_{4}\in E. If Δ(𝐉eiej)0\Delta({\mathbf{J}}e_{i}e_{j})\neq 0 for all distinct i,j{1,,4}i,j\in\{1,\dotsc,4\}, then division by Δ(𝐉e1e4)Δ(𝐉e2e3)-\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{2}e_{3}) yields

[e1e2e3e4]Δ,𝐉+[e1e3e2e4]Δ,𝐉=Δ(𝐉e1e3)Δ(𝐉e2e4)Δ(𝐉e1e4)Δ(𝐉e2e3)+Δ(𝐉e1e2)Δ(𝐉e3e4)Δ(𝐉e1e4)Δ(𝐉e3e2)= 1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\Delta,{\mathbf{J}}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\Delta,{\mathbf{J}}}}{}{}+\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{3}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{2}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\Delta,{\mathbf{J}}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{3}}\\ {e_{2}}&{e_{4}}\end{smallmatrix}\big{]}_{\Delta,{\mathbf{J}}}}{}{}\ =\ \frac{\Delta({\mathbf{J}}e_{1}e_{3})\Delta({\mathbf{J}}e_{2}e_{4})}{\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{2}e_{3})}+\frac{\Delta({\mathbf{J}}e_{1}e_{2})\Delta({\mathbf{J}}e_{3}e_{4})}{\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{3}e_{2})}\ =\ 1

for the non-degenerate cross ratios [e1e2e3e4]Δ,𝐉\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\Delta,{\mathbf{J}}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\Delta,{\mathbf{J}}}}{}{} and [e1e3e2e4]Δ,𝐉\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{3}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{2}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\Delta,{\mathbf{J}}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{3}}\\ {e_{2}}&{e_{4}}\end{smallmatrix}\big{]}_{\Delta,{\mathbf{J}}}}{}{} in P×P^{\times}.

Definition 3.10.

Let PP be a pasture. A fundamental element of PP is an element zP×z\in P^{\times} such that z+z=1z+z^{\prime}=1 for some zP×z^{\prime}\in P^{\times}.

Proposition 3.11.

Let MM be a matroid. For an element zPMz\in P_{M}, the following are equivalent:

  1. (1)

    zz is a fundamental element of PMP_{M};

  2. (2)

    z=[e1e2e3e4]Jz=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{} for some (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}};

  3. (3)

    z=[H1H2H3H4]z=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{} for some (H1,,H4)ΘM(H_{1},\dotsc,H_{4})\in\Theta_{M}^{\scalebox{0.7}{$\diamondsuit$}}.

Proof.

Our preceding discussion shows that [e1e2e3e4]J+[e1e3e2e4]J=1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}+\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{3}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{2}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{3}}\\ {e_{2}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}=1 for (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}}. Thus (2)\Rightarrow(1). The equivalence of (2) and (3) follows from Proposition 3.6.

We are left with (1)\Rightarrow(2). Assume that zPM×z\in P_{M}^{\times} is a fundamental element, i.e. z+z1=0z+z^{\prime}-1=0 for some zPM×z^{\prime}\in P_{M}^{\times}. Since the nullset of the extended universal pasture PM+P_{M}^{+} is generated by the 33-terms Plücker relations, there must be an element a(PM+)×a\in(P_{M}^{+})^{\times} such that az+aza=0az+az^{\prime}-a=0 is of the form

T𝐉e1e2T𝐉e3e4T𝐉e1e3T𝐉e2e4+T𝐉e1e4T𝐉e2e3= 0T_{{\mathbf{J}}e_{1}e_{2}}T_{{\mathbf{J}}e_{3}e_{4}}-T_{{\mathbf{J}}e_{1}e_{3}}T_{{\mathbf{J}}e_{2}e_{4}}+T_{{\mathbf{J}}e_{1}e_{4}}T_{{\mathbf{J}}e_{2}e_{3}}\ =\ 0

for some 𝐉Er2{\mathbf{J}}\in E^{r-2} and e1,,e4Ee_{1},\dotsc,e_{4}\in E such that |𝐉eiej||{\mathbf{J}}e_{i}e_{j}| is a basis of MM for all distinct i,j{1,,4}i,j\in\{1,\dotsc,4\}, i.e. (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}} where J=|𝐉|J=|{\mathbf{J}}|. After a suitable permutation of e1,,e4e_{1},\dotsc,e_{4}, we can assume that a=T𝐉e1e4T𝐉e2e3=a-a=T_{{\mathbf{J}}e_{1}e_{4}}T_{{\mathbf{J}}e_{2}e_{3}}=-a and az=T𝐉e1e3T𝐉e2e4az=-T_{{\mathbf{J}}e_{1}e_{3}}T_{{\mathbf{J}}e_{2}e_{4}}. Thus

z=aza=T𝐉e1e3T𝐉e2e4T𝐉e1e4T𝐉e2e3=[e1e2e3e4]Jz\ =\ \frac{-az}{-a}\ =\ \frac{T_{{\mathbf{J}}e_{1}e_{3}}T_{{\mathbf{J}}e_{2}e_{4}}}{T_{{\mathbf{J}}e_{1}e_{4}}T_{{\mathbf{J}}e_{2}e_{3}}}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}

is a cross ratio, as claimed. ∎

3.5. Compatibility with the Tutte group formulation of Dress and Wenzel

We provide a comparison of the different types of universal cross ratios, as introduced above, with the cross ratios introduced by Dress and Wenzel in [14, Def. 2.3].

The image of a universal cross ratio [e1e2e3e4]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{} under the isomorphism PM×𝕋MP_{M}^{\times}\to{\mathbb{T}}_{M} from Theorem 2.20 appears implicitly already in [13, Prop. 2.2], and is as follows.

Lemma 3.12.

Let MM be a matroid with Grassmann-Plücker function Δ:Er𝕂\Delta:E^{r}\to{\mathbb{K}}, Tutte group 𝕋M{\mathbb{T}}_{M} and universal pasture PMP_{M}. Let φ:PM×𝕋M\myvarphi:P_{M}^{\times}\to{\mathbb{T}}_{M} be the isomorphism of groups that sends T𝐈/T𝐈T_{\mathbf{I}}/T_{{\mathbf{I}}^{\prime}} to X𝐈/X𝐈X_{\mathbf{I}}/X_{{\mathbf{I}}^{\prime}} for 𝐈,𝐈supp(Δ){\mathbf{I}},{\mathbf{I}}^{\prime}\in\operatorname{supp}(\Delta). Then

φ([e1e2e3e4]J)=X𝐉e1e3X𝐉e2e4X𝐉e1e4X𝐉e2e3\myvarphi\bigg{(}\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}\bigg{)}\ =\ \frac{X_{{\mathbf{J}}e_{1}e_{3}}X_{{\mathbf{J}}e_{2}e_{4}}}{X_{{\mathbf{J}}e_{1}e_{4}}X_{{\mathbf{J}}e_{2}e_{3}}}

for all (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M} and 𝐉Er2{\mathbf{J}}\in E^{r-2} with |𝐉|=J|{\mathbf{J}}|=J.

Proof.

Note that the ratio (X𝐉e1e3X𝐉e2e4)(X𝐉e1e4X𝐉e2e3)1\big{(}X_{{\mathbf{J}}e_{1}e_{3}}X_{{\mathbf{J}}e_{2}e_{4}}\big{)}\big{(}{X_{{\mathbf{J}}e_{1}e_{4}}X_{{\mathbf{J}}e_{2}e_{3}}}\big{)}^{-1} does not depend on the ordering of 𝐉{\mathbf{J}}. The rest follows immediately from the definitions. ∎

Let (H1,,H4)(H_{1},\dotsc,H_{4}) be a modular quadruple of hyperplanes of MM and FF the corank 22 flat contained in all HiH_{i}. Let e3H3Fe_{3}\in H_{3}-F and e4H4Fe_{4}\in H_{4}-F. The Dress–Wenzel universal cross ratio of (H1,,H4)(H_{1},\dotsc,H_{4}) is the element

[H1H2H3H4]𝕋:=XH1,e3XH2,e4XH2,e3XH1,e4\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{{\mathbb{T}}}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}_{{\mathbb{T}}}}{}{}\ :=\ \frac{X_{H_{1},e_{3}}X_{H_{2},e_{4}}}{X_{H_{2},e_{3}}X_{H_{1},e_{4}}}

of the group 𝕋M{\mathbb{T}}_{M}^{\mathcal{H}}.

As shown in [14, Lemma 2.1], this definition is independent of the choices of e3e_{3} and e4e_{4}. Since deg([H1H2H3H4])=0\deg_{\mathcal{H}}\big{(}\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{{\mathcal{H}}}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}_{{\mathcal{H}}}}{}{}\big{)}=0, it follows from Theorem 1.3 that [H1H2H3H4]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{{\mathcal{H}}}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}_{{\mathcal{H}}}}{}{} is contained in the image of the injection ι:𝕋M𝕋M\myiota:{\mathbb{T}}_{M}\to{\mathbb{T}}_{M}^{\mathcal{H}}.

Lemma 3.13.

Let ψ:PM×𝕋M\mypsi:P_{M}^{\times}\to{\mathbb{T}}_{M}^{\mathcal{H}} be the group homomorphism that maps T𝐈eT𝐈e1T_{{\mathbf{I}}e}T_{{\mathbf{I}}e^{\prime}}^{-1} to XH,eXH,e1X_{H,e}X_{H,e^{\prime}}^{-1} where 𝐈Er1{\mathbf{I}}\in E^{r-1}, e,eEe,e^{\prime}\in E, I=|𝐈|I=|{\mathbf{I}}|, H=IH=\langle I\rangle, and Ie,IeIe,Ie^{\prime} are bases of MM. Let (H1,,H4)ΘM(H_{1},\dotsc,H_{4})\in\Theta_{M} be a modular quadruple of hyperplanes of MM. Then

ψ([H1H2H3H4])=[H1H2H3H4]𝕋.\mypsi\bigg{(}\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{}\bigg{)}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{{\mathbb{T}}}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}_{{\mathbb{T}}}}{}{}.
Proof.

It is clear from the definitions that ψ=ιφ\mypsi=\myiota\circ\myvarphi. By Lemma 3.5, there is an element (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M} with Ψ(J;e1,,e4)=(H1,,H4)\Psi(J;e_{1},\dotsc,e_{4})=(H_{1},\dotsc,H_{4}), i.e. Hi=JeiH_{i}=\langle Je_{i}\rangle for i=1,,4i=1,\dotsc,4. Using Proposition 3.6, we obtain

ψ([H1H2H3H4])=ιφ([e1e2e3e4]J)=ι(X𝐉e1e3X𝐉e2e4X𝐉e1e4X𝐉e2e3)=XH1,e3XH2,e4XH1,e4XH2,e3=[H1H2H3H4]𝕋\mypsi\bigg{(}\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{}\bigg{)}\,=\,\myiota\circ\myvarphi\bigg{(}\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}\bigg{)}\,=\,\myiota\Bigg{(}\frac{X_{{\mathbf{J}}e_{1}e_{3}}X_{{\mathbf{J}}e_{2}e_{4}}}{X_{{\mathbf{J}}e_{1}e_{4}}X_{{\mathbf{J}}e_{2}e_{3}}}\Bigg{)}\,=\,\frac{X_{H_{1},e_{3}}X_{H_{2},e_{4}}}{X_{H_{1},e_{4}}X_{H_{2},e_{3}}}\,=\,\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{{\mathbb{T}}}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}_{{\mathbb{T}}}}{}{}

as claimed. ∎

4. Foundations

The foundation FMF_{M} of a matroid MM is the subpasture of degree 0-elements of the universal pasture PMP_{M}, and it represents the functor taking a pasture PP to the set of PP-rescaling classes of MM. In particular, just as with PMP_{M}, the foundation can detect whether or not a matroid is representable over a given pasture PP in terms of the existence of a morphism from FMF_{M} to PP.

One advantage of the foundation over the universal pasture is that, because of some deep theorems due to Tutte, Dress–Wenzel, and Gelfand–Rybnikov–Stone, there is an explicit presentation of FMF_{M} in terms of generators and relations in which the relations are all inherited from “small” embedded minors. More precisely, the foundation of MM is generated by the universal cross ratios of MM, and all relations between these cross ratios are generated by a small list of relations stemming from embedded minors of MM having at most 77 elements.

We begin our discussion of foundations by reviewing some facts which were proved in the authors’ previous paper [5]. Next we explain the role of embedded minors in the study of foundations. We then exhibit, through very explicit computations, the relations between universal cross ratios inherited from small minors which enter into the presentation by generators and relations alluded to above. Finally, we use the aforementioned result of Gelfand, Rybnikov and Stone to prove that these relations generate all relations in FMF_{M} between universal cross ratios.

4.1. Definition and basic facts

Let MM be a matroid of rank rr on EE with extended universal pasture PM+P_{M}^{+}. For a subset II of EE, let δI:E\mydelta_{I}:E\to{\mathbb{Z}} be the characteristic function of II, which is an element of E{\mathbb{Z}}^{E}. The multidegree is the group homomorphism

degE:(PM+)×ET𝐈δI,\begin{array}[]{cccc}\deg_{E}:&(P_{M}^{+})^{\times}&\longrightarrow&{\mathbb{Z}}^{E}\\ &T_{\mathbf{I}}&\longmapsto&\mydelta_{I},\end{array}

where I=|𝐈|I=|{\mathbf{I}}|. It is easily verified that this map is well-defined, cf. [5, section 7.3]. The degree in ii is the function degi:(PM+)×\deg_{i}:(P_{M}^{+})^{\times}\to{\mathbb{Z}} that is the composition of degE:(PM+)×E\deg_{E}:(P_{M}^{+})^{\times}\to{\mathbb{Z}}^{E} with the canonical projection to the ii-th component, i.e. degi(T𝐈)=1\deg_{i}(T_{\mathbf{I}})=1 if iIi\in I and degi(T𝐈)=0\deg_{i}(T_{\mathbf{I}})=0 if iIi\notin I. The total degree is the function deg:(PM+)×\deg:(P_{M}^{+})^{\times}\to{\mathbb{Z}} that is the sum over degi\deg_{i} for all iEi\in E, i.e. deg(T𝐈)=iEdegi(T𝐈)=#I=r\deg(T_{\mathbf{I}})=\sum_{i\in E}\deg_{i}(T_{\mathbf{I}})=\#I=r.

Definition 4.1.

Let MM be a matroid with extended universal pasture PM+P_{M}^{+}. The foundation of MM is the subpasture FMF_{M} of PM+P_{M}^{+} that consists of 0 and all elements of multidegree 0.

Note that the universal pasture PMP_{M} of MM is the subpasture of PM+P_{M}^{+} that is generated by all units of total degree 0. Since deg(x)=0\deg(x)=0 if degE(x)=0\deg_{E}(x)=0, the foundation FMF_{M} of MM is a subpasture of PMP_{M}.

The relevance of the foundation of MM is the fact that it represents the rescaling class space

𝒳RM(P)={rescaling classes of M over P}{\mathcal{X}}^{R}_{M}(P)\ =\ \Big{\{}\text{rescaling classes of $M$ over $P$}\Big{\}}

considered as a functor in PP.

Theorem 4.2 ([5, Cor. 7.26]).

Let MM be a matroid and PP a pasture. Then there is a functorial bijection 𝒳RM(P)=Hom(FM,P){\mathcal{X}}^{R}_{M}(P)=\operatorname{Hom}(F_{M},P). In particular, MM is representable over PP if and only if there is a morphism FMPF_{M}\to P.

Recall from [13] that the inner Tutte group 𝕋(0)M{\mathbb{T}}^{(0)}_{M} of a matroid MM is defined as the subgroup of the Tutte group 𝕋M{\mathbb{T}}_{M} of MM that consists of all elements of multidegree 0, where the multidegree deg:𝕋ME\deg:{\mathbb{T}}_{M}\to{\mathbb{Z}}^{E} is defined in the same way as the multidegree deg:PME\deg:P_{M}\to{\mathbb{Z}}^{E}. This yields at once the following consequence of Theorem 2.20 (cf. [5, Cor. 7.11]).

Corollary 4.3.

The canonical isomorphism M×𝕋M{\mathbb{P}}_{M}^{\times}\to{\mathbb{T}}_{M} restricts to an isomorphism FM×𝕋(0)MF_{M}^{\times}\to{\mathbb{T}}^{(0)}_{M}.

Remark 4.4.

Wenzel observes in [34, Thm. 6.3] that a matroid representation over a fuzzy ring KK induces a group homomorphism 𝕋(0)MK×{\mathbb{T}}^{(0)}_{M}\to K^{\times}, and that this homomorphism detects the rescaling class of a representation. This can be seen as a partial analogue of Theorem 4.2 for fuzzy rings (cf. Remark 2.21).

4.2. Universal cross ratios as generators of the foundation

Let MM be a matroid of rank rr on EE and PM+P_{M}^{+} its extended universal pasture. The simplest type of elements of PM+P_{M}^{+} with multidegree 0 are universal cross ratios

[e1e2e3e4]J=T𝐉e1e3T𝐉e2e4T𝐉e1e4T𝐉e2e3\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}\ =\ \frac{T_{{\mathbf{J}}e_{1}e_{3}}T_{{\mathbf{J}}e_{2}e_{4}}}{T_{{\mathbf{J}}e_{1}e_{4}}T_{{\mathbf{J}}e_{2}e_{3}}}

where (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M} and 𝐉Er2{\mathbf{J}}\in E^{r-2} such that |𝐉|=J|{\mathbf{J}}|=J. This formula shows that the universal cross ratios are elements of the foundation FMF_{M} of MM. It is proven in [5, Cor. 7.11] that the foundation is generated by the universal cross ratios. To summarize, we have:

Theorem 4.5.

Let MM be a matroid. Then [e1e2e3e4]JFM×\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}\in F_{M}^{\times} for every (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M}, and FM×F_{M}^{\times} is generated by the collection of all such universal cross ratios.

Using Proposition 3.6, we obtain:

Corollary 4.6.

Let MM be a matroid. Then [H1H2H3H4]FM×\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{}\in F_{M}^{\times} for every (H1,,H4)ΘM(H_{1},\ldots,H_{4})\in\Theta_{M}, and FM×F_{M}^{\times} is generated by the collection of all such hyperplane universal cross ratios.

4.3. The foundation of the dual matroid

Let MM be a matroid of rank rr on EE and PMP_{M} its universal pasture. By definition the identity morphism id:PMPM\textup{id}:P_{M}\to P_{M} is the characteristic morphism of the universal PMP_{M}-matroid M^\widehat{M}; cf. Theorem 2.18. The underlying matroid of M^\widehat{M} is M^¯=M\underline{\widehat{M}}=M. The underlying matroid of the dual PMP_{M}-matroid M^\widehat{M}^{\ast} of M^\widehat{M} is the dual M^¯=M\underline{\widehat{M}^{\ast}}=M^{\ast} of MM, cf. [3, Thm. 3.24]. Let ωM:PMPM\myomega_{M}:P_{M^{\ast}}\to P_{M} be the characteristic morphism of M^\widehat{M}^{\ast}.

Proposition 4.7.

Let MM be a matroid of rank rr on EE. Then ωM:PMPM\myomega_{M}:P_{M^{\ast}}\to P_{M} is an isomorphism of pastures that restricts to an isomorphism FMFMF_{M^{\ast}}\to F_{M} between the respective foundations of MM^{\ast} and MM. Let n=#En=\#E. For every 𝐈Enr1{\mathbf{I}}\in E^{n-r-1}, 𝐉Er1{\mathbf{J}}\in E^{r-1} and e,fEe,f\in E such that E=|𝐈||𝐉|{e,f}E=|{\mathbf{I}}|\cup|{\mathbf{J}}|\cup\{e,f\}, we have

ωM(T𝐈eT𝐈f)=T𝐉fT𝐉e,\myomega_{M}\bigg{(}\frac{T_{{\mathbf{I}}e}}{T_{{\mathbf{I}}f}}\bigg{)}\ =\ -\frac{T_{{\mathbf{J}}f}}{T_{{\mathbf{J}}e}},

and for every (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M} and I=EJe1e4I=E-Je_{1}\dotsc e_{4}, we have (I;e1,,e4)ΩM(I;e_{1},\dotsc,e_{4})\in\Omega_{M^{\ast}} and

ωM([e1e2e3e4]M^,I)=[e1e2e3e4]M^,J,\myomega_{M}\bigg{(}\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\widehat{M^{\ast}},I}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\widehat{M^{\ast}},I}}{}{}\bigg{)}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\widehat{M},J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\widehat{M},J}}{}{},

where M^\widehat{M} is the universal PMP_{M}-matroid of MM and M^\widehat{M^{\ast}} is the universal PMP_{M^{\ast}}-matroid of MM^{\ast}.

Proof.

The construction of ωM\myomega_{M}, applied to MM^{\ast} in place of MM, yields a morphism ωM:PMPM\myomega_{M^{\ast}}:P_{M^{\ast\ast}}\to P_{M^{\ast}}. Since M=MM^{\ast\ast}=M, we have PM=PMP_{M^{\ast\ast}}=P_{M}. The composition ωMωM:PM=PMPMPM\myomega_{M}\circ\myomega_{M^{\ast}}:P_{M}=P_{M^{\ast\ast}}\to P_{M^{\ast}}\to P_{M} is the characteristic morphism of the double dual M^\widehat{M}^{\ast\ast} of M^\widehat{M}, which is equal to M^\widehat{M} by [3, Thm. 3.24], and thus ωMωM\myomega_{M}\circ\myomega_{M^{\ast}} is the identity of PMP_{M}. Similarly, the composition ωMωM\myomega_{M^{\ast}}\circ\myomega_{M} is the identity of PMP_{M^{\ast}}. This shows that ωM\myomega_{M} and ωM\myomega_{M^{\ast}} are mutually inverse isomorphisms.

Let Δ:ErPM\Delta:E^{r}\to P_{M} be a Grassmann-Plücker function for M^\widehat{M}. Endow EE with a total order and define sign(i1,,in)=sign(π)\operatorname{sign}(i_{1},\dotsc,i_{n})=\operatorname{sign}(\mypi) as the sign of the permutation π\mypi of EE such that π(i1)<<π(in)\mypi(i_{1})<\dotsb<\mypi(i_{n}) if i1,,inEi_{1},\dotsc,i_{n}\in E are pairwise distinct. Then by [3, Lemma 4.1], there is a Grassmann-Plücker function Δ:Enr1PM\Delta^{\ast}:E^{n-r-1}\to P_{M} for M^\widehat{M}^{\ast} that satisfies

Δ(i1,,inr)=sign(i1,,in)Δ(inr+1,,in)\Delta^{\ast}(i_{1},\dotsc,i_{n-r})\ =\ \operatorname{sign}(i_{1},\dotsc,i_{n})\Delta(i_{n-r+1},\dotsc,i_{n})

for all pairwise distinct i1,,inEi_{1},\dotsc,i_{n}\in E. Thus if 𝐈=(i1,,inr1){\mathbf{I}}=(i_{1},\dotsc,i_{n-r-1}), 𝐉=(j1,,jr1){\mathbf{J}}=(j_{1},\dotsc,j_{r-1}) and e,fEe,f\in E are as in the hypothesis of the theorem, then

ωM(T𝐈eT𝐈f)=Δ(𝐈e)Δ(𝐈f)=sign(i1,,inr1,e,j1,,jr1,f)Δ(𝐉f)sign(i1,,inr1,f,j1,,jr1,e)Δ(𝐉e)=T𝐉fT𝐉e,\myomega_{M}\bigg{(}\frac{T_{{\mathbf{I}}e}}{T_{{\mathbf{I}}f}}\bigg{)}\ =\ \frac{\Delta^{\ast}({\mathbf{I}}e)}{\Delta^{\ast}({\mathbf{I}}f)}\ =\ \frac{\operatorname{sign}(i_{1},\dotsc,i_{n-r-1},e,j_{1},\dotsc,j_{r-1},f)\Delta({\mathbf{J}}f)}{\operatorname{sign}(i_{1},\dotsc,i_{n-r-1},f,j_{1},\dotsc,j_{r-1},e)\Delta({\mathbf{J}}e)}\ =\ -\frac{T_{{\mathbf{J}}f}}{T_{{\mathbf{J}}e}},

as claimed. If (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M} and I=EJe1e4I=E-Je_{1}\dotsc e_{4}, then JeiekJe_{i}e_{k} is a basis for MM, and thus IejelIe_{j}e_{l} is a basis for MM^{\ast} for all i,j{1,2}i,j\in\{1,2\} and k,l{3,4}k,l\in\{3,4\}. Thus (I;e1,,e4)ΩM(I;e_{1},\dotsc,e_{4})\in\Omega_{M^{\ast}}. The image of the corresponding cross ratio under ωM\myomega_{M} is

ωM([e1e2e3e4]I)=Δ(𝐈e1e3)Δ(𝐈e2e4)Δ(𝐈e1e4)Δ(𝐈e2e3)=[e1e2e3e4]M^,I=[e1e2e3e4]M^,J\myomega_{M}\bigg{(}\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{I}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{I}}{}{}\bigg{)}\ =\ \frac{\Delta^{\ast}({\mathbf{I}}e_{1}e_{3})\Delta^{\ast}({\mathbf{I}}e_{2}e_{4})}{\Delta^{\ast}({\mathbf{I}}e_{1}e_{4})\Delta^{\ast}({\mathbf{I}}e_{2}e_{3})}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\widehat{M}^{\ast},I}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\widehat{M}^{\ast},I}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\widehat{M},J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\widehat{M},J}}{}{}

where 𝐈Enr2{\mathbf{I}}\in E^{n-r-2} such that |𝐈|=I|{\mathbf{I}}|=I and where we use Lemma 3.2 for the last equality. Since the foundations of MM and MM^{\ast} are generated by cross ratios, it follows at once that ωM\myomega_{M} restricts to an isomorphism FMFMF_{M^{\ast}}\to F_{M}. ∎

4.4. Foundations of embedded minors

Let MM be a matroid of rank rr on EE, and let M^\widehat{M} be the universal PMP_{M}-matroid associated with MM, whose characteristic function is the identity map on PMP_{M}; cf. Theorem 2.18. Let Δ:ErPM\Delta:E^{r}\to P_{M} be a Grassmann-Plücker function for M^\widehat{M}; e.g. we can choose some 𝐈0Er{\mathbf{I}}_{0}\in E^{r} such that |𝐈0||{\mathbf{I}}_{0}| is a basis of MM and define Δ(𝐈)=T𝐈/T𝐈0\Delta({\mathbf{I}})=T_{\mathbf{I}}/T_{{\mathbf{I}}_{0}} if |𝐈||{\mathbf{I}}| is a basis of MM and Δ(𝐈)=0\Delta({\mathbf{I}})=0 if not.

Let N=M\I/JN=M\backslash I/J be an embedded minor of MM. Let ss be its rank and EN=E(IJ)E_{N}=E-(I\cup J) its ground set. Choose an ordering J={js+1,,jr}J=\{j_{s+1},\dotsc,j_{r}\} of the elements of JJ. By [3, Lemma 4.4], the function

Δ\I/J:EsNPM𝐈Δ(𝐈js+1jr)\begin{array}[]{cccc}\Delta\backslash I/J:&E^{s}_{N}&\longrightarrow&P_{M}\\ &{\mathbf{I}}&\longmapsto&\Delta({\mathbf{I}}j_{s+1}\dotsc j_{r})\end{array}

is a Grassmann-Plücker function that represents N=M\I/JN=M\backslash I/J and its isomorphism class N^=M^\I/J\widehat{N}=\widehat{M}\backslash I/J is independent of the choice of ordering of JJ. The characteristic function of the PMP_{M}-matroid N^\widehat{N} is a morphism ψM\I/J:PNPM\mypsi_{M\backslash I/J}:P_{N}\to P_{M}; once again cf. Theorem 2.18.

Proposition 4.8.

Let MM be a matroid of rank rr on EE and N=M\I/JN=M\backslash I/J an embedded minor of rank ss on EN=E(IJ)E_{N}=E-(I\cup J). Let J={js+1,,jr}J=\{j_{s+1},\dotsc,j_{r}\}. Then the morphism ψM\I/J:PNPM\mypsi_{M\backslash I/J}:P_{N}\to P_{M} satisfies the following properties.

  1. (1)

    For all 𝐈,𝐉ENs{\mathbf{I}},{\mathbf{J}}\in E_{N}^{s} such that |𝐈||{\mathbf{I}}| and |𝐉||{\mathbf{J}}| are bases of NN, we have

    ψM\I/J(T𝐈T𝐉)=T𝐈js+1jrT𝐉js+1jr.\mypsi_{M\backslash I/J}\bigg{(}\frac{T_{{\mathbf{I}}}}{T_{{\mathbf{J}}}}\bigg{)}\ =\ \frac{T_{{\mathbf{I}}j_{s+1}\dotsc j_{r}}}{T_{{\mathbf{J}}j_{s+1}\dotsc j_{r}}}.
  2. (2)

    The identification N=M\J/IN^{\ast}=M^{\ast}\backslash J/I yields a commutative diagram

    PN{P_{N^{\ast}}}PM{P_{M^{\ast}}}PN{P_{N}}PM{P_{M}}ψM\J/I\scriptstyle{\mypsi_{M^{\ast}\backslash J/I}}ωN\scriptstyle{\myomega_{N}}ωM\scriptstyle{\myomega_{M}}ψM\I/J\scriptstyle{\mypsi_{M\backslash I/J}}

    of pastures, where ωN\myomega_{N} and ωM\myomega_{M} are the isomorphisms from Proposition 4.7.

  3. (3)

    The morphism ψM\I/J:PNPM\mypsi_{M\backslash I/J}:P_{N}\to P_{M} restricts to a morphism φM\I/J:FNFM\myvarphi_{M\backslash I/J}:F_{N}\to F_{M} between the foundations of NN and MM. For (J;e1,,e4)ΩN(J^{\prime};e_{1},\dotsc,e_{4})\in\Omega_{N}, we have (JJ;e1,,e4)ΩM(J^{\prime}\cup J;e_{1},\dotsc,e_{4})\in\Omega_{M} and

    φM\I/J([e1e2e3e4]J)=[e1e2e3e4]JJ.\myvarphi_{M\backslash I/J}\bigg{(}\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{}\bigg{)}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J\cup J^{\prime}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J\cup J^{\prime}}}{}{}.
  4. (4)

    If every element in II is a loop and if every element in JJ is a coloop, then ψM\I/J\mypsi_{M\backslash I/J} is an isomorphism. If every element in II is a loop or parallel to an element in ENE_{N} and if every element in JJ is a coloop or coparallel to an element in ENE_{N}, then φM\I/J\myvarphi_{M\backslash I/J} is an isomorphism.

