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

A semi-small decomposition of the Chow ring of a matroid

Tom Braden Department of Mathematics and Statistics, University of Massachusetts, Amherst, MA. [email protected] June Huh School of Mathematics, Institute for Advanced Study, Princeton, NJ. [email protected] Jacob P. Matherne Department of Mathematics, University of Oregon, Eugene, OR, and Max-Planck-Institut für Mathematik, Bonn, Germany. [email protected] Nicholas Proudfoot Department of Mathematics, University of Oregon, Eugene, OR. [email protected]  and  Botong Wang Department of Mathematics, University of Wisconsin-Madison, Madison, WI. [email protected]
Abstract.

We give a semi-small orthogonal decomposition of the Chow ring of a matroid M\mathrm{M}. The decomposition is used to give simple proofs of Poincaré duality, the hard Lefschetz theorem, and the Hodge–Riemann relations for the Chow ring, recovering the main result of [AHK]. We also show that a similar semi-small orthogonal decomposition holds for the augmented Chow ring of M\mathrm{M}.

June Huh received support from NSF Grant DMS-1638352 and the Ellentuck Fund. Jacob Matherne received support from NSF Grant DMS-1638352, the Association of Members of the Institute for Advanced Study, and the Max Planck Institute for Mathematics in Bonn. Nicholas Proudfoot received support from NSF Grant DMS-1565036. Botong Wang received support from NSF Grant DMS-1701305 and the Alfred P. Sloan foundation.

1. Introduction

A matroid M\mathrm{M} on a finite set EE is a nonempty collection of subsets of EE, called flats of M\mathrm{M}, that satisfies the following properties:

  1. (1)

    The intersection of any two flats is a flat.

  2. (2)

    For any flat FF, any element in EFE\setminus F is contained in exactly one flat that is minimal among the flats strictly containing FF.

Throughout, we suppose in addition that M\mathrm{M} is a loopless matroid:

  1. (3)

    The empty subset of EE is a flat.

We write (M)\mathscr{L}(\mathrm{M}) for the lattice of all flats of M\mathrm{M}. Every maximal flag of proper flats of M\mathrm{M} has the same cardinality dd, called the rank of M\mathrm{M}. A matroid can be equivalently defined in terms of its independent sets or the rank function. For background in matroid theory, we refer to [Oxley] and [Welsh].

The first aim of the present paper is to decompose the Chow ring of M\mathrm{M} as a module over the Chow ring of the deletion Mi\mathrm{M}\setminus i (Theorem 1.1). The decomposition resembles the decomposition of the cohomology ring of a projective variety induced by a semi-small map. In Section 4, we use the decomposition to give simple proofs of Poincaré duality, the hard Lefschetz theorem, and the Hodge–Riemann relations for the Chow ring, recovering the main result of [AHK].

The second aim of the present paper is to introduce the augmented Chow ring of M\mathrm{M}, which contains the graded Möbius algebra of M\mathrm{M} as a subalgebra. We give an analogous semi-small decomposition of the augmented Chow ring of M\mathrm{M} as a module over the augmented Chow ring of the deletion Mi\mathrm{M}\setminus i (Theorem 1.2), and use this to prove Poincaré duality, the hard Lefschetz theorem, and the Hodge–Riemann relations for the augmented Chow ring. These results will play a major role in the forthcoming paper [BHMPW], where we will prove the Top-Heavy conjecture along with the nonnegativity of the coefficients of the Kazhdan–Lusztig polynomial of a matroid.

1.1.

Let S¯M\underline{S}_{\mathrm{M}} be the ring of polynomials with variables labeled by the nonempty proper flats of M\mathrm{M}:

S¯M:-[xF|F is a nonempty proper flat of M].\underline{S}_{\mathrm{M}}\coloneq\mathbb{Q}[x_{F}\hskip 0.85358pt|\hskip 0.85358pt\text{$F$ is a nonempty proper flat of $\mathrm{M}$}].

The Chow ring of M\mathrm{M}, introduced by Feichtner and Yuzvinsky in [FY], is the quotient algebra111A slightly different presentation for the Chow ring of M\mathrm{M} was used in [FY] in a more general context. The present description was used in [AHK], where the Chow ring of M\mathrm{M} was denoted A(M)A(\mathrm{M}). For a comparison of the two presentations, see [BES].