Proof.

Property (1) follows from the direct computation

ψM\I/J(T𝐈T𝐉)=Δ\I/J(T𝐈)Δ\I/J(T𝐉)=T𝐈js+1jrT𝐉js+1jr.\mypsi_{M\backslash I/J}\bigg{(}\frac{T_{{\mathbf{I}}}}{T_{{\mathbf{J}}}}\bigg{)}\ =\ \frac{\Delta\backslash I/J(T_{\mathbf{I}})}{\Delta\backslash I/J(T_{\mathbf{J}})}\ =\ \frac{T_{{\mathbf{I}}j_{s+1}\dotsc j_{r}}}{T_{{\mathbf{J}}j_{s+1}\dotsc j_{r}}}.

We continue with (2). Let rr^{\ast} be the corank of MM and ss^{\ast} the corank of NN. Choose an ordering I={is+1,,ir}I=\{i_{s^{\ast}+1},\dotsc,i_{r^{\ast}}\}. Let 𝐈ENs1{\mathbf{I}}\in E_{N}^{s^{\ast}-1}, 𝐉ENs1{\mathbf{J}}\in E_{N}^{s-1} and e,fENe,f\in E_{N} be such that EN=|𝐈||𝐉|{e,f}E_{N}=|{\mathbf{I}}|\cup|{\mathbf{J}}|\cup\{e,f\}, which are the assumptions needed to apply Proposition 4.7 to ωN\myomega_{N}. Since PNP_{N^{\ast}} is generated by elements of the form T𝐈e/T𝐈fT_{{\mathbf{I}}e}/T_{{\mathbf{I}}f}, the commutativity of the diagram in question follows from

ψM\I/JωN(T𝐈eT𝐈f)=ψM\I/J(T𝐉fT𝐉e)=T𝐉fjs+1jrT𝐉ejs+1jr=ωM(T𝐈eis+1irT𝐈fis+1ir)=ωMψM\J/I(T𝐈eT𝐈f).\mypsi_{M\backslash I/J}\circ\myomega_{N}\bigg{(}\frac{T_{{\mathbf{I}}e}}{T_{{\mathbf{I}}f}}\bigg{)}\ =\ \mypsi_{M\backslash I/J}\bigg{(}-\frac{T_{{\mathbf{J}}f}}{T_{{\mathbf{J}}e}}\bigg{)}\ =\ -\frac{T_{{\mathbf{J}}fj_{s+1}\dotsc j_{r}}}{T_{{\mathbf{J}}ej_{s+1}\dotsc j_{r}}}\\ =\ \myomega_{M}\bigg{(}\frac{T_{{\mathbf{I}}ei_{s^{\ast}+1}\dotsc i_{r^{\ast}}}}{T_{{\mathbf{I}}fi_{s^{\ast}+1}\dotsc i_{r^{\ast}}}}\bigg{)}\ =\ \myomega_{M}\circ\mypsi_{M^{\ast}\backslash J/I}\bigg{(}\frac{T_{{\mathbf{I}}e}}{T_{{\mathbf{I}}f}}\bigg{)}.

Note that we can apply Proposition 4.7 to ωM\myomega_{M} since E=|𝐈||𝐉|{e,f}IJE=|{\mathbf{I}}|\cup|{\mathbf{J}}|\cup\{e,f\}\cup I\cup J.

We continue with (3). If (J;e1,,e4)ΩN(J^{\prime};e_{1},\dotsc,e_{4})\in\Omega_{N}, then for all i{1,2}i\in\{1,2\} and k{3,4}k\in\{3,4\}, the set JeiekJ^{\prime}e_{i}e_{k} is a basis of NN and thus JJ{ei,ek}J^{\prime}\cup J\cup\{e_{i},e_{k}\} is a basis of MM. Thus (JJ;e1,,e4)ΩM(J^{\prime}\cup J;e_{1},\dotsc,e_{4})\in\Omega_{M}. Let 𝐉ENs{\mathbf{J}}^{\prime}\in E_{N}^{s} such that |𝐉|=J|{\mathbf{J}}^{\prime}|=J^{\prime}. Then

ψM\I/J([e1e2e3e4]J)=ΔN(T𝐉e1e3T𝐉e2e4T𝐉e1e4T𝐉e2e3)=T𝐉e1e3js+1jrT𝐉e2e4js+1jrT𝐉e1e4js+1jrT𝐉e2e3js+1jr=[e1e2e3e4]JJ.\mypsi_{M\backslash I/J}\bigg{(}\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{}\bigg{)}=\Delta_{N}\bigg{(}\frac{T_{{\mathbf{J}}^{\prime}e_{1}e_{3}}T_{{\mathbf{J}}^{\prime}e_{2}e_{4}}}{T_{{\mathbf{J}}^{\prime}e_{1}e_{4}}T_{{\mathbf{J}}^{\prime}e_{2}e_{3}}}\bigg{)}=\frac{T_{{\mathbf{J}}^{\prime}e_{1}e_{3}j_{s+1}\dotsc j_{r}}T_{{\mathbf{J}}^{\prime}e_{2}e_{4}j_{s+1}\dotsc j_{r}}}{T_{{\mathbf{J}}^{\prime}e_{1}e_{4}j_{s+1}\dotsc j_{r}}T_{{\mathbf{J}}^{\prime}e_{2}e_{3}j_{s+1}\dotsc j_{r}}}=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J\cup J^{\prime}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J\cup J^{\prime}}}{}{}.

By Theorem 4.5, the foundation of a matroid is generated by its cross ratios. Thus the previous calculation shows that ψM\I/J\mypsi_{M\backslash I/J} restricts to a morphism φM\I/J:FNFM\myvarphi_{M\backslash I/J}:F_{N}\to F_{M} which maps [e1e2e3e4]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{} to [e1e2e3e4]JJ\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}\cup J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J^{\prime}\cup J}}{}{}.

We continue with (4). By successively deleting or contracting one element at a time, it suffices to prove the claim for #(IJ)=1\#(I\cup J)=1. Using (2), we can assume that I={e}I=\{e\} and J=J=\varnothing. If ee is a loop, then III^{\prime}\mapsto I^{\prime} defines a bijection between the set of bases IEN=E{e}I^{\prime}\subset E_{N}=E-\{e\} of NN and the set of bases of MM. Moreover, for every (J;e1,,e4)ΩM(J^{\prime};e_{1},\dotsc,e_{4})\in\Omega_{M}, we have eJe1e4e\notin J^{\prime}e_{1}\dotsc e_{4}, which provides an identification ΩN=ΩM\Omega_{N}=\Omega_{M}. Thus PNP_{N} and PMP_{M} have the same generators and the same 33-term Plücker relations, so ψM\I/J:PNPM\mypsi_{M\backslash I/J}:P_{N}\to P_{M} is an isomorphism. This argument also shows that φM\I/J:FNFM\myvarphi_{M\backslash I/J}:F_{N}\to F_{M} is an isomorphism.

If ee is parallel to an element fENf\in E_{N}, then Je=Jf\langle J^{\prime}e\rangle=\langle J^{\prime}f\rangle for every subset JJ^{\prime} of ENE_{N}. Thus for e1,,e4Ee_{1},\dotsc,e_{4}\in E and f1,,f4ENf_{1},\dotsc,f_{4}\in E_{N} with either ei=fie_{i}=f_{i} or ei=ee_{i}=e and fi=ff_{i}=f for i=1,,4i=1,\dotsc,4, we have (J;e1,,e4)ΩM(J^{\prime};e_{1},\dotsc,e_{4})\in\Omega_{M} if and only if (J;f1,,f4)ΩN(J^{\prime};f_{1},\dotsc,f_{4})\in\Omega_{N}, and φM\I/J([f1f2f3f4]J)=[e1e2e3e4]J\myvarphi_{M\backslash I/J}\Big{(}\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$f_{1}$}}&{\scalebox{0.9}{$f_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$f_{3}$}}&{{\scalebox{0.9}{$f_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{f_{1}}&{f_{2}}\\ {f_{3}}&{f_{4}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{}\Big{)}=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{}. This shows that φM\I/J:FNFM\myvarphi_{M\backslash I/J}:F_{N}\to F_{M} is an isomorphism, which completes the proof. ∎

An immediate consequence of Proposition 4.8 is the following.

Corollary 4.9.

The foundation of a matroid is isomorphic to the foundation of its simplification and isomorphic to the foundation of its cosimplification.

Proof.

This follows at once from Proposition 4.8, since the simplification of a matroid MM is an embedded minor of MM of the form M\IM\backslash I, where II consists of all loops of MM and a choice of all but one element in each class of parallel elements. Similarly, the cosimplification of MM is an embedded minor of MM of the form M/JM/J, where JJ consists of all coloops of MM and a choice of all but one element in each class of coparallel elements. ∎

Another consequence of Proposition 4.8, which we will utilize constantly in the upcoming sections, is the following observation. Since a universal cross ratio [e1e2e3e4]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{} involves only bases Jeiek{Je_{i}e_{k}} that contain JJ and have a trivial intersection with I=EJe1e2e3e4I=E-Je_{1}e_{2}e_{3}e_{4}, we have

[e1e2e3e4]J=T𝐉e1e3T𝐉e2e4T𝐉e1e4T𝐉e2e3=φ(T(e1,e3))φ(T(e2,e4))φ(T(e1,e4))φ(T(e2,e3))=φ([e1e2e3e4])\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}\ =\ \frac{T_{{\mathbf{J}}e_{1}e_{3}}T_{{\mathbf{J}}e_{2}e_{4}}}{T_{{\mathbf{J}}e_{1}e_{4}}T_{{\mathbf{J}}e_{2}e_{3}}}\ =\ \frac{\myvarphi(T_{(e_{1},e_{3})})\ \myvarphi(T_{(e_{2},e_{4})})}{\myvarphi(T_{(e_{1},e_{4})})\ \myvarphi(T_{(e_{2},e_{3})})}\ =\ \myvarphi\bigg{(}\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\varnothing}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\varnothing}}{}{}\bigg{)}

for the morphism φ=φM\I/J:FM\I/JFM\myvarphi=\myvarphi_{M\backslash I/J}:F_{M\backslash I/J}\to F_{M} from Proposition 4.8. Thus every universal cross ratio in FMF_{M} is the image of a universal cross ratio of an embedded minor N=M\I/JN=M\backslash I/J of rank 22 on a 44-element set {e1,e2,e3,e4}=E(IJ)\{e_{1},e_{2},e_{3},e_{4}\}=E-(I\cup J).

4.5. The foundation of U24U^{2}_{4}

Let M=U24M=U^{2}_{4} be the uniform minor of rank 22 on the set E={1,,4}E=\{1,\dotsc,4\}, which is represented by the Grassmann-Plücker function Δ:E2𝕂\Delta:E^{2}\to{\mathbb{K}} with Δ(i,j)=1δi,j\Delta(i,j)=1-\mydelta_{i,j}. The cross ratios of MM are of the form

[e1e2e3e4]:=[e1e2e3e4]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}}{}{}\ :=\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\varnothing}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\varnothing}}{}{}

for some permutation e:ieie:i\mapsto e_{i} of EE. Since permuting columns and rows in [e1e2e3e4]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}}{}{} does not change the cross ratio, as pointed out in section 3.1, we have

[1234]=[2143]=[3412]=[4321].\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$2$}}&{\scalebox{0.9}{$1$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{2}&{1}\\ {4}&{3}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$3$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$1$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{3}&{4}\\ {1}&{2}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$4$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$1$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{4}&{3}\\ {2}&{1}\end{smallmatrix}\big{]}}{}{}.

Thus we can assume that e1=1e_{1}=1, and with this convention, we find that each of the 24 possible cross ratios is equal to one of the following six:

[1234],[1243],[1324],[1342],[1423],[1432].\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{},\quad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {4}&{3}\end{smallmatrix}\big{]}}{}{},\quad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {2}&{4}\end{smallmatrix}\big{]}}{}{},\quad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {4}&{2}\end{smallmatrix}\big{]}}{}{},\quad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{4}\\ {2}&{3}\end{smallmatrix}\big{]}}{}{},\quad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{4}\\ {3}&{2}\end{smallmatrix}\big{]}}{}{}.

They satisfy the following two types of multiplicative relations

[1243]=[1234]1,[1243]=[1234]1,[1243]=[1234]1;\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {4}&{3}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{}^{-1},\qquad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {4}&{3}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{}^{-1},\qquad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {4}&{3}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{}^{-1};
[1234][1342][1423]=1,[1243][1324][1432]=1;\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {4}&{2}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{4}\\ {2}&{3}\end{smallmatrix}\big{]}}{}{}\ =\ -1,\qquad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {4}&{3}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {2}&{4}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{4}\\ {3}&{2}\end{smallmatrix}\big{]}}{}{}\ =\ -1;

and the Plücker relations

[1234]+[1324]=1,[1342]+[1432]=1,[1423]+[1243]=1.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{}+\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {2}&{4}\end{smallmatrix}\big{]}}{}{}=1,\quad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {4}&{2}\end{smallmatrix}\big{]}}{}{}+\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{4}\\ {3}&{2}\end{smallmatrix}\big{]}}{}{}=1,\quad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{4}\\ {2}&{3}\end{smallmatrix}\big{]}}{}{}+\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {4}&{3}\end{smallmatrix}\big{]}}{}{}=1.

These relations can be illustrated in the form of a hexagon, see Figure 1. The three edges with label * refer to relations of type (4.5), the three edges with label ++ refer to the Plücker relations (4.5), and the two inner triangles refer to the relations of type (4.5).

𝟏\mathbf{-1\ }[1234]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{}[1324]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {2}&{4}\end{smallmatrix}\big{]}}{}{}[1342]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {4}&{2}\end{smallmatrix}\big{]}}{}{}[1432]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{4}\\ {3}&{2}\end{smallmatrix}\big{]}}{}{}[1423]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{4}\\ {2}&{3}\end{smallmatrix}\big{]}}{}{}[1243]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {4}&{3}\end{smallmatrix}\big{]}}{}{}++*++*++*
Figure 1. The hexagon of cross ratios of U24U^{2}_{4}

Note that we can rewrite the relations of type (4.5) as [1234][1243]=1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {4}&{3}\end{smallmatrix}\big{]}}{}{}=1, and so forth, which highlights an analogy with the Plücker relations [1234]+[1324]=1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{}+\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {2}&{4}\end{smallmatrix}\big{]}}{}{}=1. This makes the meaning of the edge labels \ast and ++ easy to remember.

Proposition 4.10.

Let x=[1234]x=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{} and y=[1324]y=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {2}&{4}\end{smallmatrix}\big{]}}{}{}. Then the foundation of M=U24M=U^{2}_{4} is

FM=𝕌=𝔽1±x,y{x+y1}.F_{M}\ =\ {\mathbb{U}}\ =\ {{\mathbb{F}}_{1}^{\pm}}\langle x,y\rangle\!\sslash\!\{x+y-1\}.

In particular, we have

[1243]=x1,[1342]=y1,[1432]=xy1,[1423]=x1y.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\ }}{\big{[}\begin{smallmatrix}{1}&{2}\\ {4}&{3}\end{smallmatrix}\big{]}_{\ }}{}{}=\ x^{-1},\qquad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\ }}{\big{[}\begin{smallmatrix}{1}&{3}\\ {4}&{2}\end{smallmatrix}\big{]}_{\ }}{}{}=\ y^{-1},\qquad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\ }}{\big{[}\begin{smallmatrix}{1}&{4}\\ {3}&{2}\end{smallmatrix}\big{]}_{\ }}{}{}=\ -xy^{-1},\qquad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\ }}{\big{[}\begin{smallmatrix}{1}&{4}\\ {2}&{3}\end{smallmatrix}\big{]}_{\ }}{}{}=\ -x^{-1}y.
Proof.

By relation (4.5), FMF_{M} is generated by the 66 cross ratios

x=[1234],y=[1324],[1243],[1342],[1432],[1423].x\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{},\quad y\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {2}&{4}\end{smallmatrix}\big{]}}{}{},\quad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {4}&{3}\end{smallmatrix}\big{]}}{}{},\quad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {4}&{2}\end{smallmatrix}\big{]}}{}{},\quad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{4}\\ {3}&{2}\end{smallmatrix}\big{]}}{}{},\quad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{4}\\ {2}&{3}\end{smallmatrix}\big{]}}{}{}.

By relation (4.5), we have

[1243]=[1234]1=x1and[1342]=[1324]1=y1.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {4}&{3}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}^{-1}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}^{-1}}{}{}\ =\ x^{-1}\qquad\text{and}\qquad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {4}&{2}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}^{-1}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {2}&{4}\end{smallmatrix}\big{]}^{-1}}{}{}\ =\ y^{-1}.

Relation (4.5), paired with (4.5), yields

[1432]=[1423]1=[1234][1342]=xy1.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{4}\\ {3}&{2}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}^{-1}}{\big{[}\begin{smallmatrix}{1}&{4}\\ {2}&{3}\end{smallmatrix}\big{]}^{-1}}{}{}\ =\ -\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{}\ \cdot\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {4}&{2}\end{smallmatrix}\big{]}}{}{}\ =\ -xy^{-1}.

Applying (4.5) once again yields

[1423]=[1432]1=x1y.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{4}\\ {2}&{3}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$2$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}^{-1}}{\big{[}\begin{smallmatrix}{1}&{4}\\ {3}&{2}\end{smallmatrix}\big{]}^{-1}}{}{}\ =\ -x^{-1}y.

By (4.5), we have x+y1=0x+y-1=0. This shows that the foundation FMF_{M} of M=U24M=U^{2}_{4} is a quotient of 𝕌=𝔽1±x,y{x+y1}{\mathbb{U}}={{\mathbb{F}}_{1}^{\pm}}\langle x,y\rangle\!\sslash\!\{x+y-1\}.

There are several different ways to show that there are no further relations in FMF_{M} aside from those already present in 𝕌{\mathbb{U}}, for example:

  1. (1)

    One can work this out “by hand”.

  2. (2)

    One can utilize the fact that U24U^{2}_{4} is near-regular, which implies that there is a morphism FM𝕌F_{M}\to{\mathbb{U}}.

  3. (3)

    One can apply Theorem 4.20, whose proof does not rely on Proposition 4.10.

We explain a fourth route, which uses a theorem of Dress and Wenzel determining the inner Tutte group of a uniform matroid. In the case of M=U24M=U^{2}_{4}, [13, Thm. 8.1], paired with Corollary 4.3, shows that FM×𝕋(0)(/2)×2𝕌×F_{M}^{\times}\simeq{\mathbb{T}}^{(0)}\simeq({\mathbb{Z}}/2{\mathbb{Z}})\times{\mathbb{Z}}^{2}\simeq{\mathbb{U}}^{\times}. We conclude that the quotient map 𝕌FM{\mathbb{U}}\to F_{M} is an isomorphism between the underlying monoids. We are left with showing that every relation in the nullset of FMF_{M} comes from 𝕌{\mathbb{U}}, which is the intersection of the nullset NPM+N_{P_{M}^{+}} of PM+P_{M}^{+} with Sym3(FM)\operatorname{Sym}^{3}(F_{M}). Since NPM+N_{P_{M}^{+}} is generated by the single term

T1,2T3,4T1,3T2,4+T1,4T2,3=T1,4T2,3(x+y1),T_{1,2}T_{3,4}\ -\ T_{1,3}T_{2,4}\ +\ T_{1,4}T_{2,3}\ =\ -T_{1,4}T_{2,3}\cdot(x+y-1),

where we use the short-hand notation Ti,j=T(i,j)T_{i,j}=T_{(i,j)}, every term in NFMN_{F_{M}} is a multiple of x+y1x+y-1. This shows that 𝕌FM{\mathbb{U}}\to F_{M} is an isomorphism. ∎

Morphisms from 𝕌{\mathbb{U}} into another pasture can be studied in terms of pairs of fundamental elements:

Definition 4.11.

A pair of fundamental elements in PP is an ordered pair (z,z)(z,z^{\prime}) of elements z,zP×z,z^{\prime}\in P^{\times} such that z+z=1z+z^{\prime}=1.

Lemma 4.12.

Let PP be a pasture. Then there is a bijection between Hom(𝕌,P)\operatorname{Hom}({\mathbb{U}},P) with the set of pairs of fundamental elements.

Proof.

Every morphism f:𝕌=𝔽1±x,y{x+y=1}Pf:{\mathbb{U}}={{\mathbb{F}}_{1}^{\pm}}\langle x,y\rangle\!\sslash\!\{x+y=1\}\to P maps xx and yy to invertible elements in PP. Since x+y=1x+y=1, we have f(x)+f(y)=1f(x)+f(y)=1 in PP, which shows that (f(x),f(y))\big{(}f(x),f(y)\big{)} is a pair of fundamental elements. This defines a map Φ:Hom(𝕌,P)P\Phi:\operatorname{Hom}({\mathbb{U}},P)\to{\mathcal{F}}_{P}, where P{\mathcal{F}}_{P} is the set of pairs of fundamental elements in PP.

Since ff is determined by the images of xx and yy, we see that Φ\Phi is injective. On the other hand, for every pair (u,v)(u,v) of fundamental elements in PP, the map xux\mapsto u and yvy\mapsto v extends to a morphism f:𝕌Pf:{\mathbb{U}}\to P. Thus Φ\Phi is surjective as well. ∎

Recall that a reorientation class is a rescaling class over the sign hyperfield 𝕊{\mathbb{S}}. The following corollary is well known:

Corollary 4.13.

The rescaling classes of U24U^{2}_{4} over a field kk are in bijection with k{0,1}k-\{0,1\}, and U24U^{2}_{4} has 33 reorientation classes.

Proof.

If P=kP=k is a field, then y=1xy=1-x is uniquely determined by xx, and x,yx,y both belong to k×k^{\times} precisely when xk{0,1}x\in k-\{0,1\}, which establishes the first claim. The second claim follows from the observation that a+b=1a+b=1 in 𝕊{\mathbb{S}} if and only if (a,b)(a,b) is one of the 33 pairs (1,1)(1,1), (1,1)(1,-1) and (1,1)(-1,1). ∎

4.6. The tip and cotip relations

In this section, we exhibit two types of relations that occur for matroids of ranks 22 and 33, respectively, on the five element set E={1,,5}E=\{1,\dotsc,5\}.

As in the case of the uniform matroid U24U^{2}_{4}, we write [ijkl]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$k$}}&{{\scalebox{0.9}{$l$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {k}&{l}\end{smallmatrix}\big{]}}{}{} for [ijkl]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$k$}}&{{\scalebox{0.9}{$l$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\varnothing}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {k}&{l}\end{smallmatrix}\big{]}_{\varnothing}}{}{} in the case of a rank 22-matroid MM. We also use the shorthand notation Ti,j=T(i,j)T_{i,j}=T_{(i,j)} and Ti,j,k=T(i,j,k)T_{i,j,k}=T_{(i,j,k)}.

Lemma 4.14.

Let MM be a matroid of rank 22 on E={1,,5}E=\{1,\dotsc,5\}. Assume that {i,j}\{i,j\} is a basis of MM for all i{1,2}i\in\{1,2\} and all j{3,4,5}j\in\{3,4,5\}. Then

[1234][1245][1253]= 1.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {4}&{5}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$5$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {5}&{3}\end{smallmatrix}\big{]}}{}{}\ =\ 1.
Proof.

Equation (4.14) follows from the direct computation

[1234][1245][1253]=T1,3T2,4T1,4T2,3T1,4T2,5T1,5T2,4T1,5T2,3T1,3T2,5= 1.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {4}&{5}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$5$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {5}&{3}\end{smallmatrix}\big{]}}{}{}\ =\ \frac{T_{1,3}T_{2,4}}{T_{1,4}T_{2,3}}\cdot\frac{T_{1,4}T_{2,5}}{T_{1,5}T_{2,4}}\cdot\frac{T_{1,5}T_{2,3}}{T_{1,3}T_{2,5}}\ =\ 1.\qed

We call equation (4.14) the tip relation with tip {1,2}\{1,2\} and cyclic orientation (3,4,5)(3,4,5). The reason for this terminology is that in the case of the uniform matroid M=U25M=U^{2}_{5}, the three cross ratios in equation (4.14) stem from three octahedrons in the basis exchange graph of MM, which share exactly one common vertex, or tip, which is {1,2}\{1,2\}.

Note that if MM is not uniform, i.e. some 22-subsets {i,j}\{i,j\} of EE are not bases, then some of the cross ratios in equation (4.14) are trivial. We will examine this situation in more detail in section 5.1.

In the case of a matroid of rank 33, we write [ijkl]m\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$k$}}&{{\scalebox{0.9}{$l$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{m}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {k}&{l}\end{smallmatrix}\big{]}_{m}}{}{} for [ijkl]{m}\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$k$}}&{{\scalebox{0.9}{$l$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\{m\}}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {k}&{l}\end{smallmatrix}\big{]}_{\{m\}}}{}{}.

Lemma 4.15.

Let MM be a matroid of rank 33 on E={1,,5}E=\{1,\dotsc,5\}. Assume that {i,j,k}\{i,j,k\} is a basis of MM for all i{1,2}i\in\{1,2\} and all j,k{3,4,5}j,k\in\{3,4,5\} with jkj\neq k. Then

[1234]5[1245]3[1253]4= 1.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{5}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}_{5}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{3}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {4}&{5}\end{smallmatrix}\big{]}_{3}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$5$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{4}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {5}&{3}\end{smallmatrix}\big{]}_{4}}{}{}\ =\ 1.
Proof.

Equation (4.15) follows from the direct computation

[1234]5[1245]3[1253]4=T5,1,3T5,2,4T5,1,4T5,2,3T3,1,4T3,2,5T3,1,5T3,2,4T4,1,5T4,2,3T4,1,3T4,2,5=T4,1,5T4,1,5T3,2,5T3,2,5T5,1,3T5,1,3T4,2,3T4,2,3T3,1,4T3,1,4T5,2,4T5,2,4=(1)6= 1.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{5}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}_{5}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{3}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {4}&{5}\end{smallmatrix}\big{]}_{3}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$5$}}&{{\scalebox{0.9}{$3$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{4}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {5}&{3}\end{smallmatrix}\big{]}_{4}}{}{}\ =\ \frac{T_{5,1,3}\cdot T_{5,2,4}}{T_{5,1,4}\cdot T_{5,2,3}}\ \cdot\ \frac{T_{3,1,4}\cdot T_{3,2,5}}{T_{3,1,5}\cdot T_{3,2,4}}\ \cdot\ \frac{T_{4,1,5}\cdot T_{4,2,3}}{T_{4,1,3}\cdot T_{4,2,5}}\\ =\ \frac{T_{4,1,5}}{-T_{4,1,5}}\cdot\frac{T_{3,2,5}}{-T_{3,2,5}}\cdot\frac{T_{5,1,3}}{-T_{5,1,3}}\cdot\frac{T_{4,2,3}}{-T_{4,2,3}}\cdot\frac{T_{3,1,4}}{-T_{3,1,4}}\cdot\frac{T_{5,2,4}}{-T_{5,2,4}}\ =\ (-1)^{6}\ =\ 1.\mbox{\qed}

We call equation (4.15) the cotip relation with cotip {1,2}\{1,2\} and cyclic orientation (3,4,5)(3,4,5). Similar to the rank 22-case, we use this terminology since in the case of the uniform matroid M=U35M=U^{3}_{5}, the three cross ratios in equation (4.15) stem from three octahedrons in the basis exchange graph of MM, which share exactly one common vertex, which is {3,4,5}\{3,4,5\}. Therefore we call the complement {1,2}\{1,2\} of this common vertex the cotip.

Note that the tip and cotip relations are both invariant under permuting {1,2}\{1,2\} and under cyclic permutations of (3,4,5)(3,4,5). Any other permutation of EE will produce another tip or cotip relation, provided that all involved values of Δ\Delta are nonzero.

4.7. Relations for parallel elements

In this section, we exhibit a type of relation between universal cross ratios that stems from parallel elements. As in the previous section, we write [1234]5\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{5}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}_{5}}{}{} for [1234]{5}\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\{5\}}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}_{\{5\}}}{}{}.

Lemma 4.16.

Let MM be a matroid of rank 33 on E={1,,6}E=\{1,\dotsc,6\} and assume that 55 and 66 are parallel elements, i.e. {5,6}\{5,6\} is a circuit of MM. If ({k};1,,4)ΩM(\{k\};1,\dotsc,4)\in\Omega_{M} for k=5,6k=5,6, then

[1234]5=[1234]6.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{5}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}_{5}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{6}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}_{6}}{}{}.
Proof.