CH¯(M):-S¯M/(I¯M+J¯M),\underline{\mathrm{CH}}(\mathrm{M})\coloneq\underline{S}_{\mathrm{M}}/(\underline{I}_{\mathrm{M}}+\underline{J}_{\mathrm{M}}),

where I¯M\underline{I}_{\mathrm{M}} is the ideal generated by the linear forms

i1FxFi2FxF,for every pair of distinct elements i1 and i2 of E,\sum_{i_{1}\in F}x_{F}-\sum_{i_{2}\in F}x_{F},\ \ \text{for every pair of distinct elements $i_{1}$ and $i_{2}$ of $E$},

and J¯M\underline{J}_{\mathrm{M}} is the ideal generated by the quadratic monomials

xF1xF2,for every pair of incomparable nonempty proper flats F1 and F2 of M.x_{F_{1}}x_{F_{2}},\ \ \text{for every pair of incomparable nonempty proper flats $F_{1}$ and $F_{2}$ of $\mathrm{M}$.}

When EE is nonempty, the Chow ring of M\mathrm{M} admits a degree map

deg¯M:CH¯d1(M),x:-FxF1,\underline{\deg}_{\mathrm{M}}:\underline{\mathrm{CH}}^{d-1}(\mathrm{M})\longrightarrow\mathbb{Q},\qquad x_{\mathscr{F}}\coloneq\prod_{F\in\mathscr{F}}x_{F}\longmapsto 1,

where \mathscr{F} is any complete flag of nonempty proper flats of M\mathrm{M} (Definition 2.12). For any integer kk, the degree map defines the Poincaré pairing

CH¯k(M)×CH¯dk1(M),(η1,η2)deg¯M(η1η2).\underline{\mathrm{CH}}^{k}(\mathrm{M})\times\underline{\mathrm{CH}}^{d-k-1}(\mathrm{M})\longrightarrow\mathbb{Q},\quad(\eta_{1},\eta_{2})\longmapsto\underline{\deg}_{\mathrm{M}}(\eta_{1}\eta_{2}).

If M\mathrm{M} is realizable over a field,222 We say that M\mathrm{M} is realizable over a field 𝔽\mathbb{F} if there exists a linear subspace V𝔽EV\subseteq\mathbb{F}^{E} such that SES\subseteq E is independent if and only if the projection from VV to 𝔽S\mathbb{F}^{S} is surjective. Almost all matroids are not realizable over any field [Nelson]. then the Chow ring of M\mathrm{M} is isomorphic to the Chow ring of a smooth projective variety over the field (Remark 2.13).

Let ii be an element of EE, and let Mi\mathrm{M}\setminus i be the deletion of ii from M\mathrm{M}. By definition, Mi\mathrm{M}\setminus i is the matroid on EiE\setminus i whose flats are the sets of the form FiF\setminus i for a flat FF of M\mathrm{M}. The Chow rings of M\mathrm{M} and Mi\mathrm{M}\setminus i are related by the graded algebra homomorphism

θ¯i=θ¯iM:CH¯(Mi)CH¯(M),xFxF+xFi,\underline{\theta}_{i}=\underline{\theta}^{\mathrm{M}}_{i}:\underline{\mathrm{CH}}(\mathrm{M}\setminus i)\longrightarrow\underline{\mathrm{CH}}(\mathrm{M}),\qquad x_{F}\longmapsto x_{F}+x_{F\cup i},

where a variable in the target is set to zero if its label is not a flat of M\mathrm{M}. Let CH¯(i)\underline{\mathrm{CH}}_{(i)} be the image of the homomorphism θ¯i\underline{\theta}_{i}, and let 𝒮¯i\underline{\mathscr{S}}_{i} be the collection

𝒮¯i=𝒮¯i(M)={F|F is a nonempty proper subset of Ei such that F(M) and Fi(M)}.\underline{\mathscr{S}}_{i}=\underline{\mathscr{S}}_{i}(\mathrm{M})=\big{\{}F\hskip 0.85358pt|\hskip 0.85358pt\text{$F$ is a nonempty proper subset of $E\setminus i$ such that $F\in\mathscr{L}(\mathrm{M})$ and $F\cup i\in\mathscr{L}(\mathrm{M})$}\big{\}}.

The element ii is said to be a coloop of M\mathrm{M} if the ranks of M\mathrm{M} and Mi\mathrm{M}\setminus i are not equal.

Theorem 1.1.

If ii is not a coloop of M\mathrm{M}, there is a direct sum decomposition of CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}) into indecomposable graded CH¯(Mi)\underline{\mathrm{CH}}(\mathrm{M}\setminus i)-modules

CH¯(M)=CH¯(i)F𝒮¯ixFiCH¯(i).\underline{\mathrm{CH}}(\mathrm{M})=\underline{\mathrm{CH}}_{(i)}\oplus\bigoplus_{F\in\underline{\mathscr{S}}_{i}}x_{F\cup i}\underline{\mathrm{CH}}_{(i)}. (D¯1\underline{\mathrm{D}}_{1})

All pairs of distinct summands are orthogonal for the Poincaré pairing of CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}). If ii is a coloop of M\mathrm{M}, there is a direct sum decomposition of CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}) into indecomposable graded CH¯(Mi)\underline{\mathrm{CH}}(\mathrm{M}\setminus i)-modules333When E={i}E=\{i\}, we treat the symbol xx_{\varnothing} as zero in the right-hand side of (D¯2\underline{\mathrm{D}}_{2}).