By our assumptions, every subset of the form {i,j,k}\{i,j,k\} with i{1,2}i\in\{1,2\}, j{3,4}j\in\{3,4\} and k{5,6}k\in\{5,6\} is a basis of MM, but no basis contains both 55 and 66. Thus ({1};3,4,6,5)(\{1\};3,4,6,5) and ({2};3,4,5,6)(\{2\};3,4,5,6) are degenerate tuples in ΩM\Omega_{M}, and thus [3465]1=[3456]2=1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$3$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$6$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{1}}{\big{[}\begin{smallmatrix}{3}&{4}\\ {6}&{5}\end{smallmatrix}\big{]}_{1}}{}{}=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$3$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$5$}}&{{\scalebox{0.9}{$6$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{2}}{\big{[}\begin{smallmatrix}{3}&{4}\\ {5}&{6}\end{smallmatrix}\big{]}_{2}}{}{}=1. With this, equation (4.16) follows from the computation

[1234]5=[1234]5[3465]1[3456]2=T5,1,3T5,2,4T5,1,4T5,2,3T1,3,6T1,4,5T1,3,5T1,4,6T2,3,5T2,4,6T2,3,6T2,4,5=T1,4,5T1,4,5T2,3,5T2,3,5T5,1,3T5,1,3T6,1,3T6,2,4T6,1,4T6,2,3T5,2,4T5,2,4=[1234]6.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{5}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}_{5}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{5}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}_{5}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$3$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$6$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{1}}{\big{[}\begin{smallmatrix}{3}&{4}\\ {6}&{5}\end{smallmatrix}\big{]}_{1}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$3$}}&{\scalebox{0.9}{$4$}}\\[-2.0pt] {\scalebox{0.9}{$5$}}&{{\scalebox{0.9}{$6$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{2}}{\big{[}\begin{smallmatrix}{3}&{4}\\ {5}&{6}\end{smallmatrix}\big{]}_{2}}{}{}\ =\ \frac{T_{5,1,3}\cdot T_{5,2,4}}{T_{5,1,4}\cdot T_{5,2,3}}\ \cdot\ \frac{T_{1,3,6}\cdot T_{1,4,5}}{T_{1,3,5}\cdot T_{1,4,6}}\ \cdot\ \frac{T_{2,3,5}\cdot T_{2,4,6}}{T_{2,3,6}\cdot T_{2,4,5}}\\ =\ \frac{T_{1,4,5}}{T_{1,4,5}}\ \cdot\ \frac{T_{2,3,5}}{T_{2,3,5}}\ \cdot\ \frac{T_{5,1,3}}{T_{5,1,3}}\ \cdot\ \frac{T_{6,1,3}\cdot T_{6,2,4}}{T_{6,1,4}\cdot T_{6,2,3}}\ \cdot\ \frac{T_{5,2,4}}{T_{5,2,4}}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{6}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}_{6}}{}{}.\mbox{\qed}

4.8. The foundation of the Fano matroid and its dual

In this section, we show that the Fano matroid F7F_{7} and its dual F7F_{7}^{\ast} impose the relation 1=1-1=1 on their foundation, which is 𝔽2{\mathbb{F}}_{2}. This already follows from [5, Thms. 7.30 and 7.33], using the fact that F7F_{7} and F7F_{7}^{\ast} are not regular. Here we offer a proof in terms of a direct calculation that does not rely on knowledge of the representability of F7F_{7}.

The Fano matroid F7F_{7} is the rank 33 matroid on E={1,,7}E=\{1,\dotsc,7\} represented by the Grassmann-Plücker function Δ:E3𝕂\Delta:E^{3}\to{\mathbb{K}} with Δ(i,i+1,i+3)=0\Delta(i,i+1,i+3)=0 for iEi\in E, where we read i+1i+1 and i+3i+3 modulo 77, and Δ(i,j,k)=1\Delta(i,j,k)=1 otherwise. Thus the family of circuits is {{i,i+1,i+3}|iE}\Big{\{}\{i,i+1,i+3\}\,\Big{|}\,i\in E\Big{\}}, together with all 44-element subsets that do not contain one of these, which can be illustrated as follows:

11223344556677
Lemma 4.17.

The foundation of both the Fano matroid F7F_{7} and its dual F7F_{7}^{\ast} is 𝔽2{\mathbb{F}}_{2}.

Proof.

Since the foundation of F7F_{7}^{\ast} is isomorphic to the foundation of F7F_{7}, it is enough to prove the lemma for the Fano matroid. Throughout the proof, we read expressions like i+ki+k and iki-k modulo 77 for all i,kEi,k\in E.

Since the rank of F7F_{7} is 33, the set JJ of a tuple (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M} is a singleton, i.e. J={j}J=\{j\} for some jEj\in E. The element jj is contained in the three circuits C1={j,j+1,j+3}C_{1}=\{j,j+1,j+3\}, C2={j1,j,j+2}C_{2}=\{j-1,j,j+2\} and C3={j3,j2,j}C_{3}=\{j-3,j-2,j\} whose union is equal to EE. By the pigeonhole principle, we must have ek,elCie_{k},e_{l}\in C_{i} for some ii and klk\neq l. Since j,ek,elj,e_{k},e_{l} are pairwise distinct, Ci={j,ek,el}C_{i}=\{j,e_{k},e_{l}\} is not a basis. This shows that every (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M} is degenerate, and thus [e1e2e3e4]J=1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}=1. We conclude that FMF_{M} is a quotient of 𝔽1±{{\mathbb{F}}_{1}^{\pm}}.

We use the shorthand notations [ijkl]m=[ijkl]{m}\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$k$}}&{{\scalebox{0.9}{$l$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{m}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {k}&{l}\end{smallmatrix}\big{]}_{m}}{}{}=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$k$}}&{{\scalebox{0.9}{$l$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\{m\}}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {k}&{l}\end{smallmatrix}\big{]}_{\{m\}}}{}{} and Tij,k,l=T(i+j,i+k,i+l)T^{i}_{j,k,l}=T_{(i+j,i+k,i+l)} in the following. Note that Timj+m,k+m,l+m=Tij,k,lT^{i-m}_{j+m,k+m,l+m}=T^{i}_{j,k,l} and Tiσ(j),σ(k),σ(l)=sign(σ)Tij,k,lT^{i}_{\mysigma(j),\mysigma(k),\mysigma(l)}=\operatorname{sign}(\mysigma)T^{i}_{j,k,l} for every permutation σ\mysigma of {j,k,l}\{j,k,l\}. We calculate that

1\displaystyle 1\quad =i=17[i+1i+3i+2i+4]i[i+2i+6i+5i+4]i\displaystyle=\quad\prod_{i=1}^{7}\quad\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i+1$}}&{\scalebox{0.9}{$i+3$}}\\[-2.0pt] {\scalebox{0.9}{$i+2$}}&{{\scalebox{0.9}{$i+4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{i}}{\big{[}\begin{smallmatrix}{i+1}&{i+3}\\ {i+2}&{i+4}\end{smallmatrix}\big{]}_{i}}{}{}\ \cdot\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i+2$}}&{\scalebox{0.9}{$i+6$}}\\[-2.0pt] {\scalebox{0.9}{$i+5$}}&{{\scalebox{0.9}{$i+4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{i}}{\big{[}\begin{smallmatrix}{i+2}&{i+6}\\ {i+5}&{i+4}\end{smallmatrix}\big{]}_{i}}{}{}
=i=17Ti0,1,2Ti0,3,4Ti0,1,4Ti0,3,2Ti0,2,5Ti0,6,4Ti0,2,4Ti0,6,5\displaystyle=\quad\prod_{i=1}^{7}\quad\frac{T^{i}_{0,1,2}\cdot T^{i}_{0,3,4}}{T^{i}_{0,1,4}\cdot T^{i}_{0,3,2}}\ \cdot\ \frac{T^{i}_{0,2,5}\cdot T^{i}_{0,6,4}}{T^{i}_{0,2,4}\cdot T^{i}_{0,6,5}}
=i=17Ti0,1,2Ti0,3,4Ti0,2,5Ti0,6,4Ti33,4,0Ti44,0,6Ti55,0,2Ti22,1,0\displaystyle=\quad\prod_{i=1}^{7}\quad\frac{T^{i}_{0,1,2}\cdot T^{i}_{0,3,4}\cdot T^{i}_{0,2,5}\cdot T^{i}_{0,6,4}}{T^{i-3}_{3,4,0}\cdot T^{i-4}_{4,0,6}\cdot T^{i-5}_{5,0,2}\cdot T^{i-2}_{2,1,0}}
=i=17Ti0,3,4Ti30,3,4Ti0,6,4Ti40,6,4Ti0,2,5Ti50,2,5Ti0,1,2Ti20,1,2\displaystyle=\quad\prod_{i=1}^{7}\quad\frac{T^{i}_{0,3,4}}{T^{i-3}_{0,3,4}}\ \cdot\ \frac{T^{i}_{0,6,4}}{T^{i-4}_{0,6,4}}\ \cdot\ \frac{T^{i}_{0,2,5}}{T^{i-5}_{0,2,5}}\ \cdot\ \frac{T^{i}_{0,1,2}}{-T^{i-2}_{0,1,2}}
=(1)7=1.\displaystyle=\quad(-1)^{7}\quad=\quad-1.

This shows that the foundation FMF_{M} of F7F_{7} is a quotient of 𝔽2=𝔽1±{1=1}{\mathbb{F}}_{2}={{\mathbb{F}}_{1}^{\pm}}\!\sslash\!\{-1=1\}. Since F7F_{7} does not contain any U24U^{2}_{4}-minors, all cross ratios are degenerate and thus the nullset of FMF_{M} does not contain any 33-term relations. We conclude that FM=𝔽2F_{M}={\mathbb{F}}_{2}. ∎

4.9. A presentation of the foundation by hyperplanes

Gelfand, Rybnikov and Stone exhibit in [16, Thm. 4] a complete set of multiplicative relations in the inner Tutte group of MM between the cross ratios [C1C2C3C4]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$C_{1}$}}&{\scalebox{0.9}{$C_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$C_{3}$}}&{{\scalebox{0.9}{$C_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{C_{1}}&{C_{2}}\\ {C_{3}}&{C_{4}}\end{smallmatrix}\big{]}}{}{} of modular quadruples (C1,,C4)(C_{1},\dotsc,C_{4}) of circuits, which results in essence from Tutte’s homotopy theorem. Since hyperplanes are just complements of circuits of the dual matroid, this set of relations yields at once a complete set of relations for cross ratios [H1H2H3H4]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{} of modular quadruples (H1,,H4)(H_{1},\dotsc,H_{4}) of hyperplanes.

Theorem 4.18.

Let MM be a matroid with foundation FMF_{M}. Then

FM=𝔽1±[H1H2H3H4]|(H1,,H4)ΘMS,\textstyle F_{M}\ =\ {{\mathbb{F}}_{1}^{\pm}}\,\big{\langle}\,\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{}\,\big{|}\,(H_{1},\dotsc,H_{4})\in\Theta_{M}\,\big{\rangle}\,\sslash\,S,

where SS is defined by the multiplicative relations

(1)2= 1,and1=1(-1)^{2}\ =\ 1,\qquad\text{and}\qquad-1=1

if the Fano matroid F7F_{7} or its dual F7F_{7}^{\ast} is a minor of MM;

[H1H2H3H4]=[H2H1H4H3]=[H3H4H1H2]=[H4H3H2H1]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{2}$}}&{\scalebox{0.9}{$H_{1}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{4}$}}&{{\scalebox{0.9}{$H_{3}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{2}}&{H_{1}}\\ {H_{4}}&{H_{3}}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{3}$}}&{\scalebox{0.9}{$H_{4}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{1}$}}&{{\scalebox{0.9}{$H_{2}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{3}}&{H_{4}}\\ {H_{1}}&{H_{2}}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{4}$}}&{\scalebox{0.9}{$H_{3}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{2}$}}&{{\scalebox{0.9}{$H_{1}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{4}}&{H_{3}}\\ {H_{2}}&{H_{1}}\end{smallmatrix}\big{]}}{}{}

for all (H1,,H4)Θ(H_{1},\dotsc,H_{4})\in\Theta_{\mathcal{H}}^{\scalebox{0.7}{$\diamondsuit$}};

[H1H2H3H4]= 1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{}\ =\ 1

for all degenerate (H1,,H4)Θ(H_{1},\dotsc,H_{4})\in\Theta_{\mathcal{H}};

[H1H2H4H3]=[H1H2H3H4]1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{4}$}}&{{\scalebox{0.9}{$H_{3}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{4}}&{H_{3}}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}^{-1}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}^{-1}}{}{}

for all (H1,,H4)Θ(H_{1},\dotsc,H_{4})\in\Theta_{\mathcal{H}}^{\scalebox{0.7}{$\diamondsuit$}};

[H1H2H3H4][H1H3H4H2][H1H4H2H3]=1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{3}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{4}$}}&{{\scalebox{0.9}{$H_{2}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{3}}\\ {H_{4}}&{H_{2}}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{4}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{2}$}}&{{\scalebox{0.9}{$H_{3}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{4}}\\ {H_{2}}&{H_{3}}\end{smallmatrix}\big{]}}{}{}\ =\ -1

for all (H1,,H4)Θ(H_{1},\dotsc,H_{4})\in\Theta_{\mathcal{H}}^{\scalebox{0.7}{$\diamondsuit$}};

[H1H2H3H4][H1H2H4H5][H1H2H5H3]= 1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{4}$}}&{{\scalebox{0.9}{$H_{5}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{4}}&{H_{5}}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{5}$}}&{{\scalebox{0.9}{$H_{3}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{5}}&{H_{3}}\end{smallmatrix}\big{]}}{}{}\ =\ 1

for all (H1,H2,H3,H4),(H1,H2,H4,H5),(H1,H2,H5,H3)Θ(H_{1},H_{2},H_{3},H_{4}),(H_{1},H_{2},H_{4},H_{5}),(H_{1},H_{2},H_{5},H_{3})\in\Theta_{\mathcal{H}}^{\scalebox{0.7}{$\diamondsuit$}}; and

[H15H25H35H45][H13H23H43H53][H14H24H54H34]= 1,\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{15}$}}&{\scalebox{0.9}{$H_{25}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{35}$}}&{{\scalebox{0.9}{$H_{45}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{15}}&{H_{25}}\\ {H_{35}}&{H_{45}}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{13}$}}&{\scalebox{0.9}{$H_{23}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{43}$}}&{{\scalebox{0.9}{$H_{53}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{13}}&{H_{23}}\\ {H_{43}}&{H_{53}}\end{smallmatrix}\big{]}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{14}$}}&{\scalebox{0.9}{$H_{24}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{54}$}}&{{\scalebox{0.9}{$H_{34}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{14}}&{H_{24}}\\ {H_{54}}&{H_{34}}\end{smallmatrix}\big{]}}{}{}\ =\ 1,

where Hij=FiFjH_{ij}=\langle F_{i}\cup F_{j}\rangle for five pairwise distinct corank 22-flats F1,,F5F_{1},\dotsc,F_{5} that contain a common flat of corank 33 such that (H15,H25,H35,H45),(H14,H24,H54,H34)Θ(H_{15},H_{25},H_{35},H_{45}),(H_{14},H_{24},H_{54},H_{34})\in\Theta_{\mathcal{H}}^{\scalebox{0.7}{$\diamondsuit$}} and (H13,H23,H43,H53)Θ(H_{13},H_{23},H_{43},H_{53})\in\Theta_{\mathcal{H}},
as well as the additive Plücker relations

[H1H2H3H4]+[H1H3H2H4]= 1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{}+\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{3}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{2}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{3}}\\ {H_{2}}&{H_{4}}\end{smallmatrix}\big{]}}{}{}\ =\ 1

for all (J;e1,,e4)ΘM(J;e_{1},\dotsc,e_{4})\in\Theta_{M}^{\scalebox{0.7}{$\diamondsuit$}}.

Proof.

The theorem follows by translating the relations between cross ratios [C1C2C3C4]𝕋\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$C_{1}$}}&{\scalebox{0.9}{$C_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$C_{3}$}}&{{\scalebox{0.9}{$C_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{{\mathbb{T}}}}{\big{[}\begin{smallmatrix}{C_{1}}&{C_{2}}\\ {C_{3}}&{C_{4}}\end{smallmatrix}\big{]}_{{\mathbb{T}}}}{}{} in 𝕋(0)M{\mathbb{T}}^{(0)}_{M^{\ast}} for modular quadruples of cycles of the dual matroid MM^{\ast} from [16, Thm. 4] to the hyperplane formulation by replacing a cocycle CC by the hyperplane H=ECH=E-C. To pass from the inner Tutte group to the foundation, we employ Lemma 3.13, which identifies [H1H2H3H4]𝕋\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{{\mathbb{T}}}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}_{{\mathbb{T}}}}{}{} with [H1H2H3H4]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{} under the canonical isomorphism M×𝕋(0)M{\mathbb{P}}_{M}^{\times}\to{\mathbb{T}}^{(0)}_{M}.

Using this translation, relation (4.18) is equivalent to (TG0) and (CR5) in [16]. Relation (4.18) is equivalent to (CR3). Relation (4.18) is equivalent to (CR1). Relation (CR4) is equivalent to (4.18) (in the case that one cross ratio is degenerate) and (4.18) (in the case that all cross ratios are non-degenerate). Relation (4.18) is equivalent to (CR4). Relation (4.18) is equivalent to (CR6), where we observe that the degenerate case L=LL=L^{\prime} in [16] reduces (CR6) to (CR1). Finally note that the 33-term Plücker relations of FMF_{M} are captured in (4.18). ∎

Remark 4.19.

We include a discussion of relation (4.18), which has the most complicated formulation among the relations of Theorem 4.18. Since all flats contain a common flat of corank 33, this constellation comes from a minor of rank 33, which has 55 corank 22-flats corresponding to F1,,F5F_{1},\dotsc,F_{5}. In the non-degenerate situation where all hyperplanes HijH_{ij} are pairwise distinct, this minor is of type U35U^{3}_{5}, and the containment relation of the FiF_{i} and HijH_{ij} can be illustrated as on the right-hand side of Figure 2.

The original formulation of Gelfand, Rybnikov and Stone concerns points, which are circuits, and lines, which are unions of circuits having projective dimension 1. To pass from our formulation to that of Gelfand-Rybnikov-Stone, we take the complement of a hyperplane HijH_{ij}, which is a circuit CijC_{ij} of the dual matroid. Thus, in the non-degenerate case, axiom (CR6) expresses the point-line configuration of U25U^{2}_{5}, which we illustrate on the left-hand side of Figure 2. The lines LiL_{i} are the complements of the flats FiF_{i}, and therefore the union of the circuits CijC_{ij} (with varying jj).

Note that there are two degenerate situations that (CR6) allows for: (a) three lines, say L1L_{1}, L2L_{2} and L3L_{3}, intersect in one point C12=C13=C23C_{12}=C_{13}=C_{23}; this case corresponds to the point-line arrangement of a parallel extension of U24U^{2}_{4}, which we denote by C5C_{5}^{\ast} in section 5.1.3; and (b) two lines agree; this case corresponds to the point-line arrangement of U24U^{2}_{4}.

C25C_{25}C13C_{13}C24C_{24}C35C_{35}C14C_{14}C12C_{12}C23C_{23}C34C_{34}C45C_{45}C15C_{15}L1L_{1}L2L_{2}L3L_{3}L4L_{4}L5L_{5}
H34H_{34}H45H_{45}H15H_{15}H12H_{12}H23H_{23}F4F_{4}F5F_{5}F1F_{1}F2F_{2}F3F_{3}H35H_{35}H14H_{14}H25H_{25}H13H_{13}H24H_{24}
Figure 2. Point-line configuration for U25U^{2}_{5} and flat configuration for U35U^{3}_{5}

4.10. A presentation of the foundation by bases

Using the relation between cross ratios [H1H2H3H4]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{} for modular quadruples (H1,,H4)(H_{1},\dotsc,H_{4}) of hyperplanes and universal cross ratios [e1e2e3e4]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{} for (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M}, as exhibited in Proposition 3.6, we derive from Theorem 4.18 the following description of a complete set of relations between universal cross ratios. The possibility of such a deduction was observed and communicated to us by Rudi Pendavingh, who proves a similar result in the joint work [10] in progress with Brettell.

Theorem 4.20.

Let MM be a matroid with foundation FMF_{M}. Then

FM=𝔽1±[e1e2e3e4]J|(J;e1,,e4)ΩMS,\textstyle F_{M}\ =\ {{\mathbb{F}}_{1}^{\pm}}\,\big{\langle}\,\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}\,\big{|}\,(J;e_{1},\dotsc,e_{4})\in\Omega_{M}\,\big{\rangle}\,\sslash\,S,

where SS is defined by the multiplicative relations

1=1-1=1

if the Fano matroid F7F_{7} or its dual F7F_{7}^{\ast} is a minor of MM;

[e1e2e3e4]J=[e2e1e4e3]J=[e3e4e1e2]J=[e4e3e2e1]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{2}$}}&{\scalebox{0.9}{$e_{1}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{4}$}}&{{\scalebox{0.9}{$e_{3}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{2}}&{e_{1}}\\ {e_{4}}&{e_{3}}\end{smallmatrix}\big{]}_{J}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{3}$}}&{\scalebox{0.9}{$e_{4}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{1}$}}&{{\scalebox{0.9}{$e_{2}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{3}}&{e_{4}}\\ {e_{1}}&{e_{2}}\end{smallmatrix}\big{]}_{J}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{4}$}}&{\scalebox{0.9}{$e_{3}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{2}$}}&{{\scalebox{0.9}{$e_{1}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{4}}&{e_{3}}\\ {e_{2}}&{e_{1}}\end{smallmatrix}\big{]}_{J}}{}{}

for all (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}};

[e1e2e3e4]J= 1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}\ =\ 1

for all degenerate (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M};

[e1e2e4e3]J=[e1e2e3e4]J1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{4}$}}&{{\scalebox{0.9}{$e_{3}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{4}}&{e_{3}}\end{smallmatrix}\big{]}_{J}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}^{-1}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}^{-1}}{}{}

for all (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}};

[e1e2e3e4]J[e1e3e4e2]J[e1e4e2e3]J=1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{3}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{4}$}}&{{\scalebox{0.9}{$e_{2}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{3}}\\ {e_{4}}&{e_{2}}\end{smallmatrix}\big{]}_{J}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{4}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{2}$}}&{{\scalebox{0.9}{$e_{3}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{4}}\\ {e_{2}}&{e_{3}}\end{smallmatrix}\big{]}_{J}}{}{}\ =\ -1

for all (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}};

[e1e2e3e4]J[e1e2e4e5]J[e1e2e5e3]J= 1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{4}$}}&{{\scalebox{0.9}{$e_{5}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{4}}&{e_{5}}\end{smallmatrix}\big{]}_{J}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{5}$}}&{{\scalebox{0.9}{$e_{3}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{5}}&{e_{3}}\end{smallmatrix}\big{]}_{J}}{}{}\ =\ 1

for all e1,,e5Ee_{1},\dotsc,e_{5}\in E and JEJ\subset E such that each of (J;e1,e2,e3,e4)(J;e_{1},e_{2},e_{3},e_{4}), (J;e1,e2,e4,e5)(J;e_{1},e_{2},e_{4},e_{5}) and (J;e1,e2,e5,e3)(J;e_{1},e_{2},e_{5},e_{3}) is in ΩM\Omega_{M};

[e1e2e3e4]Je5[e1e2e4e5]Je3[e1e2e5e3]Je4= 1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{Je_{5}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{Je_{5}}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{4}$}}&{{\scalebox{0.9}{$e_{5}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{Je_{3}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{4}}&{e_{5}}\end{smallmatrix}\big{]}_{Je_{3}}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{5}$}}&{{\scalebox{0.9}{$e_{3}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{Je_{4}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{5}}&{e_{3}}\end{smallmatrix}\big{]}_{Je_{4}}}{}{}\ =\ 1

for all e1,,e5Ee_{1},\dotsc,e_{5}\in E and JEJ\subset E such that (Je5;e1,e2,e3,e4)(Je_{5};e_{1},e_{2},e_{3},e_{4}), (Je3;e1,e2,e4,e5)(Je_{3};e_{1},e_{2},e_{4},e_{5}) and (Je4;e1,e2,e5,e3)(Je_{4};e_{1},e_{2},e_{5},e_{3}) are in ΩM\Omega_{M};

[e1e2e3e4]Je5=[e1e2e3e4]Je6\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{Je_{5}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{Je_{5}}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{Je_{6}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{Je_{6}}}{}{}

for all e1,,e6Ee_{1},\dotsc,e_{6}\in E and JEJ\subset E such that Je5=Je6\langle Je_{5}\rangle=\langle Je_{6}\rangle and such that (Je5;e1,e2,e3,e4)(Je_{5};e_{1},e_{2},e_{3},e_{4}) and (Je6;e1,e2,e3,e4)(Je_{6};e_{1},e_{2},e_{3},e_{4}) are in ΩM\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}};
as well as the additive Plücker relations

[e1e2e3e4]J+[e1e3e2e4]J= 1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}+\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{3}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{2}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{3}}\\ {e_{2}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}\ =\ 1

for all (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}}.

Proof.

By Proposition 3.6, we have [H1H2H3H4]J=[H1H2H3H4]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}_{J}}{}{}=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{} for every (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M} and Hi=JeiH_{i}=\langle Je_{i}\rangle for i=1,,4i=1,\dotsc,4. Thus (4.20)–(4.20) follow from (4.18)–(4.18). To see that (4.20) implies (4.18), define for given j1,,jr3,e1,,e6Ej_{1},\dotsc,j_{r-3},e_{1},\dotsc,e_{6}\in E and J={j1,,jr3}J=\{j_{1},\dotsc,j_{r-3}\} as in (4.20) the corank 22-flats Fi=JeiF_{i}=\langle Je_{i}\rangle for i=1,,5i=1,\dotsc,5, which are pairwise distinct and contain the common flat J\langle J\rangle of corank 33, as required. For iji\neq j, we define hyperplanes Hij=FiFj=JeiejH_{ij}=\langle F_{i}\cup F_{j}\rangle=\langle Je_{i}e_{j}\rangle. Then we have for all identifications {i,j,k}={3,4,5}\{i,j,k\}=\{3,4,5\} that

[e1e2eiej]Jek=[H1kH2kHikHjk],\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{i}$}}&{{\scalebox{0.9}{$e_{j}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{Je_{k}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{i}}&{e_{j}}\end{smallmatrix}\big{]}_{Je_{k}}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1k}$}}&{\scalebox{0.9}{$H_{2k}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{ik}$}}&{{\scalebox{0.9}{$H_{jk}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1k}}&{H_{2k}}\\ {H_{ik}}&{H_{jk}}\end{smallmatrix}\big{]}}{}{},

which shows that (4.18) implies (4.20). The relation (4.20) follows from

[e1e2e3e4]Je5=[H1H2H3H4]=[e1e2e3e4]Je6,\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{Je_{5}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{Je_{5}}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$H_{1}$}}&{\scalebox{0.9}{$H_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$H_{3}$}}&{{\scalebox{0.9}{$H_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{H_{1}}&{H_{2}}\\ {H_{3}}&{H_{4}}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{Je_{6}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{Je_{6}}}{}{},

where Hi=Je5ei=Je6eiH_{i}=\langle Je_{5}e_{i}\rangle=\langle Je_{6}e_{i}\rangle is ii-th coefficient of the common image (H1,,H4)(H_{1},\dots,H_{4}) of (Je5;e1,,e4)(Je_{5};e_{1},\dotsc,e_{4}) and (Je6;e1,,e4)(Je_{6};e_{1},\dotsc,e_{4}) under Ψ:ΩMΘM\Psi:\Omega_{M}\to\Theta_{M}.

We are left to show that [e1e2e3e4]J=[e1e2e3e4]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}^{\prime}$}}&{\scalebox{0.9}{$e_{2}^{\prime}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}^{\prime}$}}&{{\scalebox{0.9}{$e_{4}^{\prime}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{e_{1}^{\prime}}&{e_{2}^{\prime}}\\ {e_{3}^{\prime}}&{e_{4}^{\prime}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{} if Ψ(J;e1,,e4)=Ψ(J;e1,,e4)\Psi(J;e_{1},\dotsc,e_{4})=\Psi(J^{\prime};e^{\prime}_{1},\dotsc,e^{\prime}_{4}), i.e. if Jei=Jei\langle Je_{i}\rangle=\langle J^{\prime}e_{i}^{\prime}\rangle for i=1,,4i=1,\dotsc,4. We will prove this by replacing one element of Je1e4Je_{1}\dotsc e_{4} by an element of Je1e4J^{\prime}e^{\prime}_{1}\dotsc e^{\prime}_{4} at a time. Note that both JJ and JJ^{\prime} are bases of the restriction M|F=M\(EF)M|_{F}=M\backslash(E-F), where F=J=JF=\langle J\rangle=\langle J^{\prime}\rangle is the flat of rank r2r-2 generated by JJ and JJ^{\prime}. Since the basis exchange graph of M|FM|_{F} is connected, we find a sequence J=J0,J1,,Js1,Js=JJ=J_{0},J_{1},\dotsc,J_{s-1},J_{s}=J^{\prime} of bases for M|FM|_{F} such that Jk=IkjkJ_{k}=I_{k}j_{k} and Jk+1=IkjkJ_{k+1}=I_{k}j^{\prime}_{k} for Ik=JkJk+1I_{k}=J_{k}\cap J_{k+1} and some jkJkj_{k}\in J_{k} and jkJk+1j_{k}^{\prime}\in J_{k+1}. Considered as subsets of MM, we have Jk=F\langle J_{k}\rangle=F and thus (Jk;e1,,e4)ΩM(J_{k};e_{1}^{\prime},\dotsc,e_{4}^{\prime})\in\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}} for all k=0,,sk=0,\dotsc,s. Thus we can apply (4.20), which yields

[e1e2e3e4]Jk=[e1e2e3e4]Ikjk=[e1e2e3e4]Ikjk=[e1e2e3e4]Jk+1.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J_{k}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J_{k}}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{I_{k}j_{k}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{I_{k}j_{k}}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{I_{k}j^{\prime}_{k}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{I_{k}j^{\prime}_{k}}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J_{k+1}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J_{k+1}}}{}{}.

We conclude that [e1e2e3e4]J=[e1e2e3e4]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{}.

Next we replace the eie_{i} by the eie_{i}^{\prime}, one at a time. After permuting rows and columns appropriately, which does not change the value of the cross ratio by (4.20), we are reduced to studying cross ratios of the forms [f1f2f3f4]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$f_{1}$}}&{\scalebox{0.9}{$f_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$f_{3}$}}&{{\scalebox{0.9}{$f_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{f_{1}}&{f_{2}}\\ {f_{3}}&{f_{4}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{} and [f1f2f3f4]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$f_{1}$}}&{\scalebox{0.9}{$f_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$f_{3}$}}&{{\scalebox{0.9}{$f^{\prime}_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{f_{1}}&{f_{2}}\\ {f_{3}}&{f^{\prime}_{4}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{} such that Jf4=Jf4\langle J^{\prime}f_{4}\rangle=\langle J^{\prime}f^{\prime}_{4}\rangle is a hyperplane. By (4.20), we have

[f1f2f3f4]J[f1f2f4f4]J[f1f2f4f3]J= 1.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$f_{1}$}}&{\scalebox{0.9}{$f_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$f_{3}$}}&{{\scalebox{0.9}{$f_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{f_{1}}&{f_{2}}\\ {f_{3}}&{f_{4}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$f_{1}$}}&{\scalebox{0.9}{$f_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$f_{4}$}}&{{\scalebox{0.9}{$f_{4}^{\prime}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{f_{1}}&{f_{2}}\\ {f_{4}}&{f_{4}^{\prime}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{}\cdot\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$f_{1}$}}&{\scalebox{0.9}{$f_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$f_{4}^{\prime}$}}&{{\scalebox{0.9}{$f_{3}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{f_{1}}&{f_{2}}\\ {f_{4}^{\prime}}&{f_{3}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{}\ =\ 1.

Since Jf4=Jf4\langle J^{\prime}f_{4}\rangle=\langle J^{\prime}f^{\prime}_{4}\rangle is a hyperplane, the subset Jf4f4J^{\prime}f_{4}f_{4}^{\prime} of MM has rank r1r-1 and is not a basis of MM. Thus [f1f2f4f4]J=1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$f_{1}$}}&{\scalebox{0.9}{$f_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$f_{4}$}}&{{\scalebox{0.9}{$f_{4}^{\prime}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{f_{1}}&{f_{2}}\\ {f_{4}}&{f_{4}^{\prime}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{}=1 by (4.20), which shows that

[f1f2f3f4]J=[f1f2f4f3]J1=[f1f2f3f4]J,\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$f_{1}$}}&{\scalebox{0.9}{$f_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$f_{3}$}}&{{\scalebox{0.9}{$f_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{f_{1}}&{f_{2}}\\ {f_{3}}&{f_{4}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$f_{1}$}}&{\scalebox{0.9}{$f_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$f_{4}^{\prime}$}}&{{\scalebox{0.9}{$f_{3}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}^{-1}}{\big{[}\begin{smallmatrix}{f_{1}}&{f_{2}}\\ {f_{4}^{\prime}}&{f_{3}}\end{smallmatrix}\big{]}_{J^{\prime}}^{-1}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$f_{1}$}}&{\scalebox{0.9}{$f_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$f_{3}$}}&{{\scalebox{0.9}{$f_{4}^{\prime}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{f_{1}}&{f_{2}}\\ {f_{3}}&{f_{4}^{\prime}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{},

where we use (4.20) for the last equality. We conclude that

[e1e2e3e4]J=[e1e2e3e4]J=[e1e2e3e4]J,\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e^{\prime}_{1}$}}&{\scalebox{0.9}{$e^{\prime}_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e^{\prime}_{3}$}}&{{\scalebox{0.9}{$e^{\prime}_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J^{\prime}}}{\big{[}\begin{smallmatrix}{e^{\prime}_{1}}&{e^{\prime}_{2}}\\ {e^{\prime}_{3}}&{e^{\prime}_{4}}\end{smallmatrix}\big{]}_{J^{\prime}}}{}{},

as desired. This completes the proof of the theorem. ∎

Corollary 4.21.

The foundation FMF_{M} of a matroid MM is naturally isomorphic to a quotient

FM(NM of type U24FN)SF_{M}\ \simeq\ \bigg{(}\bigotimes_{\begin{subarray}{c}N\to M\\ \text{ of type $U^{2}_{4}$}\end{subarray}}F_{N}\ \bigg{)}\!\sslash\!S

of a tensor product of foundations FN𝕌F_{N}\simeq{\mathbb{U}}, where the set SS is generated by the relations of type (4.20) in the presence of an F7F_{7} or F7F_{7}^{\ast}-minor and of types (4.20)–(4.20) that are induced by embedded minors M\I/JMM\backslash I/J\to M on at most 66 elements {e1,,e6}=E(IJ)\{e_{1},\dotsc,e_{6}\}=E-(I\cup J).

Proof.

By Theorem 4.20, the foundation is generated by the universal cross ratios [e1e2e3e4]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{} of MM, which are the images [e1e2e3e4]J=φM\I/J([e1e2e3e4])\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}=\myvarphi_{M\backslash I/J}\Big{(}\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}}{}{}\Big{)} of the universal cross ratios [e1e2e3e4]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}}{}{} of minors N=M\I/JN=M\backslash I/J on 44 elements e1,,e4e_{1},\dotsc,e_{4}; cf. Proposition 4.8. The morphisms φM\I/J:FNFM\myvarphi_{M\backslash I/J}:F_{N}\to F_{M} testify that all relations of FNF_{N} also hold in FMF_{M}, and therefore we conclude that FMF_{M} is of the form

FM(NMwith #EN=4FN)SF_{M}\ \simeq\ \bigg{(}\bigotimes_{\begin{subarray}{c}N\to M\\ \text{with $\#E_{N}=4$}\end{subarray}}F_{N}\ \bigg{)}\!\sslash\!S

for some set of 33-term relations SS, where ENE_{N} denotes the ground set of NN. A priori, this holds if we include all relations (4.20)–(4.20) of Theorem 4.20 in SS. To reduce this to the assertion of the corollary, we observe the following.

If N=M\I/JN=M\backslash I/J is a minor on 44 elements that is not of type U24U^{2}_{4}, then NN is regular and FN=𝔽1±F_{N}={{\mathbb{F}}_{1}^{\pm}}. Thus we can omit these factors from the tensor product. Note that (4.20) assures that the cross ratios coming from such a minor are trivial in FMF_{M}. Therefore we can omit (4.20) from SS.

Each of (4.20), (4.20), (4.20) and (4.20) involve only cross ratios that come from the same U24U^{2}_{4}-minor N=M\I/JN=M\backslash I/J. Therefore the analogous relations hold already in FNF_{N}, and we can omit them from the set SS.

By Theorem 4.20, the relation (4.20) holds if MM has a minor of type F7F_{7} or F7F_{7}^{\ast}. Each of the relations (4.20)–(4.20) involve cross ratios that come from the same minor on 55 or 66 elements. This shows all assertions of the corollary. ∎

4.11. A presentation of the foundation by embedded minors

Let N=M\I/JN=M\backslash I/J and N=M\I/JN^{\prime}=M\backslash I^{\prime}/J^{\prime} be two embedded minors of a matroid MM. If III^{\prime}\subset I and JJJ^{\prime}\subset J, then N=N\(II)/(JJ)N=N^{\prime}\backslash(I-I^{\prime})/(J-J^{\prime}) is an embedded minor of NN^{\prime}. We write ι:NN\myiota:N\to N^{\prime} for the inclusion as embedded minors and ι:FNFN\myiota_{\ast}:F_{N}\to F_{N^{\prime}} for the induced morphism between the respective foundations.

Theorem 4.22.

Let MM be a matroid with foundation FMF_{M}. Let {\mathcal{E}} be the collection of all embedded minors N=M\I/JN=M\backslash I/J of MM on at most 77 elements with the following properties:

  • if NN has at most 66 elements, then it contains a minor of type U24U^{2}_{4};

  • if NN has exactly 66 elements, then it contains two parallel elements;

  • if NN has 77 elements, then it is isomorphic to F7F_{7} or F7F_{7}^{\ast}.

Then

FM(NFN)S,F_{M}\ \simeq\ \bigg{(}\bigotimes_{N\in{\mathcal{E}}}F_{N}\bigg{)}\!\sslash\!S,

where the set SS is generated by the relations a=ι(a)a=\myiota_{\ast}(a) for every inclusion ι:NN\myiota:N\to N^{\prime} of embedded minors NN and NN^{\prime} in {\mathcal{E}}.

Proof.

It is clear that the morphisms φM\I/J:FM\I/JFM\myvarphi_{M\backslash I/J}:F_{M\backslash I/J}\to F_{M} from Proposition 4.8 induce a canonical morphism (NFN)SFM\Big{(}\bigotimes_{N\in{\mathcal{E}}}F_{N}\Big{)}\!\sslash\!S\to F_{M}, and since {\mathcal{E}} contains all embedded U24U^{2}_{4}-minors of MM, this morphism is surjective. Thus we are left with showing that SS contains all defining relations of MM.

Let us define i={N#EN=i}{\mathcal{E}}_{i}=\{N\in{\mathcal{E}}\mid\#E_{N}=i\} for i=4,,7i=4,\dotsc,7 where ENE_{N} denotes the ground set of the embedded minor NN. Then =47{\mathcal{E}}={\mathcal{E}}_{4}\cup\dotsc\cup{\mathcal{E}}_{7}. The set 4{\mathcal{E}}_{4} consists of the embedded U24U^{2}_{4}-minors of MM, and thus

FM(N4FN)SF_{M}\simeq\bigg{(}\bigotimes_{N\in{\mathcal{E}}_{4}}F_{N}\ \bigg{)}\!\sslash\!S^{\prime}

by Corollary 4.21, where SS^{\prime} contains all relations of types (4.20) (in the presence of an F7F_{7} or F7F_{7}^{\ast}-minor) and (4.20)–(4.20).

The relations (4.20) and (4.20) stem from embedded minors N=M\I/JN=M\backslash I/J on 55 elements, and these relations involve a nondegenerate cross ratio only if NN contains a U24U^{2}_{4}-minor, i.e. N5N\in{\mathcal{E}}_{5}. Thus (4.20) and (4.20) can be replaced by tensoring with FNF_{N} and including the relations a=ι(a)a=\myiota_{\ast}(a) for every minor embedding ι:N=N\I/JN\myiota:N^{\prime}=N\backslash I^{\prime}/J^{\prime}\to N with N4N^{\prime}\in{\mathcal{E}}_{4}.

Similarly, (4.20) stems from embedded minors N=M\I/JN=M\backslash I/J on 66 elements with two parallel elements, and involves a nondegenerate cross ratio only if NN contains a U24U^{2}_{4}-minor, i.e. N6N\in{\mathcal{E}}_{6}. Thus (4.20) can be replaced by tensoring with FNF_{N} and including the relations a=ι(a)a=\myiota_{\ast}(a) for every minor embedding ι:N=N\I/JN\myiota:N^{\prime}=N\backslash I^{\prime}/J^{\prime}\to N with N4N^{\prime}\in{\mathcal{E}}_{4}.

The set 7{\mathcal{E}}_{7} consists of all embedded minors of types F7F_{7} and F7F_{7}^{\ast}. Since FF7=FF7=𝔽2F_{F_{7}}=F_{F_{7}^{\ast}}={\mathbb{F}}_{2} and P1=1P𝔽2P\!\sslash\!\langle 1=-1\rangle\simeq P\otimes{\mathbb{F}}_{2} for every pasture PP, we can replace the relation (4.20) by FN-\otimes F_{N} if N7N\in{\mathcal{E}}_{7}. This recovers all relations in SS^{\prime} and completes the proof. ∎

5. The structure theorem

In this section, we prove the central result of this paper, Theorem 5.9, which asserts that the foundation of a matroid MM without large uniform minors is isomorphic to a tensor product of finitely many copies of the pastures 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}}, 𝔽3{\mathbb{F}}_{3} and 𝔽2{\mathbb{F}}_{2}.

This is done by first showing that in the absence of large uniform minors, the tip and cotip relations are of a particularly simple form, which eventually leads to the conclusion that the foundation of MM is the tensor product of quotients of 𝕌{\mathbb{U}} by automorphism groups, and possibly 𝔽2{\mathbb{F}}_{2}. The quotients of 𝕌{\mathbb{U}} by automorphisms are precisely 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}} and 𝔽3{\mathbb{F}}_{3}.

5.1. Foundations of matroids on 55 elements

By Theorem 4.22, the foundation of a matroid is determined completely by its minors on at most 55 elements and the embedded minors on 66 with two parallel elements.

In this section, we will determine the foundations of all matroids on at most 55 elements. Most of these matroids are regular and have foundation 𝔽1±{{\mathbb{F}}_{1}^{\pm}} by [5, Thm. 7.33]. There is only a small number of non-regular matroids on at most 55 elements, which we will inspect in detail.

Let 0rn50\leqslant r\leqslant n\leqslant 5 and MM be a matroid of rank rr on E={1,,n}E=\{1,\dotsc,n\}.

5.1.1. Regular matroids

A matroid MM is regular if and only if there is no nontrivial cross ratio, which is the case if and only if the matroid MM does not contain any minor of type U24U^{2}_{4}.

This is the case in exactly one of the following situations: (a) r{0,1,n1,n}r\in\{0,1,n-1,n\}; (b) n=4n=4, r=2r=2 and MM is not uniform; (c) n=5n=5, r=2r=2 and M\iM\backslash i is not uniform for every iEi\in E; (d) n=5n=5, r=3r=3 and M/iM/i is not uniform for every iEi\in E.

5.1.2. Matroids with exactly one embedded U24U^{2}_{4}-minor

There are several isomorphism classes of matroids with exactly one U24U^{2}_{4}-minor, which we list in the following.

Since the tip and cotip relations involve cross ratios from different embedded U24U^{2}_{4}-minors, they do not appear for matroids with only one embedded U24U^{2}_{4}-minor.

If n=4n=4, then there is exactly one such matroid, namely M=U24M=U^{2}_{4} itself, which has foundation 𝕌{\mathbb{U}} by Proposition 4.10.

Proposition 5.1.

Let MM be a matroid on 55 elements with exactly one embedded U24U^{2}_{4}-minor. Then MM is isomorphic to U24NU^{2}_{4}\oplus N where NN is a matroid on 11 element. The foundation of MM is isomorphic to 𝕌{\mathbb{U}}.

Proof.

In order to have an U24U^{2}_{4}-minor, MM must have rank 22 or 33. Since the embedded minors NMN\to M of MM correspond bijectively to the embedded minors NMN^{\ast}\to M^{\ast} and since U24U^{2}_{4} is self-dual, the matroids MM and MM^{\ast} have the same number of U24U^{2}_{4}-minors. Once we have shown that every rank 22-matroid with exactly one embedded U24U^{2}_{4}-minor is isomorphic to U24NU^{2}_{4}\oplus N for a matroid NN on one element, which has to be of rank 0, then we can conclude that MM^{\ast} is isomorphic to U24NU^{2}_{4}\oplus N^{\ast}. To complete this reduction to the rank 22-case, we note that the foundation of MM^{\ast} is canonically isomorphic to the foundation of MM, cf. Proposition 4.7.

Assume that the rank 22-matroid MM on E={1,,5}E=\{1,\dotsc,5\} has an embedded U24U^{2}_{4}-minor. After a permutation of EE, we can assume that this embedded U24U^{2}_{4}-minor is M\5=M\{5}M\backslash 5=M\backslash\{5\}, i.e. that all of the following 22-subsets

{1,2},{1,3},{1,4},{2,3},{2,4}and{3,4}\{1,2\},\quad\{1,3\},\quad\{1,4\},\quad\{2,3\},\quad\{2,4\}\quad\text{and}\quad\{3,4\}

of EE are bases. If these are all bases of MM, then 55 is a loop and MM is isomorphic to U24NU^{2}_{4}\oplus N, as claimed.

We indicate why MM cannot have more bases of the form {i,5}\{i,5\}. If MM has exactly one additional basis element, say {1,5}\{1,5\}, then the basis exchange property is violated by exchanging 11 by an element of the basis {3,4}\{3,4\}. The same reason excludes the possibility that MM has exactly two additional basis elements, say {1,5}\{1,5\} and {2,5}\{2,5\}. If MM has 99 or more basis elements, say all 22-subsets of EE but possibly {4,5}\{4,5\}, then both minors M\4M\backslash 4 and M\5M\backslash 5 are isomorphic to U24U^{2}_{4}. Thus in this case, MM has at least two embedded U24U^{2}_{4}-minors.

This shows that MM has to be isomorphic to U24NU^{2}_{4}\oplus N. Since 55 is a loop, the conditions for the tip relations are not satisfied, which means that all relations stem from the unique embedded U24U^{2}_{4}-minor M\5M\backslash 5. This shows that the foundation of MM is isomorphic to FM\5𝕌F_{M\backslash 5}\simeq{\mathbb{U}}, as claimed. ∎

5.1.3. Matroids with exactly two embedded U24U^{2}_{4}-minors

If MM has two embedded U24U^{2}_{4}-minors, then the ground set must be E={1,,5}E=\{1,\dotsc,5\}. As explained in Section 5.1.2, MM must have rank 22 or 33 if MM has an U24U^{2}_{4}-minor. We will show that if MM has exactly two embedded U24U^{2}_{4}-minors, then it must be isomorphic to the following matroid, or its dual.

Definition 5.2.

We denote by C5C_{5} the rank 33-matroid on E={1,,5}E=\{1,\dotsc,5\} whose set of bases is (E3){3,4,5}\binom{E}{3}-\{3,4,5\}.

Proposition 5.3.

A matroid MM on 55 elements has exactly two embedded U24U^{2}_{4}-minors if and only if MM is isomorphic to either C5C_{5} or its dual. The cross ratios of C5C_{5} satisfy

[ijk4]5=[ijk5]4,\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$k$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{5}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {k}&{4}\end{smallmatrix}\big{]}_{5}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$k$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{4}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {k}&{5}\end{smallmatrix}\big{]}_{4}}{}{},

and the cross ratios of C5C_{5}^{\ast} satisfy

[ijk4]=[ijk5]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$k$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {k}&{4}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$k$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {k}&{5}\end{smallmatrix}\big{]}}{}{}

for all identifications {i,j,k}={1,2,3}\{i,j,k\}=\{1,2,3\}. The foundations of both C5C_{5} and C5C_{5}^{\ast} are isomorphic to 𝕌{\mathbb{U}}.

We illustrate all non-degenerate cross ratios of C5C_{5}^{\ast} and their relations in Figure 3.

𝟏\mathbf{-1\ }[1234]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{}[1324]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {2}&{4}\end{smallmatrix}\big{]}}{}{}[3124]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$3$}}&{\scalebox{0.9}{$1$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{3}&{1}\\ {2}&{4}\end{smallmatrix}\big{]}}{}{}[3214]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$3$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$1$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{3}&{2}\\ {1}&{4}\end{smallmatrix}\big{]}}{}{}[2314]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$2$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$1$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{2}&{3}\\ {1}&{4}\end{smallmatrix}\big{]}}{}{}[2134]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$2$}}&{\scalebox{0.9}{$1$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{2}&{1}\\ {3}&{4}\end{smallmatrix}\big{]}}{}{}[1235]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{2}\\ {3}&{5}\end{smallmatrix}\big{]}}{}{}[1325]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$1$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{1}&{3}\\ {2}&{5}\end{smallmatrix}\big{]}}{}{}[3125]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$3$}}&{\scalebox{0.9}{$1$}}\\[-2.0pt] {\scalebox{0.9}{$2$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{3}&{1}\\ {2}&{5}\end{smallmatrix}\big{]}}{}{}[3215]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$3$}}&{\scalebox{0.9}{$2$}}\\[-2.0pt] {\scalebox{0.9}{$1$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{3}&{2}\\ {1}&{5}\end{smallmatrix}\big{]}}{}{}[2315]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$2$}}&{\scalebox{0.9}{$3$}}\\[-2.0pt] {\scalebox{0.9}{$1$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{2}&{3}\\ {1}&{5}\end{smallmatrix}\big{]}}{}{}[2135]\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$2$}}&{\scalebox{0.9}{$1$}}\\[-2.0pt] {\scalebox{0.9}{$3$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{2}&{1}\\ {3}&{5}\end{smallmatrix}\big{]}}{}{}++*++*++*++*++*++*============
Figure 3. The cross ratios of C5C_{5}^{\ast} and their relations
Proof.

In the proof of Proposition 5.1, we saw that C5C_{5} has at least two embedded U24U^{2}_{4}-minors, which correspond to the U24U^{2}_{4}-minors C5\4C_{5}\backslash 4 and C5\5C_{5}\backslash 5. All other minors of rank 22 on 44 elements of C5C_{5} are of the form C5\iC_{5}\backslash i for i{1,2,3}i\in\{1,2,3\}. But since {4,5}\{4,5\} is not a basis of C5C_{5}, none of these minors is isomorphic to U24U^{2}_{4}. This shows that C5C_{5} has exactly two embedded U24U^{2}_{4}-minors, as has every matroid MM that is isomorphic to C5C_{5}.

Conversely, assume that MM is a matroid on 55 elements with exactly two embedded U24U^{2}_{4}-minors. Since duality preserves U24U^{2}_{4}-minors, can assume that MM is of rank 22. After a permutation of EE, we can assume that these two embedded U24U^{2}_{4}-minors are M\4M\backslash 4 and M\5M\backslash 5. Thus all of the 22-subsets

{1,2},{1,3},{1,4},{1,5},{2,3},{2,4},{2,5},{3,4}and{3,5}\{1,2\},\ \{1,3\},\ \{1,4\},\ \{1,5\},\ \{2,3\},\ \{2,4\},\ \{2,5\},\ \{3,4\}\ \text{and}\ \{3,5\}

are bases. If {4,5}\{4,5\} was also a basis of MM, then MM would be the uniform matroid U25U^{2}_{5}, which has five U24U^{2}_{4} minors U25\iU^{2}_{5}\backslash i for i=1,,5i=1,\dotsc,5. Thus MM is isomorphic to C5C_{5}. This proves our first claim.

Let us choose an identification {i,j,k}={1,2,3}\{i,j,k\}=\{1,2,3\}. The tip relation (4.20) in Theorem 4.20 with tip {i,j}\{i,j\} and cyclic orientation (k,4,5)(k,4,5) for C5C_{5} is

[ijk4][ij45][ij5k]= 1.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$k$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {k}&{4}\end{smallmatrix}\big{]}}{}{}\ \cdot\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {4}&{5}\end{smallmatrix}\big{]}}{}{}\ \cdot\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$5$}}&{{\scalebox{0.9}{$k$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {5}&{k}\end{smallmatrix}\big{]}}{}{}\ =\ 1.

Since [ij45]=1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {4}&{5}\end{smallmatrix}\big{]}}{}{}=1 is degenerate, we obtain the claimed relation

[ijk4]=[ij5k]1=[ijk5],\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$k$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {k}&{4}\end{smallmatrix}\big{]}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$5$}}&{{\scalebox{0.9}{$k$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}^{-1}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {5}&{k}\end{smallmatrix}\big{]}^{-1}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$k$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {k}&{5}\end{smallmatrix}\big{]}}{}{},

where the second equality is relation (4.20). Similarly, the cotip relation (4.20) with cotip {i,j}\{i,j\} and cyclic orientation (k,4,5)(k,4,5) for C5C_{5}^{\ast} is

[ijk4]5[ij45]k[ij5k]4= 1.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$k$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{5}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {k}&{4}\end{smallmatrix}\big{]}_{5}}{}{}\ \cdot\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{k}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {4}&{5}\end{smallmatrix}\big{]}_{k}}{}{}\ \cdot\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$5$}}&{{\scalebox{0.9}{$k$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{4}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {5}&{k}\end{smallmatrix}\big{]}_{4}}{}{}\ =\ 1.

Since [ij45]k=1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$4$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{k}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {4}&{5}\end{smallmatrix}\big{]}_{k}}{}{}=1 is degenerate, we obtain the claimed relation

[ijk4]5=[ij5k]41=[ijk5]4.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$k$}}&{{\scalebox{0.9}{$4$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{5}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {k}&{4}\end{smallmatrix}\big{]}_{5}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$5$}}&{{\scalebox{0.9}{$k$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{4}^{-1}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {5}&{k}\end{smallmatrix}\big{]}_{4}^{-1}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$i$}}&{\scalebox{0.9}{$j$}}\\[-2.0pt] {\scalebox{0.9}{$k$}}&{{\scalebox{0.9}{$5$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{4}}{\big{[}\begin{smallmatrix}{i}&{j}\\ {k}&{5}\end{smallmatrix}\big{]}_{4}}{}{}.

Since C5C_{5}^{\ast} is a parallel extension of U24U^{2}_{4}, the foundation of C5C_{5}^{\ast} is 𝕌{\mathbb{U}} by Proposition 4.8, which concludes the proof. ∎

5.1.4. Matroids with five embedded U24U^{2}_{4}-minors

The only matroids on at most five elements that do not appear among the previous cases with at most two embedded U24U^{2}_{4}-minors are the uniform matroids U25U^{2}_{5} and U35U^{3}_{5}, which have five embedded U24U^{2}_{4}-minors.

For completeness, we describe their foundations. However, we postpone the proof to a sequel to this paper where we develop more sophisticated methods to calculate the foundations of matroids. Note that the results of this first part are independent from the following result since we consider matroids without large uniform minors.

Proposition 5.4.

The foundations of U25U^{2}_{5} and U35U^{3}_{5} are isomorphic to

𝔽1±x1,,x5{xi+xi1xi+11|i=1,,5}{{\mathbb{F}}_{1}^{\pm}}\langle x_{1},\dotsc,x_{5}\rangle\!\sslash\!\{x_{i}+x_{i-1}x_{i+1}-1\,|\,i=1,\dotsc,5\}

where x0=x5x_{0}=x_{5} and x6=x1x_{6}=x_{1}.

5.2. Symmetry quotients

The classification of foundations of matroids on up to five elements in section 5.1 shows that in a matroid without large uniform minors, all relations between cross ratios of different embedded U24U^{2}_{4}-minors arise from minors of type C5C_{5} or C5C_{5}^{\ast}. Proposition 5.3 shows that these types of minors identify the two hexagons of cross ratios, which implies an identification of two copies of the near-regular partial field 𝕌{\mathbb{U}}; cf. Figure 3. The same happens for relations of type 4.20: they identify two copies of 𝕌{\mathbb{U}}.

It can, and it will, happen that a matroid contains a chain of such minors, which creates a self-identification of the cross ratios belonging to an embedded U24U^{2}_{4}-minor of MM. By Proposition 5.3, this self-identification must respect the relations between the cross ratios in each hexagon, and induces an automorphism of 𝕌{\mathbb{U}}. Therefore we are led to study the quotients of 𝕌{\mathbb{U}} by such automorphisms.

5.2.1. Automorphisms of the near-regular partial field

In the following, we determine all automorphisms of the near-regular partial field 𝕌=𝔽1±x,y{x+y=1}{\mathbb{U}}={{\mathbb{F}}_{1}^{\pm}}\langle x,y\rangle\!\sslash\!\{x+y=1\}. By Lemma 4.12, it suffices to determine the images of xx and yy to describe an automorphism of 𝕌{\mathbb{U}}. A result equivalent to the following is also proved in [24, Lemma 4.4].

Lemma 5.5.

The elements of the form z+z1z+z^{\prime}-1 in the nullset N𝕌N_{\mathbb{U}} of 𝕌{\mathbb{U}} with z,z𝕌×z,z^{\prime}\in{\mathbb{U}}^{\times} are

x+y1,x1x1y1andy1xy11.x+y-1,\qquad x^{-1}-x^{-1}y-1\qquad\text{and}\qquad y^{-1}-xy^{-1}-1.

Thus the fundamental elements of 𝕌{\mathbb{U}} are x,y,x1,x1y,y1,xy1x,\ y,\ x^{-1},\ -x^{-1}y,\ y^{-1},\ -xy^{-1}.

Proof.

Note that the only element zz with z+11=0z+1-1=0 is z=0z=0. Thus to find all fundamental elements, it suffices to search for relations of the form z+z1N𝕌z+z^{\prime}-1\in N_{\mathbb{U}} with z,z𝕌×z,z^{\prime}\in{\mathbb{U}}^{\times}. Since N𝕌N_{\mathbb{U}} is generated by 11+01-1+0 and x+y1x+y-1, and since all terms have to be nonzero and at least one term has to be equal to 1-1 to find a relation for fundamental elements, we find exactly three relations of the form z+z1=0z+z^{\prime}-1=0, which are

x+y1,x1x1y1andy1xy11.x+y-1,\qquad x^{-1}-x^{-1}y-1\qquad\text{and}\qquad y^{-1}-xy^{-1}-1.

Thus the claim of the lemma. ∎

Proposition 5.6.

The associations

ρ:𝕌𝕌xy1yxy1andσ:𝕌𝕌,xyyx\begin{array}[]{cccc}\myrho:&{\mathbb{U}}&\longrightarrow&{\mathbb{U}}\\ &x&\longmapsto&y^{-1}\\ &y&\longmapsto&-xy^{-1}\end{array}\qquad\begin{array}[]{c}\text{and}\\ \\ {}\hfil\end{array}\qquad\begin{array}[]{cccc}\mysigma:&{\mathbb{U}}&\longrightarrow&{\mathbb{U}},\\ &x&\longmapsto&y\\ &y&\longmapsto&x\end{array}

define automorphisms of 𝕌{\mathbb{U}} that generate the automorphism group of 𝕌{\mathbb{U}} and satisfy the relations ρ3=σ2=(ρσ)2=id\myrho^{3}=\mysigma^{2}=(\myrho\mysigma)^{2}=\textup{id}. In particular, Aut(𝕌)S3\operatorname{Aut}({\mathbb{U}})\simeq S_{3}.

Proof.

By Lemma 5.5, both (y1,xy1)(y^{-1},-xy^{-1}) and (y,x)(y,x) are pairs of fundamental elements in 𝕌{\mathbb{U}}. Thus, by Lemma 4.12, ρ\myrho and σ\mysigma define morphisms from 𝕌{\mathbb{U}} to 𝕌{\mathbb{U}}. Since ρ3(x)=x\myrho^{3}(x)=x and ρ3(y)=y\myrho^{3}(y)=y, we conclude that ρ\myrho defines a group automorphism of 𝕌×{\mathbb{U}}^{\times} of order 33. Similarly, σ\mysigma defines a group automorphism of 𝕌×{\mathbb{U}}^{\times} of order 22. The relation (ρσ)2=id(\myrho\mysigma)^{2}=\textup{id} can be easily verified by evaluation on xx and yy.

We conclude that the automorphism group of 𝕌{\mathbb{U}} contains ρ,σS3\langle\myrho,\mysigma\rangle\simeq S_{3} as a subgroup. By Lemma 5.5, 𝕌{\mathbb{U}} contains precisely 66 fundamental elements, which implies easily that Aut(𝕌)\operatorname{Aut}({\mathbb{U}}) is generated by ρ\myrho and σ\mysigma. ∎

Remark 5.7.

It follows from Lemma 5.5 that the isomorphism FU24𝕌F_{U^{2}_{4}}\to{\mathbb{U}} from Proposition 4.10 maps the cross ratios of U24U^{2}_{4} bijectively to the fundamental elements of 𝕌{\mathbb{U}}. We can arrange these fundamental elements in a hexagon

𝟏\mathbf{-1\ }xxyyy1y^{-1}xy1-xy^{-1}x1y-x^{-1}yx1x^{-1}++*++*++*

in the same way as we arrange the cross ratios in Figure 1. It follows from Proposition 5.6 that the automorphisms of 𝕌{\mathbb{U}} correspond bijectively to the symmetries of this hexagon that preserve the edge labels and the inner triangles.

5.2.2. Classification of the symmetry quotients of 𝕌{\mathbb{U}}

A symmetry quotient of 𝕌{\mathbb{U}} is the quotient of 𝕌{\mathbb{U}} by a group of automorphisms. More precisely, if HH is a subgroup of Aut(𝕌)\operatorname{Aut}({\mathbb{U}}), then the quotient of 𝕌{\mathbb{U}} by HH is

𝕌/H=𝕌{x=τ(x),y=τ(y)|τH}.{\mathbb{U}}/H\ =\ {\mathbb{U}}\!\sslash\!\{\,x=\mytau(x),y=\mytau(y)\,|\,\mytau\in H\,\}.

In fact, we have 𝕌/H=𝕌{x=τ(x),y=τ(y)|τS}{\mathbb{U}}/H={\mathbb{U}}\!\sslash\!\{x=\mytau(x),y=\mytau(y)|\mytau\in S\} if SS is a set of generators of HH.

Recall from section 2.1.2 that 𝔽3=𝔽1±{1+1+1}{\mathbb{F}}_{3}={{\mathbb{F}}_{1}^{\pm}}\!\sslash\!\{1+1+1\},

𝔻=𝔽1±z{z+z1}and=𝔽1±z{z3+1,zz21}.{\mathbb{D}}\ =\ {{\mathbb{F}}_{1}^{\pm}}\langle z\rangle\!\sslash\!\{z+z-1\}\qquad\text{and}\qquad{\mathbb{H}}\ =\ {{\mathbb{F}}_{1}^{\pm}}\langle z\rangle\!\sslash\!\{z^{3}+1,\,z-z^{2}-1\}.

Note that this implies that z3=1z^{3}=-1 and z6=1z^{6}=1 in {\mathbb{H}}.

Proposition 5.8.

The symmetry quotients of 𝕌{\mathbb{U}} are, up to isomorphism,

𝕌/id𝕌,𝕌/σ𝔻,𝕌/ρ,𝕌/ρ,σ𝔽3.{\mathbb{U}}/\langle\textup{id}\rangle\ \simeq\ {\mathbb{U}},\qquad{\mathbb{U}}/\langle\mysigma\rangle\ \simeq\ {\mathbb{D}},\qquad{\mathbb{U}}/\langle\myrho\rangle\ \simeq\ {\mathbb{H}},\qquad{\mathbb{U}}/\langle\myrho,\mysigma\rangle\ \simeq\ {\mathbb{F}}_{3}.
Proof.

In the following, we show that the quotients of 𝕌{\mathbb{U}} by different subgroups HH of Aut(𝕌)S3\operatorname{Aut}({\mathbb{U}})\simeq S_{3} are exactly the pastures 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}} and 𝔽3{\mathbb{F}}_{3}, up to isomorphism. Clearly 𝕌=𝕌/id{\mathbb{U}}={\mathbb{U}}/\langle\textup{id}\rangle is the quotient of 𝕌{\mathbb{U}} by the trivial subgroup.

Note that if HH^{\prime} is a subgroup conjugate to HH, i.e. H=τHτ1H^{\prime}=\mytau H\mytau^{-1} for some τAut(𝕌)\mytau\in\operatorname{Aut}({\mathbb{U}}), then the quotient of 𝕌{\mathbb{U}} by HH^{\prime} equals the quotient of τ(𝕌)=𝕌\mytau({\mathbb{U}})={\mathbb{U}} by HH. This means that it suffices to determine the isomorphism classes of the quotients of 𝕌{\mathbb{U}} by the groups σ\langle\mysigma\rangle, ρ\langle\myrho\rangle and Aut(𝕌)=ρ,σ\operatorname{Aut}({\mathbb{U}})=\langle\myrho,\mysigma\rangle, which represent all conjugacy classes of nontrivial subgroups of Aut(𝕌)\operatorname{Aut}({\mathbb{U}}).

Let H=σH=\langle\mysigma\rangle. We denote the residue classes of xx and yy in 𝕌/σ{\mathbb{U}}/\langle\mysigma\rangle by x¯\bar{x} and y¯\bar{y}, respectively. We claim that the association

f:𝕌/σ𝔻x¯zy¯z\begin{array}[]{cccc}f:&{\mathbb{U}}/\langle\mysigma\rangle&\longrightarrow&{\mathbb{D}}\\ &\bar{x}&\longmapsto&z\\ &\bar{y}&\longmapsto&z\end{array}

defines an isomorphism of pastures. We begin with the verification that ff defines a morphism. The map f^:𝕌𝔻\hat{f}:{\mathbb{U}}\to{\mathbb{D}} with f^(x)=f^(y)=z\hat{f}(x)=\hat{f}(y)=z is a morphism, since the generator x+y1x+y-1 of the nullset of 𝕌{\mathbb{U}} is mapped to z+z1z+z-1, which is in the nullset of 𝔻{\mathbb{D}}. Since f^(σ(x))=z=f^(x)\hat{f}\big{(}\mysigma(x)\big{)}=z=\hat{f}(x) and f^(σ(y))=z=f^(y)\hat{f}\big{(}\mysigma(y)\big{)}=z=\hat{f}(y), the morphism f^\hat{f} induces a morphism f:𝕌/σ𝔻f:{\mathbb{U}}/\langle\mysigma\rangle\to{\mathbb{D}} by the universal property of the quotient 𝕌/σ=𝕌{σ(x)=y,σ(y)=x}{\mathbb{U}}/\langle\mysigma\rangle={\mathbb{U}}\!\sslash\!\{\mysigma(x)=y,\mysigma(y)=x\}, cf. Proposition 2.6.

We define the inverse to ff as the association g:zx¯g:z\mapsto\bar{x}. This defines a multiplicative map since 𝔻×{\mathbb{D}}^{\times} is freely generated by zz. Since

g(z)+g(z)1=x¯+x¯1=x¯+y¯1g(z)+g(z)-1\ =\ \bar{x}+\bar{x}-1\ =\ \bar{x}+\bar{y}-1

is null in 𝕌/σ{\mathbb{U}}/\langle\mysigma\rangle, this defines a morphism g:𝔻𝕌/σg:{\mathbb{D}}\to{\mathbb{U}}/\langle\mysigma\rangle. It is obvious that gg is an inverse to ff, which shows that ff is an isomorphism.

We continue with the automorphism group H=ρH=\langle\myrho\rangle. We claim that the association

f:𝕌/ρx¯zy¯z2\begin{array}[]{cccc}f:&{\mathbb{U}}/\langle\myrho\rangle&\longrightarrow&{\mathbb{H}}\\ &\bar{x}&\longmapsto&z\\ &\bar{y}&\longmapsto&-z^{2}\end{array}

defines an isomorphism of pastures. We begin with the verification that ff defines a morphism. The map f^:𝕌\hat{f}:{\mathbb{U}}\to{\mathbb{H}} with f^(x)=z\hat{f}(x)=z and f^(y)=z2\hat{f}(y)=-z^{2} is a morphism, since the generator x+y1x+y-1 of the nullset of 𝕌{\mathbb{U}} is mapped to zz21z-z^{2}-1, which is in the nullset of {\mathbb{H}}. Since f^(ρ(x))=f^(y1)=z=f^(x)\hat{f}\big{(}\myrho(x)\big{)}=\hat{f}(y^{-1})=z=\hat{f}(x) and f^(ρ(y))=f^(xy1)=z2=f^(y)\hat{f}\big{(}\myrho(y)\big{)}=\hat{f}(-xy^{-1})=-z^{2}=\hat{f}(y), the morphism f^\hat{f} induces a morphism f:𝕌/ρ𝔻f:{\mathbb{U}}/\langle\myrho\rangle\to{\mathbb{D}} by the universal property of the quotient 𝕌/ρ=𝕌{ρ(x)=y,ρ(y)=x}{\mathbb{U}}/\langle\myrho\rangle={\mathbb{U}}\!\sslash\!\{\myrho(x)=y,\myrho(y)=x\}.

We define the inverse of ff as follows. Let g^:𝔽1±z𝕌/ρ\hat{g}:{{\mathbb{F}}_{1}^{\pm}}\langle z\rangle\to{\mathbb{U}}/\langle\myrho\rangle be the morphism that maps zz to x¯\bar{x}. The defining relations of 𝕌/ρ{\mathbb{U}}/\langle\myrho\rangle are x¯=y¯1\bar{x}=\bar{y}^{-1} and y¯=x¯y¯1\bar{y}=-\bar{x}\bar{y}^{-1}. Thus

g^(z3)+g^(1)=x¯3+1=y¯2x¯+1=x¯1y¯y¯1x¯+1=1+1,\hat{g}(z^{3})+\hat{g}(1)\ =\ \bar{x}^{3}+1\ =\ \bar{y}^{-2}\bar{x}+1\ =\ -\bar{x}^{-1}\bar{y}\bar{y}^{-1}\bar{x}+1\ =\ -1+1,

which is in the nullset of 𝕌/ρ{\mathbb{U}}/\myrho. Since z3=1z^{3}=-1 in {\mathbb{H}}, we have z2=z1-z^{2}=z^{-1} and thus

g^(z)+g^(z2)1=x¯+x¯11=x¯+y¯1,\hat{g}(z)+\hat{g}(-z^{2})-1\ =\ \bar{x}+\bar{x}^{-1}-1\ =\ \bar{x}+\bar{y}-1,

which is also in the nullset of 𝕌/ρ{\mathbb{U}}/\langle\myrho\rangle. This shows that the morphism g^\hat{g} defines a morphism g:𝕌/ρg:{\mathbb{H}}\to{\mathbb{U}}/\langle\myrho\rangle, which is obviously inverse to ff.

Finally we show that 𝕌/ρ,σ{\mathbb{U}}/\langle\myrho,\mysigma\rangle is isomorphic to 𝔽3{\mathbb{F}}_{3}. Since 𝕌/ρ,σ(𝕌/ρ)/σ{\mathbb{U}}/\langle\myrho,\mysigma\rangle\simeq\big{(}{\mathbb{U}}/\langle\myrho\rangle\big{)}/\langle\mysigma\rangle, it suffices to show that the association

f:/σ𝔽3z¯1\begin{array}[]{cccc}f:&{\mathbb{H}}/\langle\mysigma\rangle&\longrightarrow&{\mathbb{F}}_{3}\\ &\bar{z}&\longmapsto&-1\end{array}

is an isomorphism. Since σ(z)=σ(x¯)=y¯=z1\mysigma(z)=\mysigma(\bar{x})=\bar{y}=z^{-1} and f(z¯)=f(z¯1)f(\bar{z})=f(\bar{z}^{-1}), and since f(z6)=(1)6=1=f(1)f(z^{6})=(-1)^{6}=1=f(1), the assignment f(z¯)=1f(\bar{z})=-1 extends to a multiplicative map. Since f(z3)+f(1)=(1)3+1=1+1f(z^{3})+f(1)=(-1)^{3}+1=-1+1 and f(z)+f(z2)1=111f(z)+f(-z^{2})-1=-1-1-1 are null in 𝔽3{\mathbb{F}}_{3}, the map ff is a morphism. Note that in /σ{\mathbb{H}}/\langle\mysigma\rangle, we have z¯3=1\bar{z}^{3}=-1 and z¯=z¯1\bar{z}=\bar{z}^{-1}, and thus z¯=1\bar{z}=-1. We conclude that the assignment g:11=z¯g:1\mapsto 1=-\bar{z} defines a morphism g:𝔽3/σg:{\mathbb{F}}_{3}\to{\mathbb{H}}/\langle\mysigma\rangle, since

g(1)+g(1)+g(1)= 1+1+1=(z¯z¯21)g(1)+g(1)+g(1)\ =\ 1+1+1\ =\ -\big{(}\bar{z}-\bar{z}^{2}-1\big{)}

is null in /σ{\mathbb{H}}/\langle\mysigma\rangle. It is clear that gg is an inverse of ff, which shows that ff is an isomorphism. This concludes the proof of the proposition. ∎

5.3. The structure theorem for matroids without large uniform minors

We are prepared to prove the central result of this paper. In the following, the empty tensor product stands for the initial object in Pastures\operatorname{Pastures}, which is 𝔽1±{{\mathbb{F}}_{1}^{\pm}}.

Theorem 5.9.

Let MM be a matroid without large uniform minors and FMF_{M} its foundation. Then

FMF1FrF_{M}\ \simeq\ F_{1}\otimes\dotsb\otimes F_{r}

for some r0r\geqslant 0 and pastures F1,,Fr{𝕌,𝔻,,𝔽3,𝔽2}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}},{\mathbb{F}}_{3},{\mathbb{F}}_{2}\}.

Proof.

Let {\mathcal{E}} be the collection of embedded minors NN of MM from Theorem 4.22. Then

FM(NFN)S,F_{M}\ \simeq\ \bigg{(}\bigotimes_{N\in{\mathcal{E}}}F_{N}\bigg{)}\!\sslash\!S,

where the set SS is generated by the relations a=ι(a)a=\myiota_{\ast}(a) for every inclusion ι:NN\myiota:N\to N^{\prime} of embedded minors NN and NN^{\prime} in {\mathcal{E}}.

From the analysis in section 5.1, it follows that the foundation FNF_{N} of every embedded minor NN of MM with at most 55 elements is either 𝔽1±{{\mathbb{F}}_{1}^{\pm}} or 𝕌{\mathbb{U}}, where we use the assumption that MM is without minors of types U25U^{2}_{5} and U35U^{3}_{5}. A matroid with foundation 𝔽1±{{\mathbb{F}}_{1}^{\pm}} is regular and has thus no minor of type U24U^{2}_{4}. We conclude that every embedded minor in {\mathcal{E}} on at most 55 elements has foundation 𝕌{\mathbb{U}}.

If an embedded minor NN in {\mathcal{E}} has 66 elements, and thus two of them are parallel, then deleting one of these parallel elements yields an embedded minor N=N\eN^{\prime}=N\backslash e of NN, and the induced morphism FNFNF_{N^{\prime}}\to F_{N} is an isomorphism. Thus also every embedded minor in {\mathcal{E}} with 66 elements has foundation 𝕌{\mathbb{U}}.

Since neither F7F_{7} nor F7F_{7}^{\ast} contains a minor of type 𝕌{\mathbb{U}}, an embedded minor NN in {\mathcal{E}} with 77 elements cannot contain another embedded minor NN^{\prime} in {\mathcal{E}}. Consequently the isomorphism of Theorem 4.22 implies that

FMN7FN(NFN)S,F_{M}\ \simeq\ \bigotimes_{N\in{\mathcal{E}}_{7}}F_{N}\otimes\bigg{(}\bigotimes_{N\in{\mathcal{E}}^{\prime}}F_{N}\bigg{)}\!\sslash\!S^{\prime},

where 7{\mathcal{E}}_{7} is the subset of {\mathcal{E}} that contains all embedded minors with 77 elements, {\mathcal{E}}^{\prime} is the subset of {\mathcal{E}} that contains all embedded minors with at most 66 elements and SS is the set generated by the relations a=ι(a)a=\myiota_{\ast}(a) for every inclusion ι:NN\myiota:N\to N^{\prime} of embedded minors NN and NN^{\prime} in {\mathcal{E}}^{\prime}.

By what we have seen, an inclusion NNN\to N^{\prime} of embedded minors in {\mathcal{E}}^{\prime} is an isomorphism, and either foundation is isomorphic to 𝕌{\mathbb{U}}. Thus all identifications in SS^{\prime} stem from isomorphisms between some factors FNF_{N} of the tensor product. What can, and does, happen is that a chain of such isomorphisms imposes a self-identification of a factor FN𝕌F_{N}\simeq{\mathbb{U}} with itself by a non-trivial automorphism. This leads to a symmetry quotient of 𝕌{\mathbb{U}}, which is one of 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}} and 𝔽3{\mathbb{F}}_{3}. Thus

(NFN)S\bigg{(}\bigotimes_{N\in{\mathcal{E}}^{\prime}}F_{N}\bigg{)}\!\sslash\!S^{\prime}

is a tensor product of copies of 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}} and 𝔽3{\mathbb{F}}_{3}.

This leaves us with the factors FNF_{N} for N7N\in{\mathcal{E}}_{7}. By Theorem 4.20, we have 1=1-1=1, and all cross ratios are trivial since there are no U24U^{2}_{4}-minors. Thus FN𝔽1±{1=1}=𝔽2F_{N}\simeq{{\mathbb{F}}_{1}^{\pm}}\!\sslash\!\{1=-1\}={\mathbb{F}}_{2}. This concludes the proof of the theorem. ∎

Theorem 5.9 can be reformulated as follows, which expresses the dependencies of the factors FiF_{i} on MM.

Corollary 5.10.

Let MM be a matroid without large uniform minors, FMF_{M} its foundation. Then

FMF0F1FrF_{M}\ \simeq\ F_{0}\otimes F_{1}\otimes\dotsb\otimes F_{r}

for a uniquely determined r0r\geqslant 0 and uniquely determined pastures F0{𝔽1±,𝔽2,𝔽3,𝕂}F_{0}\in\{{{\mathbb{F}}_{1}^{\pm}},{\mathbb{F}}_{2},{\mathbb{F}}_{3},{\mathbb{K}}\} and F1,,Fr{𝕌,𝔻,}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}}\}, up to a permutation of the indices 1,,r1,\dotsc,r. We have F0=𝔽2F_{0}={\mathbb{F}}_{2} or F0=𝕂F_{0}={\mathbb{K}} if and only if MM contains a minor of type F7F_{7} or F7F_{7}^{\ast}.

Proof.

By Theorem 5.9, the foundation FMF_{M} of a matroid MM without large uniform minors is isomorphic to a tensor product of copies of 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}}, 𝔽3{\mathbb{F}}_{3} and 𝔽2{\mathbb{F}}_{2}.

Since morphisms from 𝔽2{\mathbb{F}}_{2} and 𝔽3{\mathbb{F}}_{3} into other pastures are uniquely determined, if they exist, we conclude that 𝔽2𝔽2=𝔽2{\mathbb{F}}_{2}\otimes\dotsb\otimes{\mathbb{F}}_{2}={\mathbb{F}}_{2} and 𝔽3𝔽3=𝔽3{\mathbb{F}}_{3}\otimes\dotsb\otimes{\mathbb{F}}_{3}={\mathbb{F}}_{3}. Thus the pasture

𝔽2𝔽2r times𝔽3𝔽3s times\underbrace{{\mathbb{F}}_{2}\otimes\dotsb\otimes{\mathbb{F}}_{2}}_{r\text{ times}}\otimes\underbrace{{\mathbb{F}}_{3}\otimes\dotsb\otimes{\mathbb{F}}_{3}}_{s\text{ times}}

is isomorphic to

𝔽1± if r=s=0;𝔽2 if r>s=0;𝔽3 if s>r=0;𝔽2𝔽3=𝕂 if r,s>0;{{\mathbb{F}}_{1}^{\pm}}\ \text{ if $r=s=0$;}\quad{\mathbb{F}}_{2}\ \text{ if $r>s=0$;}\quad{\mathbb{F}}_{3}\ \text{ if $s>r=0$;}\quad{\mathbb{F}}_{2}\otimes{\mathbb{F}}_{3}={\mathbb{K}}\ \text{ if $r,s>0$;}

cf. Example 2.8 for the equality 𝔽2𝔽3=𝕂{\mathbb{F}}_{2}\otimes{\mathbb{F}}_{3}={\mathbb{K}}. This explains the list of possible isomorphism types for F0F_{0}. Since 𝔽2{\mathbb{F}}_{2} appears as a factor of FMF_{M} if and only if MM has a minor of type F7F_{7} or F7F_{7}^{\ast}, this verifies the last claim of the corollary.

It follows that FMF_{M} is isomorphic to a tensor product of F0F_{0} with pastures F1,,Fr{𝕌,𝔻,}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}}\}.

We are left with establishing the uniqueness claims. To begin with, F0F_{0} is uniquely determined by the presence or absence of the relations 1+1=01+1=0 and 1+1+1=01+1+1=0, which correspond to the relations r>0r>0 and s>0s>0, respectively, in our previous case consideration. Thus F0F_{0} is uniquely determined.

The factors F1,,FrF_{1},\dotsc,F_{r} are determined by the fundamental elements of FMF_{M}, as we explain in the following. Let ιi:FiFjFM\myiota_{i}:F_{i}\to\bigotimes F_{j}\simeq F_{M} be the canonical inclusion. By the construction of the tensor product, the nullset of FMF_{M} consists of all terms of the form dιi(a)+dιi(b)+dιi(c)d\myiota_{i}(a)+d\myiota_{i}(b)+d\myiota_{i}(c) for some i{0,,r}i\in\{0,\dotsc,r\}, dFjd\in\bigotimes F_{j} and a,b,cFia,b,c\in F_{i} such that a+b+ca+b+c is in the nullset of FiF_{i}. The fundamental elements of FMF_{M} stem from such equations for which dιi(a)d\myiota_{i}(a) and dιi(b)d\myiota_{i}(b) are nonzero and dιi(c)=1d\myiota_{i}(c)=-1. Thus d=ιi(c)1=ιi(c1)d=-\myiota_{i}(c)^{-1}=\myiota_{i}(-c^{-1}) is in the image of ιi\myiota_{i}, and therefore dιi(a)=ιi(c1a)d\myiota_{i}(a)=\myiota_{i}(-c^{-1}a) and dιi(b)=ιi(c1b)d\myiota_{i}(b)=\myiota_{i}(-c^{-1}b). Since c1ac1b1-c^{-1}a-c^{-1}b-1 is in the nullset of FiF_{i}, we conclude that all fundamental elements in FMF_{M} are of the form ιi(z)\myiota_{i}(z) for some ii and some fundamental element zz of FiF_{i}.

To make a distinction between the different isomorphism types of the factors, we note that every fundamental element xx with relation x+y1=0x+y-1=0 gives rise to a set {x,x1,y,y1,x1y,xy1}\Big{\{}x,\,x^{-1},\,y,\,y^{-1},\,-x^{-1}y,\,-xy^{-1}\Big{\}} of fundamental elements. If these six fundamental elements come from a factor Fi𝕌F_{i}\simeq{\mathbb{U}}, then they are pairwise different. If they come from a factor Fi𝔻F_{i}\simeq{\mathbb{D}}, then

{x,x1,y,y1,x1y,xy1}={x,y1,x1y}\Big{\{}x,\,x^{-1},\,y,\,y^{-1},\,-x^{-1}y,\,-xy^{-1}\Big{\}}\ =\ \Big{\{}x,\,y^{-1},\,-x^{-1}y\Big{\}}

is a set with three distinct elements. If they come from a factor Fi𝔻F_{i}\simeq{\mathbb{D}}, then

{x,x1,y,y1,x1y,xy1}={x,y}\Big{\{}x,\,x^{-1},\,y,\,y^{-1},\,-x^{-1}y,\,-xy^{-1}\Big{\}}\ =\ \Big{\{}x,y\Big{\}}

is a set with two distinct elements. Note that if F0=𝔽3F_{0}={\mathbb{F}}_{3} or F0=𝕂F_{0}={\mathbb{K}}, then x=1x=-1 is also a fundamental element, and in this case x1=y=y1=x1y=xy1=1x^{-1}=y=y^{-1}=-x^{-1}y=-xy^{-1}=-1 are all equal. This shows that the number of factors of types 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}} and {\mathbb{H}} are determined by the fundamental elements of FMF_{M}, which completes the proof of our uniqueness claims. ∎

Remark 5.11.

In a sequel to this paper, we will show that for all r0r\geqslant 0 and F1,,Fr{𝕌,𝔻,,𝔽3,𝔽2}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}},{\mathbb{F}}_{3},{\mathbb{F}}_{2}\}, there is a matroid MM without large uniform minors whose foundation is isomorphic to the tensor product F1FrF_{1}\otimes\dotsb\otimes F_{r}.

6. Applications

In this concluding part of the paper, we explain various applications of our central result Theorem 5.9. Along with some new results and strengthenings of known facts, we also present short conceptual proofs for a number of established theorems which illustrate the versatility of our structure theory for foundations.

The main technique in most of the upcoming proofs is the following. A matroid MM is representable over a pasture PP if and only there is a morphism from the foundation FMF_{M} of MM to PP. If MM is without large uniform minors, then we know by Theorem 5.9 that FMF_{M} is isomorphic to the tensor product of copies FiF_{i} of 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}}, 𝔽3{\mathbb{F}}_{3} and 𝔽2{\mathbb{F}}_{2}. Thus a morphism from FMF_{M} to PP exists if and only there is a morphism from each FiF_{i} to PP, which in practice is quite easy to determine.

For reference in the later sections, we will provide some general criteria for such morphisms in the following result, and list the outcome for a series of prominent pastures in Table 2.

Lemma 6.1.

Let PP be a pasture.

  1. (1)

    There is a morphism 𝕌P{\mathbb{U}}\to P if and only if PP contains a fundamental element. For a field kk, this is the case if and only if #k3\#k\geqslant 3.

  2. (2)

    There is a morphism 𝔻P{\mathbb{D}}\to P if and only if there is an element uP×u\in P^{\times} such that u+u=1u+u=1. For a field kk, this is the case if and only if chark2\textup{char}\;k\neq 2.

  3. (3)

    There is a morphism P{\mathbb{H}}\to P if and only if there is an element uP×u\in P^{\times} such that u3=1u^{3}=-1 and uu2=1u-u^{2}=1. For a field kk, this is the case if and only if chark=3\textup{char}\;k=3 or if kk contains a primitive third root of unity.

  4. (4)

    There is a morphism 𝔽3P{\mathbb{F}}_{3}\to P if and only if 1+1+1=01+1+1=0 in PP. For a field kk, this is the case if and only if chark=3\textup{char}\;k=3.

  5. (5)

    There is a morphism 𝔽2P{\mathbb{F}}_{2}\to P if and only if 1=1-1=1 in PP. For a field kk, this is the case if and only if chark=2\textup{char}\;k=2.

There exist morphisms from 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}}, 𝔽3{\mathbb{F}}_{3} and 𝔽2{\mathbb{F}}_{2} into the pastures 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}}, 𝔽q{\mathbb{F}}_{q} for q=2,,8q=2,\dotsc,8, {\mathbb{Q}}, {\mathbb{C}}, 𝕊{\mathbb{S}}, {\mathbb{P}} and 𝕎{\mathbb{W}} where Table 2 contains a check mark—a dash indicates that there is no morphism.

Table 2. Existence of morphisms from 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}}, 𝔽3{\mathbb{F}}_{3} and 𝔽2{\mathbb{F}}_{2} into other pastures
𝕌{\mathbb{U}} 𝔻{\mathbb{D}} {\mathbb{H}} 𝔽2{\mathbb{F}}_{2} 𝔽3{\mathbb{F}}_{3} 𝔽4{\mathbb{F}}_{4} 𝔽5{\mathbb{F}}_{5} 𝔽7{\mathbb{F}}_{7} 𝔽8{\mathbb{F}}_{8} {\mathbb{Q}} {\mathbb{C}} 𝕊{\mathbb{S}} {\mathbb{P}} 𝕎{\mathbb{W}}
𝕌{\mathbb{U}} \checkmark \checkmark \checkmark - \checkmark \checkmark \checkmark \checkmark \checkmark \checkmark \checkmark \checkmark \checkmark \checkmark
𝔻{\mathbb{D}} - \checkmark - - \checkmark - \checkmark \checkmark - \checkmark \checkmark \checkmark \checkmark \checkmark
{\mathbb{H}} - - \checkmark - \checkmark \checkmark - \checkmark - - \checkmark - \checkmark \checkmark
𝔽3{\mathbb{F}}_{3} - - - - \checkmark - - - - - - - - \checkmark
𝔽2{\mathbb{F}}_{2} - - - \checkmark - \checkmark - - \checkmark - - - - -
Proof.

We briefly indicate the reasons for claims (1)–(5). We begin with claim (1). The universal property from Proposition 2.6 implies that there is a morphism from 𝕌=𝔽1±x,y{x+y1}{\mathbb{U}}={{\mathbb{F}}_{1}^{\pm}}\langle x,y\rangle\!\sslash\!\{x+y-1\} to PP if and only if there are u,vPu,v\in P such that u+v=1u+v=1. By definition, such elements are fundamental elements of PP. If P=kP=k is a field, then a pair (u,v)(u,v) of fundamental elements is a point of the line L={(w,1w))|wk}L=\{(w,1-w))|w\in k\} in k2k^{2}. Since LL contains precisely two points (0,1)(0,1) and (0,1)(0,1) with vanishing coordinates, the elements of L(k×)2L\cap(k^{\times})^{2} are in bijection with k{0,1}k-\{0,1\}. Thus kk has a fundamental element if and only if #k3\#k\geqslant 3.

We continue with claim (2). The first assertion follows at once from the universal property for 𝔻=𝔽1±z{z+z1}{\mathbb{D}}={{\mathbb{F}}_{1}^{\pm}}\langle z\rangle\!\sslash\!\{z+z-1\}. A field P=kP=k contains an element uu with u+u=1u+u=1 if and only if 1+11+1 is invertible in kk, which is the case if and only if kk is of characteristic different from 22.

We continue with claim (3). The first assertion follows at once from the universal property for =𝔽1±z{z31,zz21}{\mathbb{H}}={{\mathbb{F}}_{1}^{\pm}}\langle z\rangle\!\sslash\!\{z^{3}-1,z-z^{2}-1\}. In a field P=kP=k of characteristic 33, the element u=1u=-1 satisfies u3=1u^{3}=-1 and uu2=1u-u^{2}=1. If kk has characteristic different from 33, then v=uv=-u satisfies the equation v2+v+1=0v^{2}+v+1=0, which characterizes a primitive third root of unity. Note that we have automatically u3=v3=1u^{3}=-v^{3}=-1 in a field if vv is a third root of unity.

Claims (4) and (5) are obvious. The existence or non-existence of morphisms as displayed in Table 2 can be easily verified using (1)–(5). ∎

6.1. Forbidden minors for regular, binary and ternary matroids

The techniques of this paper allow for short arguments to re-establish the known characterizations of regular, binary and ternary matroids in terms of forbidden minors, as they have been proven by Tutte in [31] and [32] for regular and binary matroids, and independently by Bixby in [6] and by Seymour in [29] for ternary matroids.

We spell out the following basic fact for its importance for many of the upcoming theorems.

Lemma 6.2.

Binary matroids and ternary matroids are without large uniform minors.

Proof.

All minors of a binary or ternary matroid are binary or ternary, respectively. Since U25U^{2}_{5} and U35U^{3}_{5} are neither binary nor ternary, the result follows. ∎

Next we turn to the proofs of the excluded minor characterizations of regular, binary and ternary matroids.

Theorem 6.3 (Tutte ’58).

A matroid is regular if and only if it contains no minor of types U24U^{2}_{4}, F7F_{7} or F7F_{7}^{\ast}. A matroid is binary if and only if it contains no minor of type U24U^{2}_{4}.

Proof.

By Corollary 4.13, U24U^{2}_{4} is not binary and therefore also not regular. It follows from Theorem 4.20 that the foundations of F7F_{7} and F7F_{7}^{\ast} contain the relation 1=1-1=1, which means that they do not admit a morphism to 𝔽1±{{\mathbb{F}}_{1}^{\pm}}. Thus F7F_{7} and F7F_{7}^{\ast} are not regular.

We are left with showing that the respective lists of forbidden minors are complete. If a matroid MM does not contain a minor of type U24U^{2}_{4}, then Corollary 4.21 implies that the foundation FMF_{M} of MM is equal to 𝔽1±{{\mathbb{F}}_{1}^{\pm}} or 𝔽1±{1=1}=𝔽2{{\mathbb{F}}_{1}^{\pm}}\!\sslash\!\{-1=1\}={\mathbb{F}}_{2}. In either case, there is a morphism from FMF_{M} to 𝔽2{\mathbb{F}}_{2}, which shows that MM is binary if it has no minor of type U24U^{2}_{4}.

If, in addition, MM has no minor of types F7F_{7} or F7F_{7}^{\ast}, then Corollary 4.21 implies that FM=𝔽1±F_{M}={{\mathbb{F}}_{1}^{\pm}}, and thus MM is regular. ∎

Theorem 6.4 (Bixby ’79, Seymour ’79).

A matroid is ternary if and only if it does not contain a minor of type U25U^{2}_{5}, U35U^{3}_{5}, F7F_{7} or F7F_{7}^{\ast}.

Proof.

If MM is ternary, then it does not have a minor of type U25U^{2}_{5} or U35U^{3}_{5} by Lemma 6.2. Thus Theorem 4.20 applies, and since 11-1\neq 1 in 𝔽3{\mathbb{F}}_{3}, MM does not have a minor of type F7F_{7} or F7F_{7}^{\ast}. This establishes all forbidden minors as listed in the theorem.

To show that the list of forbidden minors is complete, we assume that MM contains no minors of these types. Then Corollary 5.10 implies that the foundation of MM is isomorphic to F1FrF_{1}\otimes\dotsb\otimes F_{r} with Fi{𝕌,𝔻,,𝔽3}F_{i}\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}},{\mathbb{F}}_{3}\}. Since each of 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}}, 𝔽3{\mathbb{F}}_{3} admits a morphism to 𝔽3{\mathbb{F}}_{3}, there is a morphism FM𝔽3F_{M}\to{\mathbb{F}}_{3}, which shows that MM is ternary. ∎

6.2. Uniqueness of the rescaling class over 𝔽3{\mathbb{F}}_{3}

Brylawski and Lucas show in [11] that a representation of a matroid over 𝔽3{\mathbb{F}}_{3} is uniquely determined up to rescaling. Our method yields a short proof of the following generalization.

Theorem 6.5.

Let PP be a pasture with at most one fundamental element. Then every matroid has at most one rescaling class over PP.

Proof.

Let MM be a matroid with foundation FMF_{M}. Since the rescaling classes of MM over PP are in bijective correspondence with the morphisms FMPF_{M}\to P, it suffices to show that there is at most one such morphism.

By Proposition 3.11, every cross ratio of FMF_{M} is a fundamental element of FMF_{M}, and thus must be mapped to a fundamental element zz of PP. By the uniqueness of zz (if it exists), the image of every cross ratio is uniquely determined. Since FMF_{M} is generated over 𝔽1±{{\mathbb{F}}_{1}^{\pm}} by cross ratios, the result follows. ∎

Remark 6.6.

Examples of pastures with at most one fundamental element are 𝔽1±{{\mathbb{F}}_{1}^{\pm}}, 𝔽2{\mathbb{F}}_{2}, 𝔽3{\mathbb{F}}_{3} and 𝕂{\mathbb{K}}. In fact it is not hard to prove that every pasture with at most one fundamental element contains one of these pastures as a subpasture, and that the fundamental element is 1-1 (if it exists). Note that Brylawski and Lucas’s theorem concerns the case P=𝔽3P={\mathbb{F}}_{3}.

6.3. Criteria for representability over certain fields

Our theory allows us to deduce at once that matroids without large minors that are representable over certain pastures are automatically representable over certain (partial) fields. For instance, we find such criteria in the cases of the sign hyperfield 𝕊{\mathbb{S}}, the phase hyperfield {\mathbb{P}} and the weak sign hyperfield 𝕎{\mathbb{W}}.

Note that the proof of Criterion (1) in the following theorem strengthens Lee and Scobee’s result that every ternary and orientable matroid is dyadic; see [17, Cor. 1]. In fact, we further improve on this result in Theorem 6.9 where we show that every orientation is uniquely liftable to 𝔻{\mathbb{D}} up to rescaling.

In the statement of the following theorem, recall that a matroid is said to be weakly orientable if it is representable over 𝕎{\mathbb{W}}.

Theorem 6.7.

Let MM be a matroid without large uniform minors.

  1. (1)

    If MM is orientable, then it is representable over every field of characteristic different from 22.

  2. (2)

    If MM is representable over {\mathbb{P}}, then it is representable over fields of every characteristic except possibly 22.

  3. (3)

    If MM is weakly orientable, then it is ternary.

Proof.

Let FMF_{M} be the foundation of MM and FMF1FrF_{M}\simeq F_{1}\otimes\dotsb\otimes F_{r} the decomposition from Theorem 5.9 into factors Fi{𝕌,𝔻,,𝔽3,𝔽2}F_{i}\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}},{\mathbb{F}}_{3},{\mathbb{F}}_{2}\}. If MM is representable over a pasture PP, then there is a morphism FMPF_{M}\to P, and thus there is a morphism FiPF_{i}\to P for every i=1,,ri=1,\dotsc,r. Conversely, if one of the building blocks 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}}, 𝔽3{\mathbb{F}}_{3} and 𝔽2{\mathbb{F}}_{2} does not map to PP, we conclude that this building block does not occur among the FiF_{i}.

Claim (1) follows since there are no morphisms from {\mathbb{H}}, 𝔽3{\mathbb{F}}_{3} or 𝔽2{\mathbb{F}}_{2} to 𝕊{\mathbb{S}}, and both 𝕌{\mathbb{U}} and 𝔻{\mathbb{D}} map to every field of characteristic different from 22. Claim (2) follows since there are no morphisms from 𝔽3{\mathbb{F}}_{3} or 𝔽2{\mathbb{F}}_{2} to {\mathbb{P}}, and since each of 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}} and {\mathbb{H}} maps to a field kk if its characteristic is 33 or if it is different from 22 and if kk contains a primitive third root of unity. Claim (3) follows since there is no morphism from 𝔽2{\mathbb{F}}_{2} to 𝕎{\mathbb{W}}, and each of 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}} and 𝔽3{\mathbb{F}}_{3} maps to 𝔽3{\mathbb{F}}_{3}. ∎

Remark 6.8.

The proof of Theorem 6.7 shows that similar conclusions can be formulated for other pastures PP that do not receive morphisms from some of the building blocks of the foundation FMF_{M} of a matroid MM without large uniform minors. If MM is representable over PP, then we can conclude the following, for instance:

  • if there is no morphism from 𝔻{\mathbb{D}} to PP, then MM is quaternary;

  • if there is no morphism from either 𝔽2{\mathbb{F}}_{2} or 𝔻{\mathbb{D}} to PP, then MM is hexagonal.

6.4. Oriented matroids without large minors are uniquely dyadic

Our techniques allow us to strengthen the result of Lee and Scobee ([17, Thm. 1]) that an oriented matroid is dyadic if its underlying matroid is ternary. At the end of this section, we deduce Lee and Scobee’s result from ours.

An oriented matroid is an 𝕊{\mathbb{S}}-matroid, i.e. the class M=[Δ]M=[\Delta] of a Grassmann-Plücker function Δ:Er𝕊\Delta:E^{r}\to{\mathbb{S}}, where rr is the rank of MM and EE its ground set. The underlying matroid of MM is the matroid M¯=t𝕊,(M)\underline{M}=t_{{\mathbb{S}},\ast}(M), where t𝕊:𝕊𝕂t_{\mathbb{S}}:{\mathbb{S}}\to{\mathbb{K}} is the terminal morphism, cf. section 2.1.3. Recall that a reorientation class is a rescaling class over 𝕊{\mathbb{S}}.

Let sign:𝔻𝕊\operatorname{sign}:{\mathbb{D}}\to{\mathbb{S}} be the morphism from the dyadic partial field 𝔻=𝔽1±z{z+z1}{\mathbb{D}}={{\mathbb{F}}_{1}^{\pm}}\langle z\rangle\!\sslash\!\{z+z-1\} to 𝕊{\mathbb{S}} that maps zz to 11. An oriented matroid M=[Δ]M=[\Delta] is dyadic if there is a 𝔻{\mathbb{D}}-matroid M^\widehat{M} such that M=sign(M^)M=\operatorname{sign}_{\ast}(\widehat{M}). We call M^\widehat{M} a lift of MM along sign:𝔻𝕊\operatorname{sign}:{\mathbb{D}}\to{\mathbb{S}}.

Theorem 6.9.

Let MM be an oriented matroid whose underlying matroid M¯\underline{M} is without large uniform minors. Then there is a unique rescaling class [M^][\widehat{M}] of dyadic matroids such that sign(M^)=M\operatorname{sign}_{\ast}(\widehat{M})=M.

Proof.

Let FM¯F_{\underline{M}} be the foundation of M¯\underline{M}. The oriented matroid MM determines a reorientation class [M][M] and thus a morphism f:FM¯𝕊f:F_{\underline{M}}\to{\mathbb{S}}. Since rescaling classes of M¯\underline{M} over 𝔻{\mathbb{D}} correspond bijectively to morphisms FM¯𝔻F_{\underline{M}}\to{\mathbb{D}}, we need to show that the morphism f:FM¯𝕊f:F_{\underline{M}}\to{\mathbb{S}} lifts uniquely to 𝔻{\mathbb{D}}, i.e. that there is a unique morphism f^:FM¯𝔻\hat{f}:F_{\underline{M}}\to{\mathbb{D}} such that the diagram

FM¯{F_{\underline{M}}}𝔻{{\mathbb{D}}}𝕊{{\mathbb{S}}}f^\scriptstyle{\hat{f}}f\scriptstyle{f}sign\scriptstyle{\operatorname{sign}}

commutes.

Note that this implies only that there is a unique rescaling class [M^][\widehat{M}] such that the reorientation classes [sign(M^)][\operatorname{sign}_{\ast}(\widehat{M})] and [M][M] are equal. In order to conclude that we can choose M^\widehat{M} such that sign(M^)=M\operatorname{sign}_{\ast}(\widehat{M})=M, we note that the morphism sign:𝔻𝕊\operatorname{sign}:{\mathbb{D}}\to{\mathbb{S}} is surjective, and thus any reorientation M=sign(M^)M^{\prime}=\operatorname{sign}_{\ast}(\widehat{M}) of MM can be inverted by a rescaling of M^\widehat{M} over 𝔻{\mathbb{D}}. This shows that we have proven everything, once we show that ff lifts uniquely to 𝔻{\mathbb{D}}.

Since M¯\underline{M} is without large uniform minors, Theorem 5.9 implies that FM¯F_{\underline{M}} is isomorphic to F1FrF_{1}\otimes\dotsb\otimes F_{r} for some F1,,Fr{𝕌,𝔻,,𝔽3,𝔽2}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}},{\mathbb{F}}_{3},{\mathbb{F}}_{2}\}. Composing f:FM¯𝕊f:F_{\underline{M}}\to{\mathbb{S}} with the canonical inclusions ιi:Fi𝔽M¯\myiota_{i}:F_{i}\to{\mathbb{F}}_{\underline{M}} yields morphisms fi=fιi:Fi𝕊f_{i}=f\circ\myiota_{i}:F_{i}\to{\mathbb{S}} for i=1,,ri=1,\dotsc,r. As visible in Table 2, there are no morphisms from {\mathbb{H}}, 𝔽3{\mathbb{F}}_{3} or 𝔽2{\mathbb{F}}_{2} to 𝕊{\mathbb{S}}. This means that F1,,Fr{𝕌,𝔻}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{D}}\}.

By the universal property of the tensor product, the morphisms 𝔽M¯𝔻{\mathbb{F}}_{\underline{M}}\to{\mathbb{D}} correspond bijectively to the tuples of morphisms fi:Fi𝔻f_{i}:F_{i}\to{\mathbb{D}}. Thus there is a unique lift of ff to 𝔻{\mathbb{D}} if and only if for every ii, there is a unique lift of fif_{i} to 𝔻{\mathbb{D}}. This reduces our task to an inspection of the two cases Fi=𝔻F_{i}={\mathbb{D}} and Fi=𝕌F_{i}={\mathbb{U}}.

Consider the case fi:Fi=𝔻𝕊f_{i}:F_{i}={\mathbb{D}}\to{\mathbb{S}}. Since z+z=1z+z=1 in 𝔻{\mathbb{D}}, we must have f(z)+f(z)=1f(z)+f(z)=1 in 𝕊{\mathbb{S}}, which is only possible if f(z)=1f(z)=1. Thus fi=signf_{i}=\operatorname{sign}, which means that the identity morphism f^i=id:𝔻𝔻\hat{f}_{i}=\textup{id}:{\mathbb{D}}\to{\mathbb{D}} lifts fif_{i}, i.e.

𝔻{{\mathbb{D}}}𝔻{{\mathbb{D}}}𝕊{{\mathbb{S}}}f^i=id\scriptstyle{\hat{f}_{i}=\textup{id}}fi\scriptstyle{f_{i}}sign\scriptstyle{\operatorname{sign}}

commutes. This lift is unique since u+u=1u+u=1 is only satisfied by u=z𝔻u=z\in{\mathbb{D}}, and thus f^i(z)=z\hat{f}_{i}(z)=z is determined.

We are left with the case fi:Fi=𝕌𝕊f_{i}:F_{i}={\mathbb{U}}\to{\mathbb{S}}, for which we inspect the possible images of the fundamental elements xx and yy of 𝕌{\mathbb{U}} in 𝕊{\mathbb{S}} and 𝔻{\mathbb{D}}. The relations of the form u+v1=0u+v-1=0 in 𝕊{\mathbb{S}} are 1+11=01+1-1=0 and 111=01-1-1=0. Thus fif_{i} maps (x,y)(x,y) to one of (1,1)(1,1), (1,1)(1,-1) and (1,1)(-1,1). This means that there are precisely 33 morphisms 𝕌𝕊{\mathbb{U}}\to{\mathbb{S}}, and fif_{i} has to be one of them.

The relations of the form u+v1=0u+v-1=0 in 𝔻{\mathbb{D}} are z+z1=0z+z-1=0 and z111=0z^{-1}-1-1=0. Thus the morphisms 𝕌𝕌{\mathbb{U}}\to{\mathbb{U}} correspond to a choice of mapping (x,y)(x,y) to one of (z,z)(z,z), (z1,1)(z^{-1},-1) and (1,z1)(-1,z^{-1}). Considering the respective images sign(z)=sign(z1)=1\operatorname{sign}(z)=\operatorname{sign}(z^{-1})=1 and sign(1)=1\operatorname{sign}(-1)=-1 in 𝕊{\mathbb{S}}, we conclude that every morphism fi:𝕌𝕊f_{i}:{\mathbb{U}}\to{\mathbb{S}} lifts uniquely to a morphism f^i:𝕌𝔻\hat{f}_{i}:{\mathbb{U}}\to{\mathbb{D}}, i.e.

𝕌{{\mathbb{U}}}𝔻{{\mathbb{D}}}𝕊{{\mathbb{S}}}f^i\scriptstyle{\hat{f}_{i}}fi\scriptstyle{f_{i}}sign\scriptstyle{\operatorname{sign}}

commutes. This completes the proof of the theorem. ∎

As an application, we show how Theorem 6.9 implies the result [17, Thm. 1] of Lee and Scobee.

Theorem 6.10 (Lee–Scobee ’99).

An oriented matroid is dyadic if and only if its underlying matroid is ternary.

Proof.

Let MM be an oriented matroid and let M¯\underline{M} be its underlying matroid. If M¯\underline{M} is ternary, then it is without large uniform minors. Thus MM is dyadic by Theorem 6.9.

Conversely, assume that MM is dyadic, i.e. it has a lift M^\widehat{M} along sign:𝔻𝕊\operatorname{sign}:{\mathbb{D}}\to{\mathbb{S}}. Since there is a morphism f:𝔻𝔽3f:{\mathbb{D}}\to{\mathbb{F}}_{3}, and since t𝔽3f=t𝕊signt_{{\mathbb{F}}_{3}}\circ f=t_{\mathbb{S}}\circ\operatorname{sign}, the 𝔽3{\mathbb{F}}_{3}-matroid f(M^)f_{\ast}(\widehat{M}) is a representation of M¯=t𝕊,(M)\underline{M}=t_{{\mathbb{S}},\ast}(M) over 𝔽3{\mathbb{F}}_{3}. Thus M¯\underline{M} is ternary. ∎

6.5. Positively oriented matroids without large uniform minors are near-regular

In their 2017 paper [2], Ardila, Rincón and Williams prove that every positively oriented matroid can be represented over {\mathbb{R}} (and a posteriori, by a theorem of Postnikov, over {\mathbb{Q}}), which solves a conjecture from da Silva’s thesis [12] from 1987. A second proof has recently been obtained by Speyer and Williams in [30]. Neither of these proofs yields information about the structure of the lifts of positive orientations to {\mathbb{Q}} or {\mathbb{R}}.

With our techniques, we can recover and strengthen the result for positively oriented matroids whose underlying matroid is without large uniform minors. To begin with, let us recall the definition of positively oriented matroids.

Definition 6.11.

Let MM be a matroid of rank rr on the ground set E={1,,n}E=\{1,\dotsc,n\}. A positive orientation of MM (with respect to EE) is a Grassmann-Plücker function Δ:Er𝕊\Delta:E^{r}\to{\mathbb{S}} such that t,𝕊([Δ])=Mt_{\ast,{\mathbb{S}}}([\Delta])=M and such that Δ(j1,,jr){0,1}\Delta(j_{1},\dotsc,j_{r})\in\{0,1\} for every (j1,,jr)Er(j_{1},\dotsc,j_{r})\in E^{r} with j1<<jrj_{1}<\dotsc<j_{r}.

An oriented matroid MM of rank rr on EE is positively oriented if its underlying matroid has a positive orientation Δ:Er𝕊\Delta:E^{r}\to{\mathbb{S}} with respect to some identification E{1,,n}E\simeq\{1,\dotsc,n\} such that M=[Δ]M=[\Delta].

A key tool for proof of Theorem 6.15 is the following notion.

Definition 6.12.

Let MM be a matroid of rank rr on the ground set E={1,,n}E=\{1,\dotsc,n\}. Let VV be the Klein 44-group, considered as a subgroup of S4S_{4}. The Ω\Omega-signature of MM (with respect to EE) is the map

Σ:ΩMS4/V\Sigma:\ \Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}}\ \longrightarrow\ S_{4}/V

that sends (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}} to the class [ϵ]S4/V[\myepsilon]\in S_{4}/V of the uniquely determined permutation ϵS4\myepsilon\in S_{4} that

{e1,,e4}{1,,4}eiϵ(i)\begin{array}[]{ccc}\{e_{1},\dotsc,e_{4}\}&\longrightarrow&\{1,\dotsc,4\}\\ e_{i}&\longmapsto&\myepsilon(i)\end{array}

is an order-preserving bijection.

Example 6.13.

The key example to understand the relevance of the Ω\Omega-signature is the uniform matroid M=U24M=U^{2}_{4}, whose foundation is FM=𝕌F_{M}={\mathbb{U}}. In this case, ΩM\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}} consists of the tuples (;e1,,e4)(\varnothing;e_{1},\dots,e_{4}) for which (e1,,e4)(e_{1},\dotsc,e_{4}) is a permutation of (1,,4)(1,\dotsc,4). Since the cross ratio [e1e2e3e4]FM\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}}{}{}\in F_{M} determines (e1,e2,e3,e4)(e_{1},e_{2},e_{3},e_{4}) up to a permutation in VV, which corresponds to a permutation of the rows and the columns of the cross ratio, the Ω\Omega-signature induces a well-defined bijection

{cross ratios in FM}S4/V[e1e2e3e4]Σ(;e1,,e4).\begin{array}[]{ccc}\Big{\{}\text{cross ratios in }F_{M}\Big{\}}&\longrightarrow&S_{4}/V\\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}}{}{}&\longmapsto&\Sigma(\varnothing;e_{1},\dotsc,e_{4}).\end{array}
Lemma 6.14.

Let MM be a matroid of rank rr on the ground set E={1,,n}E=\{1,\dotsc,n\} and let Δ:Er𝕊\Delta:E^{r}\to{\mathbb{S}} be a positive orientation of MM. Let (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega^{\scalebox{0.7}{$\diamondsuit$}}_{M} and ϵS4\myepsilon\in S_{4} be such that [ϵ]=Σ(J;e1,,e4)[\myepsilon]=\Sigma(J;e_{1},\dotsc,e_{4}). Then

[e1e2e3e4]Δ,J=(1)ϵ(1)+ϵ(2)+1.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\Delta,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\Delta,J}}{}{}\ =\ (-1)^{\myepsilon(1)+\myepsilon(2)+1}.
Proof.

Choose 𝐉=(j1,,jr2)Er2{\mathbf{J}}=(j_{1},\dotsc,j_{r-2})\in E^{r-2} so that |𝐉|=J|{\mathbf{J}}|=J. Since Δ\Delta is a positive orientation, we have for all i{1,2}i\in\{1,2\} and j{3,4}j\in\{3,4\} that Δ(𝐉eiej)=signπi,j\Delta({\mathbf{J}}e_{i}e_{j})=\operatorname{sign}\mypi_{i,j}, where πi,j:JeiejJeiej\mypi_{i,j}:Je_{i}e_{j}\to Je_{i}e_{j} is the unique permutation such that

πi,j(j1)<<πi,j(jr2)<πi,j(ei)<πi,j(ej).\mypi_{i,j}(j_{1})\ <\ \dotsc\ <\ \mypi_{i,j}(j_{r-2})\ <\ \mypi_{i,j}(e_{i})\ <\ \mypi_{i,j}(e_{j}).

Since the cross ratio [e1e2e3e4]Δ,J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\Delta,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\Delta,J}}{}{} is invariant under permutations of JJ, we can assume that j1<<jr2j_{1}<\dotsc<j_{r-2}. Thus we can write πi,j=σi,jϵi,j\mypi_{i,j}=\mysigma_{i,j}\circ\myepsilon_{i,j} as the composition of σi,j=πi,jϵi,j1\mysigma_{i,j}=\mypi_{i,j}\circ\myepsilon_{i,j}^{-1} with the permutation ϵi,j\myepsilon_{i,j} of JeiejJe_{i}e_{j} that fixes j1,,jr2j_{1},\dotsc,j_{r-2} and satisfies ϵi,j(ei)<ϵi,j(ej)\myepsilon_{i,j}(e_{i})<\myepsilon_{i,j}(e_{j}). A minimal decomposition of σi,j\mysigma_{i,j} into transpositions is

σi,j=(jkjej)(jr2ej)(jkiei)(jr2ei),\mysigma_{i,j}\ =\ (j_{k_{j}}\ e_{j})\dotsb(j_{r-2}\ e_{j})\ (j_{k_{i}}\ e_{i})\dotsb(j_{r-2}\ e_{i}),

where kik_{i} is such that jki1<ei<jkij_{k_{i}-1}<e_{i}<j_{k_{i}}. Thus

sign(σi,j)=(1)(r1ki)+(r1kj)=(1)ki+kj,\operatorname{sign}(\mysigma_{i,j})\ =\ (-1)^{\big{(}r-1-k_{i}\big{)}+\big{(}r-1-k_{j}\big{)}}\ =\ (-1)^{k_{i}+k_{j}},

and

[e1e2e3e4]Δ,J\displaystyle\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\Delta,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\Delta,J}}{}{}\ =Δ(𝐉e1e3)Δ(𝐉e2e4)Δ(𝐉e1e4)Δ(𝐉e2e3)\displaystyle=\ \frac{\Delta({\mathbf{J}}e_{1}e_{3})\Delta({\mathbf{J}}e_{2}e_{4})}{\Delta({\mathbf{J}}e_{1}e_{4})\Delta({\mathbf{J}}e_{2}e_{3})}
=sign(π1,3)sign(π2,4)sign(π1,4)sign(π2,3)\displaystyle=\ \frac{\operatorname{sign}(\mypi_{1,3})\operatorname{sign}(\mypi_{2,4})}{\operatorname{sign}(\mypi_{1,4})\operatorname{sign}(\mypi_{2,3})}
=(1)k1+k3(1)k2+k4(1)k1+k4(1)k2+k3sign(ϵ1,3)sign(ϵ2,4)sign(ϵ1,4)sign(ϵ2,3)\displaystyle=\ \frac{(-1)^{k_{1}+k_{3}}(-1)^{k_{2}+k_{4}}}{(-1)^{k_{1}+k_{4}}(-1)^{k_{2}+k_{3}}}\ \cdot\ \frac{\operatorname{sign}(\myepsilon_{1,3})\operatorname{sign}(\myepsilon_{2,4})}{\operatorname{sign}(\myepsilon_{1,4})\operatorname{sign}(\myepsilon_{2,3})}
=sign(ϵ1,3)sign(ϵ2,4)sign(ϵ1,4)sign(ϵ2,3).\displaystyle=\ {\operatorname{sign}(\myepsilon_{1,3})\operatorname{sign}(\myepsilon_{2,4})}{\operatorname{sign}(\myepsilon_{1,4})\operatorname{sign}(\myepsilon_{2,3})}.

Since the parity of ϵ(1)+ϵ(2)+1\myepsilon^{\prime}(1)+\myepsilon^{\prime}(2)+1 is even for every ϵV\myepsilon^{\prime}\in V, we can assume that ϵ\myepsilon is the representative that occurs in the definition of Σ\Sigma, i.e. we can assume that eiϵ(i)e_{i}\mapsto\myepsilon(i) defines an order preserving bijection {e1,,e4}{1,,4}\{e_{1},\dotsc,e_{4}\}\to\{1,\dotsc,4\}. Then ϵi,j\myepsilon_{i,j} is the identity if ϵ(i)<ϵ(j)\myepsilon(i)<\myepsilon(j) and ϵi,j=(eiej)\myepsilon_{i,j}=(e_{i}\ e_{j}) if ϵ(i)>ϵ(j)\myepsilon(i)>\myepsilon(j). Thus sign(ϵi,j)=1\operatorname{sign}(\myepsilon_{i,j})=1 if ϵ(i)<ϵ(j)\myepsilon(i)<\myepsilon(j) and sign(ϵi,j)=1\operatorname{sign}(\myepsilon_{i,j})=-1 if ϵ(i)>ϵ(j)\myepsilon(i)>\myepsilon(j).

Since [e1e2e3e4]Δ,J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\Delta,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\Delta,J}}{}{} is invariant under exchanging rows and columns, we can assume that e1e_{1} is the minimal element in {e1,,e4}\{e_{1},\dotsc,e_{4}\}, i.e. ϵ(1)=1\myepsilon(1)=1 and sign(ϵ1,j)=1\operatorname{sign}(\myepsilon_{1,j})=1 for j{3,4}j\in\{3,4\}. We verify the claim of the lemma by a case consideration for the value of ϵ(2)\myepsilon(2).

If ϵ(2)=2\myepsilon(2)=2, then e2e_{2} is minimal in {e2,e3,e4}\{e_{2},e_{3},e_{4}\} and sign(ϵ2,j)=1\operatorname{sign}(\myepsilon_{2,j})=1 for all j{3,4}j\in\{3,4\}. Thus

[e1e2e3e4]Δ,J= 1=(1)1+2+1=(1)ϵ(1)+ϵ(2)+1.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\Delta,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\Delta,J}}{}{}\ =\ 1\ =\ (-1)^{1+2+1}\ =\ (-1)^{\myepsilon(1)+\myepsilon(2)+1}.

If ϵ(2)=3\myepsilon(2)=3, then e3<e2<e4e_{3}<e_{2}<e_{4} or e4<e2<e3e_{4}<e_{2}<e_{3}. Thus sign(ϵ2,3)sign(ϵ2,4)=1\operatorname{sign}(\myepsilon_{2,3})\operatorname{sign}(\myepsilon_{2,4})=-1 and

[e1e2e3e4]Δ,J=1=(1)1+3+1=(1)ϵ(1)+ϵ(2)+1.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\Delta,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\Delta,J}}{}{}\ =\ -1\ =\ (-1)^{1+3+1}\ =\ (-1)^{\myepsilon(1)+\myepsilon(2)+1}.

If ϵ(2)=4\myepsilon(2)=4, then e2e_{2} is maximal in {e2,e3,e4}\{e_{2},e_{3},e_{4}\} and sign(ϵ2,j)=1\operatorname{sign}(\myepsilon_{2,j})=-1 for all j{3,4}j\in\{3,4\}. Thus

[e1e2e3e4]Δ,J=(1)2=(1)1+4+1=(1)ϵ(1)+ϵ(2)+1,\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\Delta,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\Delta,J}}{}{}\ =\ (-1)^{2}\ =\ (-1)^{1+4+1}\ =\ (-1)^{\myepsilon(1)+\myepsilon(2)+1},

which completes the proof. ∎

Let f:P𝕊f:P\to{\mathbb{S}} be a morphism of pastures. A lift of MM to PP (along ff) is a PP-matroid M^\widehat{M} such that f(M^)=Mf_{\ast}(\widehat{M})=M. In the following result, we will implicitly understand that a subfield kk of {\mathbb{R}} comes with the sign map sign:k𝕊\operatorname{sign}:k\to{\mathbb{S}}.

As explained in Corollary 4.13, the near-regular partial field 𝕌=𝔽1±x,y{x+y1}{\mathbb{U}}={{\mathbb{F}}_{1}^{\pm}}\langle x,y\rangle\!\sslash\!\{x+y-1\} admits three morphisms to 𝕊{\mathbb{S}}. Since the automorphism group Aut(𝕌)\operatorname{Aut}({\mathbb{U}}) acts transitively on these three morphisms, we can fix one of them without restricting the generality of our results. Thus we will implicitly understand that 𝕌{\mathbb{U}} comes with the morphism sign:𝕌𝕊\operatorname{sign}:{\mathbb{U}}\to{\mathbb{S}} given by sign(x)=sign(y)=1\operatorname{sign}(x)=\operatorname{sign}(y)=1.

Theorem 6.15.

Let MM be a positively oriented matroid whose underlying matroid M¯\underline{M} is without large uniform minors. Then M¯\underline{M} is near-regular and FM¯𝕌rF_{\underline{M}}\simeq{\mathbb{U}}^{\otimes r} for some r0r\geqslant 0. Up to rescaling equivalence, there are precisely 2r2^{r} lifts of MM to 𝕌{\mathbb{U}}, and for every subfield kk of {\mathbb{R}}, the lifts of MM to kk modulo rescaling equivalence correspond bijectively to ((0,1)k)r\Big{(}(0,1)\cap k\Big{)}^{r}.

Proof.

By Theorem 5.9, the foundation FM¯F_{\underline{M}} is isomorphic to a tensor product F1FrF_{1}\otimes\dotsb\otimes F_{r} of copies FiF_{i} of 𝔽2{\mathbb{F}}_{2} and symmetry quotients of 𝕌{\mathbb{U}}. The rescaling class of MM induces a morphism FM¯𝕊F_{\underline{M}}\to{\mathbb{S}}. Since there is no morphism from 𝔽2{\mathbb{F}}_{2} to 𝕊{\mathbb{S}}, each of the factors FiF_{i} has to be a symmetry quotient of 𝕌{\mathbb{U}}.

From the proof of Theorem 5.9, it follows that each symmetry quotient Fi=𝕌/HiF_{i}={\mathbb{U}}/H_{i} of 𝕌{\mathbb{U}} is the image of the induced morphism 𝕌FNFM{\mathbb{U}}\simeq F_{N}\to F_{M} of foundations for an embedded U24U^{2}_{4}-minor N=M\I/JN=M\backslash I/J of MM. This means that for every σHi\mysigma\in H_{i} and every (J;e1,,e4)ΩM(J;e_{1},\dotsc,e_{4})\in\Omega_{M}, we have an identity of universal cross ratios

[e1e2e3e4]J=[σ(e1)σ(e2)σ(e3)σ(e4)]J.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$\mysigma(e_{1})$}}&{\scalebox{0.9}{$\mysigma(e_{2})$}}\\[-2.0pt] {\scalebox{0.9}{$\mysigma(e_{3})$}}&{{\scalebox{0.9}{$\mysigma(e_{4})$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{\mysigma(e_{1})}&{\mysigma(e_{2})}\\ {\mysigma(e_{3})}&{\mysigma(e_{4})}\end{smallmatrix}\big{]}_{J}}{}{}.

We claim that if [e1e2e3e4]J=[e1e2e3e4]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}^{\prime}$}}&{\scalebox{0.9}{$e_{2}^{\prime}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}^{\prime}$}}&{{\scalebox{0.9}{$e_{4}^{\prime}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}^{\prime}}&{e_{2}^{\prime}}\\ {e_{3}^{\prime}}&{e_{4}^{\prime}}\end{smallmatrix}\big{]}_{J}}{}{} then Σ(e1,,e4)=Σ(e1,,e4)\Sigma(e_{1},\dotsc,e_{4})=\Sigma(e_{1}^{\prime},\dotsc,e_{4}^{\prime}), where Σ:ΩMS4/V\Sigma:\Omega_{M}^{\scalebox{0.7}{$\diamondsuit$}}\to S_{4}/V is the Ω\Omega-signature. We verify this in the following for all the defining relations of FM¯F_{\underline{M}} that involve non-degenerate cross ratios, as they appear in Theorem 4.20.

The relations (4.20) and (4.20) do not involve non-degenerate cross ratios (and (4.20) does not occur in our case since neither the Fano matroid not its dual are orientable). The relations (4.20), (4.20), (4.20) and (4.20) are already incorporated in 𝕌{\mathbb{U}} and can thus be ignored. For relation (4.20), it is obvious that both involved cross ratios have the same Ω\Omega-signature.

Thus we are left the relations (4.20) and (4.20). Since M¯\underline{M} is without large uniform minors, each of these relations reduces to an identity of two universal cross ratios. We begin with the tip relation (4.20), which is of the form

[e1e2e3e4]J=[e1e2e3e5]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{5}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{5}}\end{smallmatrix}\big{]}_{J}}{}{}

in our case, where we use (4.20) to express [e1e2e5e3]J1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{5}$}}&{{\scalebox{0.9}{$e_{3}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}^{-1}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{5}}&{e_{3}}\end{smallmatrix}\big{]}_{J}^{-1}}{}{} as [e1e2e3e5]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{5}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{5}}\end{smallmatrix}\big{]}_{J}}{}{}. After a permutation of {e1,,e4}\{e_{1},\dotsc,e_{4}\}, we can assume that e1<e4<e2<e3e_{1}<e_{4}<e_{2}<e_{3}, and thus

[e1e2e3e4]Δ,J=(1)1+3+1=1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\Delta,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\Delta,J}}{}{}\ =\ (-1)^{1+3+1}\ =\ -1

by Lemma 6.14. Therefore also [e1e2e3e5]Δ,J=1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{5}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\Delta,J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{5}}\end{smallmatrix}\big{]}_{\Delta,J}}{}{}=-1, which means that the unique order preserving bijection π:{e1,e2,e3,e5}{1,,4}\mypi:\{e_{1},e_{2},e_{3},e_{5}\}\to\{1,\dots,4\} must satisfy π(e1)=π(e2)\mypi(e_{1})=\mypi(e_{2}) according to Lemma 6.14. Since e1<e2<e3e_{1}<e_{2}<e_{3} by our assumptions, this implies that e1<e5<e2e_{1}<e_{5}<e_{2}. Thus Σ(e1,e2,e3,e4)=Σ(e1,e2,e3,e5)\Sigma(e_{1},e_{2},e_{3},e_{4})=\Sigma(e_{1},e_{2},e_{3},e_{5}).

The cotip relations (4.20) are in our case of the form

[e1e2e3e4]Je5=[e1e2e3e5]Je4.\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{Je_{5}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{Je_{5}}}{}{}\ =\ \mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{5}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{Je_{4}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{5}}\end{smallmatrix}\big{]}_{Je_{4}}}{}{}.

As before, we can assume that e1<e4<e2<e3e_{1}<e_{4}<e_{2}<e_{3} and thus [e1e2e3e4]Δ,Je5=1\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{\Delta,Je_{5}}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{\Delta,Je_{5}}}{}{}=-1. By the same reasoning, this implies that e1<e5<e2<e3e_{1}<e_{5}<e_{2}<e_{3} and thus Σ(e1,e2,e3,e4)=Σ(e1,e2,e3,e5)\Sigma(e_{1},e_{2},e_{3},e_{4})=\Sigma(e_{1},e_{2},e_{3},e_{5}). This establishes our claim that Σ(e1,,e4)=Σ(e1,,e4)\Sigma(e_{1},\dotsc,e_{4})=\Sigma(e_{1}^{\prime},\dotsc,e_{4}^{\prime}) whenever [e1e2e3e4]J=[e1e2e3e4]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}^{\prime}$}}&{\scalebox{0.9}{$e_{2}^{\prime}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}^{\prime}$}}&{{\scalebox{0.9}{$e_{4}^{\prime}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}^{\prime}}&{e_{2}^{\prime}}\\ {e_{3}^{\prime}}&{e_{4}^{\prime}}\end{smallmatrix}\big{]}_{J}}{}{}.

In particular, if [e1e2e3e4]J=[σ(e1)σ(e2)σ(e3)σ(e4)]J\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$e_{1}$}}&{\scalebox{0.9}{$e_{2}$}}\\[-2.0pt] {\scalebox{0.9}{$e_{3}$}}&{{\scalebox{0.9}{$e_{4}$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{e_{1}}&{e_{2}}\\ {e_{3}}&{e_{4}}\end{smallmatrix}\big{]}_{J}}{}{}=\mathchoice{\scalebox{1.3}{$\big{[}$}\,\raisebox{1.0pt}{$\begin{matrix}{\scalebox{0.9}{$\mysigma(e_{1})$}}&{\scalebox{0.9}{$\mysigma(e_{2})$}}\\[-2.0pt] {\scalebox{0.9}{$\mysigma(e_{3})$}}&{{\scalebox{0.9}{$\mysigma(e_{4})$}}}\end{matrix}$}\,\scalebox{1.3}{$\big{]}$}_{J}}{\big{[}\begin{smallmatrix}{\mysigma(e_{1})}&{\mysigma(e_{2})}\\ {\mysigma(e_{3})}&{\mysigma(e_{4})}\end{smallmatrix}\big{]}_{J}}{}{} then Σ(e1,,e4)=Σ(σ(e1),,σ(e4))\Sigma(e_{1},\dotsc,e_{4})=\Sigma\Big{(}{\mysigma(e_{1})},\dotsc,{\mysigma(e_{4})}\Big{)}, which means that σ\mysigma is in VV. These are precisely the relations in (4.20), which are already satisfied in 𝕌{\mathbb{U}}. We conclude that σ\mysigma is the identity on 𝕌{\mathbb{U}}.

This shows that every factor FiF_{i} of FM¯F_{\underline{M}} is a trivial quotient of 𝕌{\mathbb{U}} and thus FM¯𝕌rF_{\underline{M}}\simeq{\mathbb{U}}^{\otimes r}, as claimed in the theorem. It also implies at once that M¯\underline{M} is near-regular.

Let χM:FM¯𝕊\mychi_{M}:F_{\underline{M}}\to{\mathbb{S}} be the morphism of pastures induced by the rescaling class of MM. The lifts of MM to 𝕌{\mathbb{U}} and kk, up to rescaling, correspond to the lifts of χM\mychi_{M} to 𝕌{\mathbb{U}} and kk, respectively. We can study this question for each factor Fi=𝕌F_{i}={\mathbb{U}} of FM¯F_{\underline{M}} individually.

A lift of f:𝕌𝕊f:{\mathbb{U}}\to{\mathbb{S}} to 𝕌{\mathbb{U}} is a morphism f^:𝕌𝕌\hat{f}:{\mathbb{U}}\to{\mathbb{U}} such that sign(f^(x))=sign(f^(y))=1\operatorname{sign}\big{(}\hat{f}(x)\big{)}=\operatorname{sign}\big{(}\hat{f}(y)\big{)}=1. This determines f^\hat{f} up to a permutation of xx and yy, which shows that there are precisely two lifts of f:𝕌𝕊f:{\mathbb{U}}\to{\mathbb{S}} to 𝕌{\mathbb{U}}. Thus there are precisely 2r2^{r} lifts of MM to 𝕌{\mathbb{U}} up to rescaling equivalence.

A lift of f:𝕌𝕊f:{\mathbb{U}}\to{\mathbb{S}} to kk is a morphism f^:𝕌k\hat{f}:{\mathbb{U}}\to k such that sign(f^(x))=sign(f^(y))=1\operatorname{sign}\big{(}\hat{f}(x)\big{)}=\operatorname{sign}\big{(}\hat{f}(y)\big{)}=1. Since f^(y)=1f^(x)\hat{f}(y)=1-\hat{f}(x), this means that f^(x)((0,1)k)\hat{f}(x)\in\Big{(}(0,1)\cap k\Big{)} and, conversely, every choice of image f^(x)((0,1)k)\hat{f}(x)\in\Big{(}(0,1)\cap k\Big{)} determines a lift f^\hat{f} of ff to kk. Thus the lifts of MM to kk up to rescaling equivalence correspond bijectively to ((0,1)k)r\Big{(}(0,1)\cap k\Big{)}^{r}. This completes the proof of the theorem. ∎

6.6. Representation classes of matroids without large uniform minors

Given a matroid MM, we can ask over which pastures MM is representable. This defines a class of pastures that we call the representation class of MM.

For cardinality reasons, it is clear that not every class of pastures can be the representation class of a matroid. The theorems in Section 6.7 make clear that this fails in an even more drastic way—for example, a matroid that is representable over 𝔽2{\mathbb{F}}_{2} and 𝔽3{\mathbb{F}}_{3} is representable over all pastures; cf. Theorem 6.26.

In this section, we determine the representation classes that are defined by matroids without large uniform minors. It turns out that there are only twelve of them; see Table 3 for a characterization.

Definition 6.16.

Let MM be a matroid. The representation class of MM is the class 𝒫M{\mathcal{P}}_{M} of all pastures PP over which MM is representable. Two matroids MM and MM^{\prime} are representation equivalent if 𝒫M=𝒫M{\mathcal{P}}_{M}={\mathcal{P}}_{M^{\prime}}.

Note that the representation class 𝒫M{\mathcal{P}}_{M} of a matroid MM consists of precisely those pastures for which there is a morphism from the foundation FMF_{M} of MM to PP. This means that the representation class of a matroid is determined by its foundation. Evidently, 𝒫M=𝒫M{\mathcal{P}}_{M}={\mathcal{P}}_{M^{\prime}} if MM and MM^{\prime} are representation equivalent, which justifies the notation 𝒫C=𝒫M{\mathcal{P}}_{C}={\mathcal{P}}_{M} where CC is the representation class of MM.

Often there are simpler pastures than the foundation that characterize representation classes in the same way, which leads to the following notion.

Definition 6.17.

Let MM be a matroid with representation class 𝒫M{\mathcal{P}}_{M}. A characteristic pasture for MM is a pasture Π\Pi for which a pasture PP is in 𝒫M{\mathcal{P}}_{M} if and only if there is a morphism ΠP\Pi\to P. A matroid MM is strictly representable over a pasture PP if PP is a characteristic pasture for MM.

By the existence of the identity morphism id:ΠΠ\textup{id}:\Pi\to\Pi, strictly representable implies representable. And the foundation of a matroid MM is clearly a characteristic pasture for MM. The following result characterizes all characteristic pastures:

Lemma 6.18.

Let MM be a matroid with foundation FMF_{M}. A pasture Π\Pi is a characteristic pasture of MM if and only if there exist morphisms FMΠF_{M}\to\Pi and ΠFM\Pi\to F_{M}.

Proof.

Assume that Π\Pi is a characteristic pasture for MM. Since also FMF_{M} is a characteristic pasture, we have FM,Π𝒫MF_{M},\Pi\in{\mathcal{P}}_{M}, and by the defining property of characteristic pastures, there are morphisms FMΠF_{M}\to\Pi and ΠFM\Pi\to F_{M}.

Conversely, assume that there are morphisms FMΠF_{M}\to\Pi and ΠFM\Pi\to F_{M}. If P𝒫MP\in{\mathcal{P}}_{M}, then there is a morphism FMPF_{M}\to P, which yields a morphism ΠFMP\Pi\to F_{M}\to P. If there is a morphism ΠP\Pi\to P, then there is a morphism FMΠPF_{M}\to\Pi\to P, and thus P𝒫MP\in{\mathcal{P}}_{M}. This shows that Π\Pi is a characteristic pasture for MM. ∎

The next result describes an explicit condition for representation equivalent matroids.

Lemma 6.19.

Let MM and MM be two matroids with respective representation classes 𝒫M{\mathcal{P}}_{M} and 𝒫M{\mathcal{P}}_{M^{\prime}} and respective characteristic pastures Π\Pi and Π\Pi^{\prime}. Then 𝒫M{\mathcal{P}}_{M^{\prime}} is contained in 𝒫M{\mathcal{P}}_{M} if and only if there is a morphism ΠΠ\Pi\to\Pi^{\prime}. In particular, MM and NN are representation equivalent if and only if there exist morphisms ΠΠ\Pi\to\Pi^{\prime} and ΠΠ\Pi^{\prime}\to\Pi.

Proof.

If there is a morphism f:ΠΠf:\Pi\to\Pi^{\prime}, then we can compose every morphism ΠP\Pi^{\prime}\to P with ff, which implies that 𝒫M𝒫M{\mathcal{P}}_{M^{\prime}}\subset{\mathcal{P}}_{M}. Assume conversely that 𝒫M𝒫M{\mathcal{P}}_{M^{\prime}}\subset{\mathcal{P}}_{M}. Then Π𝒫M\Pi^{\prime}\in{\mathcal{P}}_{M}, which means that there is a morphism ΠΠ\Pi\to\Pi^{\prime}. The additional claim of the lemma is obvious. ∎

In the following, we say that a matroid MM is

  • strictly binary if 𝔽2{\mathbb{F}}_{2} is a characteristic pasture for MM;

  • strictly ternary if 𝔽3{\mathbb{F}}_{3} is a characteristic pasture for MM;

  • strictly near-regular if 𝕌{\mathbb{U}} is a characteristic pasture for MM;

  • strictly dyadic if 𝔻{\mathbb{D}} is a characteristic pasture for MM;

  • strictly hexagonal if {\mathbb{H}} is a characteristic pasture for MM;

  • strictly 𝔻{\mathbb{D}}\otimes{\mathbb{H}}-representable if 𝔻{\mathbb{D}}\otimes{\mathbb{H}} is a characteristic pasture for MM;

  • idempotent if 𝕂{\mathbb{K}} is a characteristic pasture for MM.

Note that an idempotent matroid MM is representable over a pasture PP if and only if PP is idempotent, by which we mean that both 1=1-1=1 and 1+1=11+1=1 hold in PP.

Theorem 6.20.

Let MM be a matroid without large uniform minors. Then MM belongs to precisely one of the 1212 classes that are described in Table 3. The six columns of Table 3 describe the following information:

  1. (1)

    a label for each class CC;

  2. (2)

    a name (as far as we have introduced one);

  3. (3)

    a characteristic pasture ΠC\Pi_{C} that is minimal in the sense that the foundation of every matroid MM in the class CC is of isomorphism type FMΠCF1FrF_{M}\simeq\Pi_{C}\otimes F_{1}\otimes\dotsb\otimes F_{r} for some r0r\geqslant 0 and F1,,Fr{𝕌,𝔻,}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}}\};

  4. (4)

    the type of factors FiF_{i} that can occur in the expression FMΠCF1FrF_{M}\simeq\Pi_{C}\otimes F_{1}\otimes\dotsb\otimes F_{r} for MM in CC;

  5. (5)

    a characterization of the pastures PP in the representation class 𝒫C{\mathcal{P}}_{C};

  6. (6)

    whether the matroids in this class are representable over some field.

The left diagram in Figure 4 illustrates the existence of morphisms between the different characteristic pastures ΠC\Pi_{C} in Table 3. The right diagram illustrates the inclusion relation between the representation classes 𝒫i=𝒫Ci{\mathcal{P}}_{i}={\mathcal{P}}_{C_{i}} (for i=1,,12i=1,\dotsc,12)—an edge indicates that the class on the bottom end of the edge is contained in the class at the top end of the edge.

Table 3. The equivalence classes of matroids without large uniform minors
CC Name minimal ΠC\Pi_{C} add. FiF_{i} P𝒫CP\in{\mathcal{P}}_{C} iff. u,vP×\exists u,v\in P^{\times} s.t. field?
C1C_{1} regular 𝔽1±{{\mathbb{F}}_{1}^{\pm}} yes
C2C_{2} str. near-regular 𝕌{\mathbb{U}} 𝕌{\mathbb{U}} u+v=1u+v=1 yes
C3C_{3} strictly dyadic 𝔻{\mathbb{D}} 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}} u+u=1u+u=1 yes
C4C_{4} str. hexagonal {\mathbb{H}} 𝕌{\mathbb{U}}, {\mathbb{H}} vv2=v3=1v-v^{2}=-v^{3}=1 yes
C5C_{5} str. 𝔻{\mathbb{D}}\otimes{\mathbb{H}}-repr. 𝔻{\mathbb{D}}\otimes{\mathbb{H}} 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}} u+u=vv2=v3=1u+u=v-v^{2}=-v^{3}=1 yes
C6C_{6} strictly ternary 𝔽3{\mathbb{F}}_{3} 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}} 1+1=11+1=1 yes
C7C_{7} strictly binary 𝔽2{\mathbb{F}}_{2} 1=1-1=1 yes
C8C_{8} 𝔽2𝕌{\mathbb{F}}_{2}\otimes{\mathbb{U}} 𝕌{\mathbb{U}} 1=u+v=1-1=u+v=1 yes
C9C_{9} 𝔽2𝔻{\mathbb{F}}_{2}\otimes{\mathbb{D}} 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}} 1=u+u=1-1=u+u=1 no
C10C_{10} 𝔽2{\mathbb{F}}_{2}\otimes{\mathbb{H}} 𝕌{\mathbb{U}}, {\mathbb{H}} 1=vv2=v3=1-1=v-v^{2}=v^{3}=1 yes
C11C_{11} 𝔽2𝔻{\mathbb{F}}_{2}\otimes{\mathbb{D}}\otimes{\mathbb{H}} 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}} 1=u+u=vv2=v3=1-1=u+u=v-v^{2}=v^{3}=1 no
C12C_{12} idempotent 𝔽2𝔽3{\mathbb{F}}_{2}\otimes{\mathbb{F}}_{3} 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}} 1=1+1=1-1=1+1=1 no
𝔽1±{{\mathbb{F}}_{1}^{\pm}}𝕌{\mathbb{U}}𝔻{\mathbb{D}}{\mathbb{H}}𝔻{\mathbb{D}}\otimes{\mathbb{H}}𝔽3{\mathbb{F}}_{3}𝔽2{\mathbb{F}}_{2}𝔽2𝕌{\mathbb{F}}_{2}\otimes{\mathbb{U}}𝔽2𝔻{\mathbb{F}}_{2}\otimes{\mathbb{D}}𝔽2{\mathbb{F}}_{2}\otimes{\mathbb{H}}𝔽2𝔻{\mathbb{F}}_{2}\otimes{\mathbb{D}}\otimes{\mathbb{H}}𝔽2𝔽3{\mathbb{F}}_{2}\otimes{\mathbb{F}}_{3}
𝒫1{\mathcal{P}}_{1}𝒫2{\mathcal{P}}_{2}𝒫3{\mathcal{P}}_{3}𝒫4{\mathcal{P}}_{4}𝒫5{\mathcal{P}}_{5}𝒫6{\mathcal{P}}_{6}𝒫7{\mathcal{P}}_{7}𝒫8{\mathcal{P}}_{8}𝒫9{\mathcal{P}}_{9}𝒫10{\mathcal{P}}_{10}𝒫11{\mathcal{P}}_{11}𝒫12{\mathcal{P}}_{12}
Figure 4. Morphisms between characteristic pastures and containment of the representation classes for matroids without large uniform minors
Proof.

For the sake of this proof, we say that two pastures PP and PP^{\prime} are equivalent, and write PPP\sim P^{\prime}, if there are morphisms PPP\to P^{\prime} and PPP^{\prime}\to P.

If there is a morphism PPP^{\prime}\to P, then there are morphisms PPPP\to P\otimes P^{\prime} and PPPP\otimes P^{\prime}\to P, which means that PPPP\otimes P^{\prime}\sim P. This applies in particular to P=PP^{\prime}=P. This shows that P1PrP1PsP_{1}\otimes\dotsb\otimes P_{r}\sim P_{1}\otimes\dotsb\otimes P_{s} for srs\leqslant r and pastures P1,,PrP_{1},\dotsc,P_{r} if, for every i{s+1,,r}i\in\{s+1,\dotsc,r\}, there is a j{1,,r}j\in\{1,\dotsc,r\} and a morphism PiPjP_{i}\to P_{j}.

Let MM be a matroid without large uniform minors and FMF_{M} its foundation. By Theorem 5.9, FMF1FrF_{M}\simeq F_{1}\otimes\dotsb\otimes F_{r} for some F1,,Fr{𝕌,𝔻,,𝔽3,𝔽2}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}},{\mathbb{F}}_{3},{\mathbb{F}}_{2}\}, where we can assume that 𝔽2{\mathbb{F}}_{2} appears at most once as a factor. By the previous considerations, FMF1FsF_{M}\sim F_{1}\otimes\dotsb\otimes F_{s} for pairwise distinct F1,,Fs{𝕌,𝔻,,𝔽3,𝔽2}F_{1},\dotsc,F_{s}\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}},{\mathbb{F}}_{3},{\mathbb{F}}_{2}\}. Since there are morphisms

𝔻{{\mathbb{D}}}𝕌{{\mathbb{U}}}𝔽3,{{\mathbb{F}}_{3},}{{\mathbb{H}}}

we have 𝔻𝕌𝔻{\mathbb{D}}\otimes{\mathbb{U}}\sim{\mathbb{D}}, 𝕌{\mathbb{H}}\otimes{\mathbb{U}}\sim{\mathbb{H}} and 𝔽3F𝔽3{\mathbb{F}}_{3}\otimes F\sim{\mathbb{F}}_{3} for F{𝕌,𝔻,}F\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}}\}. Thus we can assume that in the expression F1FsF_{1}\otimes\dotsb\otimes F_{s} at most one of 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}} and 𝔽3{\mathbb{F}}_{3} appears, with the exception of 𝔻{\mathbb{D}}\otimes{\mathbb{H}}.

Thus we are limited to the twelve different expressions for F1FsF_{1}\otimes\dotsb\otimes F_{s} that appear in Figure 4. We conclude that FMF_{M} is equivalent to one of those and that Π=F1Fs\Pi=F_{1}\otimes\dotsb\otimes F_{s} is a characteristic pasture for MM.

An easy case-by-case verification based on Table 2, which we shall not carry out, shows that there is a morphism between two pastures if and only if there is a directed path between these pastures in the diagram on the left hand side of Figure 4. By Lemma 6.19, this diagram determines at once the inclusion behaviour of the associated representation classes 𝒫1{\mathcal{P}}_{1}𝒫12{\mathcal{P}}_{12} as illustrated on the right hand side of Figure 4.

Note that the way we found the twelve characteristic pastures Π\Pi shows that they are minimal in the sense of part (3) of the theorem, and it shows that the types of additional factors displayed in the forth column of Table 3 are correct. The conditions in the fifth column of Table 3 follows at once from Lemma 6.1.

For the verification of the last column, note that there is a morphism ΠC𝔽3\Pi_{C}\to{\mathbb{F}}_{3} for the classes C{C1,,C6}C\in\{C_{1},\dotsc,C_{6}\} and that there is a morphism Π𝔽4\Pi\to{\mathbb{F}}_{4} for C{C7,C8,C10}C\in\{C_{7},C_{8},C_{10}\}. Thus the matroids in the classes C1C_{1}C8C_{8} and C10C_{10} are representable over a field. There is no morphism from 𝔽2𝔻{\mathbb{F}}_{2}\otimes{\mathbb{D}} to any field since in a field only one of 1+1=01+1=0 and 1+1=z11+1=z^{-1} for some z0z\neq 0 can hold. Thus matroids in the classes C9C_{9}, C11C_{11} and C12C_{12} are not representable over any field, which concludes the proof of the theorem. ∎

As a sample application, we formulate the following strengthening of the result [37, Thm. 3.3] by Whittle. Recall that a matroid is called representable if it is representable over some field.

Theorem 6.21.

Let 𝒫8={𝔽q|q8 a prime power}{\mathcal{P}}_{\leqslant 8}=\big{\{}{\mathbb{F}}_{q}\,\big{|}\,q\leqslant 8\text{ a prime power}\big{\}}. Then two representable matroids MM and MM^{\prime} without large uniform minors are representation equivalent if and only if 𝒫M𝒫8=𝒫M𝒫8{\mathcal{P}}_{M}\cap{\mathcal{P}}_{\leqslant 8}={\mathcal{P}}_{M^{\prime}}\cap{\mathcal{P}}_{\leqslant 8}. More precisely, for i{1,,8,10}i\in\{1,\dotsc,8,10\} and pip_{i} and qiq_{i} as in Table 4, the class 𝒫Ci{\mathcal{P}}_{C_{i}} is the intersection of the representation classes 𝒫M{\mathcal{P}}_{M} of all matroids MM without large uniform minors that are representable over 𝔽pi{\mathbb{F}}_{p_{i}} and 𝔽qi{\mathbb{F}}_{q_{i}}.

Table 4. Prime powers such that 𝒫Ci={𝒫M|M is representable over 𝔽pi and 𝔽qi}{\mathcal{P}}_{C_{i}}=\bigcap\big{\{}{\mathcal{P}}_{M}\,\big{|}\,M\text{ is representable over ${\mathbb{F}}_{p_{i}}$ and ${\mathbb{F}}_{q_{i}}$}\big{\}}
ii 1 2 3 4 5 6 7 8 10
pip_{i} 2 3 3 3 3 3 2 8 4
qiq_{i} 3 8 5 4 7 3 2 8 4
Proof.

For i{1,,8,10}i\in\{1,\dotsc,8,10\} and MM in CiC_{i}, let 𝒰i{\mathcal{U}}_{i} be the subset of {𝕌,𝔻,,𝔽3,𝔽2}\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}},{\mathbb{F}}_{3},{\mathbb{F}}_{2}\} such that Πi=P𝒰iP\Pi_{i}=\bigotimes_{P\in{\mathcal{U}}_{i}}P is a characteristic pasture for MM, cf. Table 3. Then we can read off from Table 2 that there are morphisms P𝔽piP\to{\mathbb{F}}_{p_{i}} and P𝔽qiP\to{\mathbb{F}}_{q_{i}} for all P𝒰iP\in{\mathcal{U}}_{i}, and that for all P{𝕌,𝔻,,𝔽3,𝔽2}P\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}},{\mathbb{F}}_{3},{\mathbb{F}}_{2}\} that are not in 𝒰i{\mathcal{U}}_{i}, there is either no morphism from PP to 𝔽pi{\mathbb{F}}_{p_{i}} or no morphism from PP to 𝔽qi{\mathbb{F}}_{q_{i}}. This shows that the existence of morphisms into 𝔽pi{\mathbb{F}}_{p_{i}} and 𝔽qi{\mathbb{F}}_{q_{i}} characterize the factors of the characteristic pasture Πi\Pi_{i} and establishes the claims of the theorem. ∎

Remark 6.22.

Note that the representation class 𝒫1{\mathcal{P}}_{1} of regular matroids contains all pastures and is therefore the largest possible representation class. The representation class 𝒫12{\mathcal{P}}_{12} of idempotent pastures is the smallest representation class, since every matroid is by definition representable over 𝕂{\mathbb{K}} and thus over every idempotent pasture. (Recall that a pasture PP is called idempotent if there is a morphism from 𝕂{\mathbb{K}} to PP.) Every other representation class thus lies between 𝒫12{\mathcal{P}}_{12} and 𝒫1{\mathcal{P}}_{1}.

Remark 6.23.

We will show in a sequel to this paper that every tensor product of copies of the pastures 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}}, 𝔽3{\mathbb{F}}_{3} and 𝔽2{\mathbb{F}}_{2} occurs as the foundation of a matroid. Consequently each of the classes C1C_{1}C12C_{12} is nonempty.

Alternatively, we can use known results to deduce this. Since there are matroids that are regular, strictly near-regular (e.g. U24U^{2}_{4}), strictly dyadic (e.g. the non-Fano matroid F7F_{7}^{-}), strictly hexagonal (e.g. the ternary affine plane AG(2,3)AG(2,3)), strictly ternary (e.g. the matroid T8T_{8} from Oxley’s book [21]) and strictly binary (e.g. the Fano matroid F7F_{7}), the classes C1C_{1}, C2C_{2}, C3C_{3}, C4C_{4}, C6C_{6} and C7C_{7} are nonempty.

Since the characteristic pastures of the remaining classes in Table 3 are tensor products of characteristic pastures of one of the aforementioned matroids, we can deduce that the other classes are also nonempty by observing that

{P|FMFMP}={P|FMP}{P|FMP}=𝒫M𝒫M=𝒫MM\big{\{}P\,\big{|}\,F_{M}\otimes F_{M^{\prime}}\stackrel{{\scriptstyle\scalebox{0.7}{$\exists$}}}{{\rightarrow}}P\big{\}}\ =\ \big{\{}P\,\big{|}\,F_{M}\stackrel{{\scriptstyle\scalebox{0.7}{$\exists$}}}{{\rightarrow}}P\big{\}}\cap\big{\{}P\,\big{|}\,F_{M^{\prime}}\stackrel{{\scriptstyle\scalebox{0.7}{$\exists$}}}{{\rightarrow}}P\big{\}}\ =\ {\mathcal{P}}_{M}\cap{\mathcal{P}}_{M^{\prime}}\ =\ {\mathcal{P}}_{M\oplus M^{\prime}}

for two matroids MM and MM^{\prime}.

Remark 6.24.

Since all binary and ternary matroids are without large uniform minors, all matroids in the classes C1C_{1}C7C_{7} are without large uniform minors. This is not true for all classes though. For instance the direct sum of an idempotent matroid with U25U^{2}_{5} is also idempotent and thus in C12C_{12}, but has a minor of type U25U^{2}_{5}; cf. Remark 6.23 for the existence of idempotent matroids.

In fact, a similar construction yield matroids with U25U^{2}_{5}-minors in the classes C10C_{10} and C11C_{11}. By contrast, all matroids in C8C_{8} and C9C_{9} are without large uniform minors. This latter fact can be proven as follows: a class CiC_{i} contains a matroid MM with a U25U^{2}_{5}- or a U35U^{3}_{5}-minor if and only if there is morphism from the foundation of U25U^{2}_{5} (cf. Proposition 5.4) to the minimal characteristic pasture for MM. There is no morphism from the foundation of U25U^{2}_{5} to 𝔽2𝕌{\mathbb{F}}_{2}\otimes{\mathbb{U}} or to 𝔽2𝔻{\mathbb{F}}_{2}\otimes{\mathbb{D}}, but there are morphisms to 𝔽2{\mathbb{F}}_{2}\otimes{\mathbb{H}} and 𝔽2𝔻{\mathbb{F}}_{2}\otimes{\mathbb{D}}\otimes{\mathbb{H}}.

6.7. Characterization of classes of matroids

In this section, we use our results to provide different characterizations of some prominent classes of matroids, such as regular, near-regular, binary, ternary, quaternary, dyadic, and hexagonal matroids. In particular, we find new proofs for results by Tutte, Bland and Las Vergnas, and Whittle, which we refer to in detail at the beginnings of the appropriate sections. Moreover, we obtain new characterizations, which often involve the pastures 𝕊{\mathbb{S}}, {\mathbb{P}} and 𝕎{\mathbb{W}}.

All these characterizations are immediate applications of Theorem 5.9 in combination with Table 2. It is possible to work out additional descriptions for the classes of matroids under consideration, or to study other classes with the same techniques. For example, our technique allows for an easy proof of the following results found in Theorems 5.1 and 5.2 of Semple and Whittle’s paper [27].

Theorem 6.25 (Semple–Whittle ’96).

Let 𝒞P{\mathcal{C}}_{P} denote the class of matroids without large uniform minors that are representable over a pasture PP. Then the following hold true.

  1. (1)

    𝒞𝔽2r𝒞𝔽3=𝒞𝕌{\mathcal{C}}_{{\mathbb{F}}_{2^{r}}}\cap{\mathcal{C}}_{{\mathbb{F}}_{3}}={\mathcal{C}}_{\mathbb{U}} for odd r2r\geqslant 2.

  2. (2)

    𝒞𝔽2r𝒞𝔽3=𝒞{\mathcal{C}}_{{\mathbb{F}}_{2^{r}}}\cap{\mathcal{C}}_{{\mathbb{F}}_{3}}={\mathcal{C}}_{\mathbb{H}} for even r2r\geqslant 2.

  3. (3)

    𝒞k𝒞𝔽3{\mathcal{C}}_{k}\subset{\mathcal{C}}_{{\mathbb{F}}_{3}} for every field kk of characteristic different from 22, and 𝒞k=𝒞𝔻{\mathcal{C}}_{k}={\mathcal{C}}_{\mathbb{D}} if, in addition, kk does not contain a primitive sixth root of unity.

6.7.1. Regular matroids

The following theorem extends a number of classical results that characterize regular matroids, namely as binary matroids that are representable over a field kk with chark2\textup{char}\;k\neq 2 by Tutte in [31] and [32] (use P=kP=k in (5)) and as binary and orientable matroids by Bland and Las Vergnas in [8] (use P=𝕊P={\mathbb{S}} in (5)). Up to the characterization (3), the authors of this paper have proven Theorem 6.26 in its full generality in [5, Thm. 7.33] with a slightly different proof.

Theorem 6.26.

Let MM be a matroid with foundation FMF_{M}. Then the following assertions are equivalent:

  1. (1)

    MM is regular.

  2. (2)

    FM=𝔽1±F_{M}={{\mathbb{F}}_{1}^{\pm}}.

  3. (3)

    MM belongs to C1C_{1}.

  4. (4)

    MM is representable over all pastures.

  5. (5)

    MM is representable over 𝔽2{\mathbb{F}}_{2} and a pasture with 11-1\neq 1.

Proof.

The logical structure of this proof is (1)\Rightarrow(3)\Rightarrow(4)\Rightarrow(5)\Rightarrow(2)\Rightarrow(1). The implications (2)\Rightarrow(1)\Rightarrow(3)\Rightarrow(4) follow from Theorem 6.20 and (4)\Rightarrow(5) is trivial.

We close the circle by showing (5)\Rightarrow(2). If MM is binary, then it is without large uniform minors by Lemma 6.2. Thus, by Theorem 5.9, FMF_{M} is a tensor product of copies of 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}}, 𝔽3{\mathbb{F}}_{3} and 𝔽2{\mathbb{F}}_{2}. But none of 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}} or 𝔽3{\mathbb{F}}_{3} admits a morphism to 𝔽2{\mathbb{F}}_{2}, and 𝔽2{\mathbb{F}}_{2} admits no morphism into a pasture PP with 11-1\neq 1. Thus FM=𝔽1±F_{M}={{\mathbb{F}}_{1}^{\pm}}, as claimed. ∎

6.7.2. Binary matroids

We find the following equivalent characterizations of binary matroids.

Theorem 6.27.

Let MM be a matroid with foundation FMF_{M}. Then the following assertions are equivalent:

  1. (1)

    MM is binary.

  2. (2)

    FM𝔽1±F_{M}\simeq{{\mathbb{F}}_{1}^{\pm}} or FM𝔽2F_{M}\simeq{\mathbb{F}}_{2}.

  3. (3)

    MM belongs to C1C_{1} or C7C_{7}.

  4. (4)

    MM is representable over every pasture for which 1=1-1=1.

  5. (5)

    All fundamental elements of FMF_{M} are trivial.

Proof.

We prove (1)\Rightarrow(3)\Rightarrow(2)\Rightarrow(5)\Rightarrow(2)\Rightarrow(4)\Rightarrow(1). Steps (1)\Rightarrow(3)\Rightarrow(2) follow from Theorem 6.20, step (5)\Rightarrow(2) follows from part (1) of Lemma 6.1 and Corollary 5.10, and steps (2)\Rightarrow(5) and (2)\Rightarrow(4)\Rightarrow(1) are trivial. ∎

6.7.3. Ternary matroids

We find the following equivalent characterizations of ternary matroids.

Theorem 6.28.

Let MM be a matroid with foundation FMF_{M}. Then the following assertions are equivalent:

  1. (1)

    MM is ternary.

  2. (2)

    FMF1FrF_{M}\simeq F_{1}\otimes\dotsb\otimes F_{r} for r0r\geqslant 0 and F1,,Fr{𝕌,𝔻,,𝔽3}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}},{\mathbb{F}}_{3}\}.

  3. (3)

    MM belongs to one of C1C_{1}C6C_{6}.

  4. (4)

    MM is representable over every pasture for which 1+1+1=01+1+1=0.

  5. (5)

    MM is without large uniform minors and representable over a field of characteristic 33.

  6. (6)

    MM is without large uniform minors and weakly orientable.

  7. (7)

    MM is without large uniform minors and there is no morphism from 𝔽2{\mathbb{F}}_{2} to FMF_{M}.

Proof.

We show (2)\Leftrightarrow(3), (1)\Leftrightarrow(4) and (2)\Rightarrow(1)\Rightarrow(5) / (6) / (7)\Rightarrow(2). The implications (2)\Rightarrow(1)\Leftrightarrow(4) are trivial. The equivalence (2)\Leftrightarrow(3) follows from Theorem 6.20.

Assuming (1), then MM is without large uniform minors by Lemma 6.2. Since there are morphisms 𝔽3k{\mathbb{F}}_{3}\to k for every field kk of characteristic 33 and 𝔽3𝕎{\mathbb{F}}_{3}\to{\mathbb{W}}, this implies (5) and (6).

If MM is without large uniform minors, then Theorem 5.9 implies that FMF_{M} is the tensor product of copies of 𝕌{\mathbb{U}}, 𝔻{\mathbb{D}}, {\mathbb{H}}, 𝔽3{\mathbb{F}}_{3} and 𝔽2{\mathbb{F}}_{2}. Thus (1) and the fact that 𝔽2{\mathbb{F}}_{2} does not map to 𝔽3{\mathbb{F}}_{3} implies (7). Conversely, each condition of (5), (6) and (7) implies that 𝔽2{\mathbb{F}}_{2} cannot occur as a building block of FMF_{M}, and thus (2). ∎

6.7.4. Quaternary matroids without large uniform minors

We find the following equivalent characterizations of quaternary matroids without large uniform minors.

Theorem 6.29.

Let MM be a matroid without large uniform minors and FMF_{M} its foundation. Then the following assertions are equivalent:

  1. (1)

    MM is quaternary.

  2. (2)

    FMF1FrF_{M}\simeq F_{1}\otimes\dotsb\otimes F_{r} for r0r\geqslant 0 and F1,,Fr{𝕌,,𝔽2}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{H}},{\mathbb{F}}_{2}\}.

  3. (3)

    MM belongs to C1C_{1}, C2C_{2}, C4C_{4}, C7C_{7}, C8C_{8} or C10C_{10}.

  4. (4)

    MM is representable over every pasture for which 1+1=01+1=0 and that contains an element uu for which u2+u+1=0u^{2}+u+1=0.

  5. (5)

    MM is representable over all field extensions of 𝔽4{\mathbb{F}}_{4}.

  6. (6)

    There is no morphism from 𝔻{\mathbb{D}} to FMF_{M}.

Proof.

We show (2)\Leftrightarrow(3) and (2)\Rightarrow(4)\Rightarrow(1)\Rightarrow(5)\Rightarrow(6)\Rightarrow(2). The equivalence (2)\Leftrightarrow(3) follows from Theorem 6.20. The implications (2)\Rightarrow(4)\Rightarrow(1)\Rightarrow(5) are trivial. The implication (5)\Rightarrow(6) follows since there is no morphism from 𝔻{\mathbb{D}} to 𝔽4{\mathbb{F}}_{4} by Lemma 6.1. The implication (6)\Rightarrow(2) follows by Theorem 5.9, together with the fact that there is a morphism 𝔻𝔽3{\mathbb{D}}\to{\mathbb{F}}_{3} but not to 𝕌{\mathbb{U}}, {\mathbb{H}} and 𝔽2{\mathbb{F}}_{2}, and thus only the latter three pastures can occur as factors of FMF_{M}. ∎

6.7.5. Near-regular matroids

In this section, we provide several characterizations of near-regular matroids. The descriptions (5) and (6) appear in Whittle’s paper [36, Thm. 1.4].

Theorem 6.30.

Let MM be a matroid with foundation FMF_{M}. Then the following assertions are equivalent:

  1. (1)

    MM is near-regular.

  2. (2)

    FMF1FrF_{M}\simeq F_{1}\otimes\dotsb\otimes F_{r} for r0r\geqslant 0 and F1==Fr=𝕌F_{1}=\dotsb=F_{r}={\mathbb{U}}.

  3. (3)

    MM belongs to C1C_{1} or C2C_{2}.

  4. (4)

    MM is representable over all pastures with a fundamental element.

  5. (5)

    MM is representable over fields with at least 33 elements.

  6. (6)

    MM is representable over 𝔽3{\mathbb{F}}_{3} and 𝔽8{\mathbb{F}}_{8}.

  7. (7)

    MM is without large uniform minors and representable over 𝔽4{\mathbb{F}}_{4} and 𝔽5{\mathbb{F}}_{5}.

  8. (8)

    MM is without large uniform minors and representable over 𝔽4{\mathbb{F}}_{4} and 𝕊{\mathbb{S}}.

  9. (9)

    MM is without large uniform minors and representable over 𝔽8{\mathbb{F}}_{8} and 𝕎{\mathbb{W}}.

  10. (10)

    MM is dyadic and hexagonal.

  11. (11)

    MM is without large uniform minors and there are no morphisms 𝔽2FM{\mathbb{F}}_{2}\to F_{M}, 𝔻FM{\mathbb{D}}\to F_{M}, or FM{\mathbb{H}}\to F_{M}.

Proof.

We show (2)\Leftrightarrow(3), (2)\Rightarrow(1)\Rightarrow(4)\Rightarrow(5)\Rightarrow(2) and the equivalence of (2) with each of (6)–(11). The equivalence (2)\Leftrightarrow(3) follows from Theorem 6.20, (2)\Rightarrow(1) and (4)\Rightarrow(5) are trivial and (1)\Rightarrow(4) follows from Lemma 6.1. That (2) implies (6)–(11) can be read off from Table 2. Conversely, each of (5)–(11) implies that MM is without large uniform minors and thus Theorem 5.9 applies. In turn, each of (5)–(11) excludes that any of 𝔻{\mathbb{D}}, {\mathbb{H}}, 𝔽3{\mathbb{F}}_{3} and 𝔽2{\mathbb{F}}_{2} occur as a factor FMF_{M}, and thus (2). ∎

6.7.6. Dyadic matroids

In this section, we provide several characterizations of dyadic matroids. Description (6) has been given by Whittle in [35, Thm. 7.1]. Descriptions (4) and (5) have been given by Whittle in [36, Thm. 1.1].

Theorem 6.31.

Let MM be a matroid with foundation FMF_{M}. Then the following assertions are equivalent:

  1. (1)

    MM is dyadic.

  2. (2)

    FMF1FrF_{M}\simeq F_{1}\otimes\dotsb\otimes F_{r} for r0r\geqslant 0 and F1,,Fr{𝕌,𝔻}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{D}}\}.

  3. (3)

    MM belongs to C1C_{1}, C2C_{2} or C3C_{3}.

  4. (4)

    MM is representable over every pasture PP such that 1+1=u1+1=u for some uP×u\in P^{\times}.

  5. (5)

    MM is representable over every field of characteristic different from 22.

  6. (6)

    MM is representable over 𝔽3{\mathbb{F}}_{3} and 𝔽q{\mathbb{F}}_{q}, where qq is an odd prime power such that q1q-1 is not divisible by 33.

  7. (7)

    MM is representable over 𝔽3{\mathbb{F}}_{3} and {\mathbb{Q}}.

  8. (8)

    MM is representable over 𝔽3{\mathbb{F}}_{3} and 𝕊{\mathbb{S}}.

  9. (9)

    MM is without large uniform minors and there are no morphisms 𝔽2FM{\mathbb{F}}_{2}\to F_{M} or FM{\mathbb{H}}\to F_{M}.

Proof.

We show (2)\Leftrightarrow(3), (2)\Rightarrow(1)\Rightarrow(4)\Rightarrow(5)\Rightarrow(2) and the equivalence of (2) with each of (6)–(9). The equivalence (2)\Leftrightarrow(3) follows from Theorem 6.20, (2)\Rightarrow(1) and (4)\Rightarrow(5) are trivial and (1)\Rightarrow(4) follows from Lemma 6.1. That (2) implies (6)–(9) follows from Lemma 6.1 and Table 2. Conversely, each of (5)–(9) implies that MM is without large uniform minors and thus Theorem 5.9 applies. In turn, each of (5)–(9) excludes that any of {\mathbb{H}}, 𝔽3{\mathbb{F}}_{3} and 𝔽2{\mathbb{F}}_{2} occur as a factor FMF_{M}, and thus (2). ∎

6.7.7. Hexagonal matroids

In this section, we provide several characterizations of hexagonal matroids. Description (5) has been given by Whittle in [36, Thm. 1.2].

Theorem 6.32.

Let MM be a matroid with foundation FMF_{M}. Then the following assertions are equivalent:

  1. (1)

    MM is hexagonal.

  2. (2)

    FMF1FrF_{M}\simeq F_{1}\otimes\dotsb\otimes F_{r} for r0r\geqslant 0 and F1,,Fr{𝕌,}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{H}}\}.

  3. (3)

    MM belongs to C1C_{1}, C2C_{2} or C4C_{4}.

  4. (4)

    MM is representable over every pasture that contains an element uu with u3=1u^{3}=-1 and u2u+1=0u^{2}-u+1=0.

  5. (5)

    MM is representable over every field that is of characteristic 33 or contains a primitive sixth root of unity.

  6. (6)

    MM is representable over 𝔽3{\mathbb{F}}_{3} and 𝔽4{\mathbb{F}}_{4}.

  7. (7)

    MM is without large uniform minors, weakly orientable, and representable over 𝔽4{\mathbb{F}}_{4}.

  8. (8)

    MM is without large uniform minors and there are no morphisms 𝔽2FM{\mathbb{F}}_{2}\to F_{M} or 𝔻FM{\mathbb{D}}\to F_{M}.

Proof.

We show (2)\Leftrightarrow(3), (2)\Rightarrow(1)\Rightarrow(4)\Rightarrow(5)\Rightarrow(2) and the equivalence of (2) with each of (6)–(8). The equivalence (2)\Leftrightarrow(3) follows from Theorem 6.20, (2)\Rightarrow(1) and (4)\Rightarrow(5) are trivial and (1)\Rightarrow(4) follows from Lemma 6.1. That (2) implies (6)–(8) follows from Lemma 6.1 and Table 2. Conversely, each of (5)–(8) implies that MM is without large uniform minors and thus Theorem 5.9 applies. In turn, each of (5)–(8) excludes that any of 𝔻{\mathbb{D}}, 𝔽3{\mathbb{F}}_{3} and 𝔽2{\mathbb{F}}_{2} occur as a factor FMF_{M}, and thus (2). ∎

6.7.8. 𝔻{\mathbb{D}}\otimes{\mathbb{H}}-representable matroids

Whittle describes in [36, Thm. 1.3] equivalent conditions that are satisfied by 𝔻{\mathbb{D}}\otimes{\mathbb{H}}-representable matroids, which are conditions (4) and (5) below. We augment Whittle’s result with the following theorem.

Theorem 6.33.

Let MM be a matroid with foundation FMF_{M}. Then the following assertions are equivalent:

  1. (1)

    MM is 𝔻{\mathbb{D}}\otimes{\mathbb{H}}-representable.

  2. (2)

    FMF1FrF_{M}\simeq F_{1}\otimes\dotsb\otimes F_{r} for r0r\geqslant 0 and F1,,Fr{𝕌,𝔻,}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}}\}.

  3. (3)

    MM belongs to one of C1C_{1}C5C_{5}.

  4. (4)

    MM is representable over 𝔽3{\mathbb{F}}_{3} and {\mathbb{C}}.

  5. (5)

    MM is representable over 𝔽3{\mathbb{F}}_{3} and 𝔽q{\mathbb{F}}_{q}, where qq is an odd prime power congruent to 11 modulo 33.

  6. (6)

    MM is representable over 𝔽3{\mathbb{F}}_{3} and {\mathbb{P}}.

Proof.

We show (1)\Rightarrow(2)\Rightarrow(3)\Rightarrow(1) and the equivalence of (2) with each of (4)–(6). The implications (1)\Rightarrow(2)\Rightarrow(3)\Rightarrow(1) follow from Theorem 6.20. That (2) implies (4)–(6) follows from Lemma 6.1 and Table 2. Conversely, each of (4)–(6) implies that MM is without large uniform minors by Lemma 6.2, and thus Theorem 5.9 applies. In turn, each of (4)–(6) excludes the possibility that either 𝔽3{\mathbb{F}}_{3} or 𝔽2{\mathbb{F}}_{2} occurs as a factor FMF_{M}, and thus (2). ∎

6.7.9. Representable matroids without large uniform minors

As a final application, we find the following equivalent characterization of matroids without large uniform minors which are representable over some field.

Theorem 6.34.

Let MM be a matroid without large uniform minors and FMF_{M} its foundation. Then the following assertions are equivalent:

  1. (1)

    MM is representable over some field.

  2. (2)

    FMF1FrF_{M}\simeq F_{1}\otimes\dotsb\otimes F_{r} for r0r\geqslant 0 and either F1,,Fr{𝕌,𝔻,,𝔽3}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{D}},{\mathbb{H}},{\mathbb{F}}_{3}\} or F1,,Fr{𝕌,,𝔽2}F_{1},\dotsc,F_{r}\in\{{\mathbb{U}},{\mathbb{H}},{\mathbb{F}}_{2}\}.

  3. (3)

    MM belongs to one of C1C_{1}C8C_{8} or C10C_{10}.

  4. (4)

    MM is ternary or quaternary.

  5. (5)

    There is no morphism from 𝔽2𝔻{\mathbb{F}}_{2}\otimes{\mathbb{D}} to FMF_{M}.

Proof.

The equivalences (1)\Leftrightarrow(2)\Leftrightarrow(3) follow from Theorem 6.20. The implications (2)\Rightarrow(4)\Rightarrow(5)\Rightarrow(2) can be derived by combining the implications (2)\Rightarrow(1)\Rightarrow(7)\Rightarrow(2) from Theorem 6.28 and (2)\Rightarrow(1)\Rightarrow(6)\Rightarrow(2) from Theorem 6.29. ∎

References

  • [1] Laura Anderson. Vectors of matroids over tracts. J. Combin. Theory Ser. A, 161:236–270, 2019.
  • [2] Federico Ardila, Felipe Rincón, and Lauren Williams. Positively oriented matroids are realizable. J. Eur. Math. Soc. (JEMS), 19(3):815–833, 2017.
  • [3] Matthew Baker and Nathan Bowler. Matroids over partial hyperstructures. Adv. Math., 343:821–863, 2019.
  • [4] Matthew Baker and Tong Jin. On the structure of hyperfields obtained as quotients of fields. To appear in Proc. Amer. Math. Soc., arXiv:1912.11496, 2019.
  • [5] Matthew Baker and Oliver Lorscheid. The moduli space of matroids. Preprint, arXiv:1809.03542, 2018.
  • [6] Robert E. Bixby. On Reid’s characterization of the ternary matroids. J. Combin. Theory Ser. B, 26(2):174–204, 1979.
  • [7] Robert G. Bland and David L. Jensen. Weakly oriented matroids. Cornell University School of OR/IE Technical Report No. 732, 1987.
  • [8] Robert G. Bland and Michel Las Vergnas. Orientability of matroids. J. Combinatorial Theory Ser. B, 24(1):94–123, 1978.
  • [9] Nathan Bowler and Rudi Pendavingh. Perfect matroids over hyperfields. Preprint, arXiv:1908.03420, 2019.
  • [10] Nick Brettell and Rudi Pendavingh. Computing excluded minors for classes of matroids representable over partial fields. In preparation.
  • [11] Tom Brylawski and Dean Lucas. Uniquely representable combinatorial geometries. In Proceedings of the International Colloquium on Combinatorial Theory. Rome, 1973.
  • [12] Ilda P.F. da Silva. Quelques propriétés des matroides orientés. PhD thesis, Université Paris VI, 1987.
  • [13] Andreas W. M. Dress and Walter Wenzel. Geometric algebra for combinatorial geometries. Adv. Math., 77(1):1–36, 1989.
  • [14] Andreas W. M. Dress and Walter Wenzel. On combinatorial and projective geometry. Geom. Dedicata, 34(2):161–197, 1990.
  • [15] Andreas W. M. Dress and Walter Wenzel. Perfect matroids. Adv. Math., 91(2):158–208, 1992.
  • [16] Israel M. Gelfand, Grigori L. Rybnikov, and David A. Stone. Projective orientations of matroids. Adv. Math., 113(1):118–150, 1995.
  • [17] Jon Lee and Matt Scobee. A characterization of the orientations of ternary matroids. J. Combin. Theory Ser. B, 77(2):263–291, 1999.
  • [18] Oliver Lorscheid. Scheme theoretic tropicalization. Preprint, arXiv:1508.07949v2, 2015.
  • [19] Ch. G. Massouros. Methods of constructing hyperfields. Internat. J. Math. Math. Sci., 8(4):725–728, 1985.
  • [20] Peter Nelson. Almost all matroids are nonrepresentable. Bull. Lond. Math. Soc., 50(2):245–248, 2018.
  • [21] James G. Oxley. Matroid theory. Oxford Science Publications. The Clarendon Press, Oxford University Press, New York, 1992.
  • [22] Rudi Pendavingh. Field extensions, derivations, and matroids over skew hyperfields. Preprint, arXiv:1802.02447, 2018.
  • [23] Rudi A. Pendavingh and Stefan H. M. van Zwam. Confinement of matroid representations to subsets of partial fields. J. Combin. Theory Ser. B, 100(6):510–545, 2010.
  • [24] Rudi A. Pendavingh and Stefan H. M. van Zwam. Lifts of matroid representations over partial fields. J. Combin. Theory Ser. B, 100(1):36–67, 2010.
  • [25] Rudi A. Pendavingh and Stefan H. M. van Zwam. Representing some non-representable matroids. Preprint, arXiv:1106.3088, 2011.
  • [26] Charles Semple. kk-regular matroids. In Combinatorics, complexity, & logic (Auckland, 1996), Springer Ser. Discrete Math. Theor. Comput. Sci., pages 376–386. Springer, Singapore, 1997.
  • [27] Charles Semple and Geoff Whittle. On representable matroids having neither U2,5U_{2,5}- nor U3,5U_{3,5}-minors. In Matroid theory (Seattle, WA, 1995), volume 197 of Contemp. Math., pages 377–386. Amer. Math. Soc., Providence, RI, 1996.
  • [28] Charles Semple and Geoff Whittle. Partial fields and matroid representation. Adv. in Appl. Math., 17(2):184–208, 1996.
  • [29] P. D. Seymour. Matroid representation over GF(3){\rm GF}(3). J. Combin. Theory Ser. B, 26(2):159–173, 1979.
  • [30] David Speyer and Lauren K. Williams. The positive Dressian equals the positive tropical Grassmannian. Preprint, arXiv:2003.10231, 2020.
  • [31] William T. Tutte. A homotopy theorem for matroids, I. Trans. Amer. Math. Soc., 88:144–160, 1958.
  • [32] William T. Tutte. A homotopy theorem for matroids, II. Trans. Amer. Math. Soc., 88:161–174, 1958.
  • [33] William T. Tutte. Lectures on matroids. J. Res. Nat. Bur. Standards Sect. B, 69B:1–47, 1965.
  • [34] Walter Wenzel. Projective equivalence of matroids with coefficients. J. Combin. Theory Ser. A, 57(1):15–45, 1991.
  • [35] Geoff Whittle. A characterisation of the matroids representable over GF(3){\rm GF}(3) and the rationals. J. Combin. Theory Ser. B, 65(2):222–261, 1995.
  • [36] Geoff Whittle. On matroids representable over GF(3){\rm GF}(3) and other fields. Trans. Amer. Math. Soc., 349(2):579–603, 1997.
  • [37] Geoff Whittle. Recent work in matroid representation theory. Discrete Math., 302(1-3):285–296, 2005.