CH¯(M)=CH¯(i)xEiCH¯(i)F𝒮¯ixFiCH¯(i).\underline{\mathrm{CH}}(\mathrm{M})=\underline{\mathrm{CH}}_{(i)}\oplus x_{E\setminus i}\underline{\mathrm{CH}}_{(i)}\oplus\bigoplus_{F\in\underline{\mathscr{S}}_{i}}x_{F\cup i}\underline{\mathrm{CH}}_{(i)}. (D¯2\underline{\mathrm{D}}_{2})

All pairs of distinct summands except for the first two are orthogonal for the Poincaré pairing of CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}).

We write rkM:2E\text{rk}_{\mathrm{M}}:2^{E}\to\mathbb{N} for the rank function of M\mathrm{M}. For any proper flat FF of M\mathrm{M}, we set444The symbols MF\mathrm{M}^{F} and MF\mathrm{M}_{F} appear inconsistently in the literature, sometimes this way and sometimes interchanged. The localization is frequently called the restriction. On the other hand, the contraction is also sometimes called the restriction, especially in the context of hyperplane arrangements, so we avoid the word restriction to minimize ambiguity.

MF\displaystyle\mathrm{M}^{F} :-the localization of M at F, a loopless matroid on F of rank equal to rkM(F),\displaystyle\coloneq\text{the localization of $\mathrm{M}$ at $F$, a loopless matroid on $F$ of rank equal to $\text{rk}_{\mathrm{M}}(F)$},
MF\displaystyle\mathrm{M}_{F} :-the contraction of M by F, a loopless matroid on EF of rank equal to drkM(F).\displaystyle\coloneq\text{the contraction of $\mathrm{M}$ by $F$, a loopless matroid on $E\setminus F$ of rank equal to $d-\text{rk}_{\mathrm{M}}(F)$}.

The lattice of flats of MF\mathrm{M}^{F} can be identified with the lattice of flats of M\mathrm{M} that are contained in FF, and the lattice of flats of MF\mathrm{M}_{F} can be identified with the lattice of flats of M\mathrm{M} that contain FF. The CH¯(Mi)\underline{\operatorname{CH}}(\mathrm{M}\setminus i)-module summands in the decompositions (D¯1\underline{\mathrm{D}}_{1}) and (D¯2\underline{\mathrm{D}}_{2}) admit isomorphisms

CH¯(i)CH¯(Mi)andxFiCH¯(i)CH¯(MFi)CH¯(MF)[1],\underline{\mathrm{CH}}_{(i)}\cong\underline{\operatorname{CH}}(\mathrm{M}\setminus i)\ \ \text{and}\ \ x_{F\cup i}\underline{\mathrm{CH}}_{(i)}\cong\underline{\mathrm{CH}}(\mathrm{M}_{F\cup i})\otimes\underline{\mathrm{CH}}(\mathrm{M}^{F})[-1],

where [1][-1] indicates a degree shift (Propositions 3.4 and 3.5). In addition, if ii is a coloop of M\mathrm{M},

xEiCH¯(i)CH¯(Mi)[1].x_{E\setminus i}\underline{\mathrm{CH}}_{(i)}\cong\underline{\operatorname{CH}}(\mathrm{M}\setminus i)[-1].

Numerically, the semi-smallness of the decomposition (D¯1\underline{\mathrm{D}}_{1}) is reflected in the identity

dimxFiCH¯(i)k1=dimxFiCH¯(i)dk2for F𝒮¯i.\operatorname{dim}x_{F\cup i}\underline{\mathrm{CH}}^{k-1}_{(i)}=\operatorname{dim}x_{F\cup i}\underline{\mathrm{CH}}_{(i)}^{d-k-2}\ \ \text{for $F\in\underline{\mathscr{S}}_{i}$}.

When M\mathrm{M} is the Boolean matroid on EE, the graded dimension of CH¯(M)\underline{\operatorname{CH}}(\mathrm{M}) is given by the Eulerian numbers dk\genfrac{<}{>}{0.0pt}{}{d}{k}, and the decomposition (D¯2\underline{\mathrm{D}}_{2}) specializes to the known quadratic recurrence relation

sd(t)=sd1(t)+tk=0d2(d1k)sk(t)sdk1(t),s0(t)=1,s_{d}(t)=s_{d-1}(t)+t\sum_{k=0}^{d-2}{d-1\choose k}s_{k}(t)s_{d-k-1}(t),\qquad s_{0}(t)=1,

where sk(t)s_{k}(t) is the kk-th Eulerian polynomial [Petersen, Theorem 1.5].

1.2.

We also give similar decompositions for the augmented Chow ring of M\mathrm{M}, which we now introduce. Let SMS_{\mathrm{M}} be the ring of polynomials in two sets of variables

SM:-[yi|i is an element of E][xF|F is a proper flat of M].S_{\mathrm{M}}\coloneq\mathbb{Q}[y_{i}\hskip 0.85358pt|\hskip 0.85358pt\text{$i$ is an element of $E$}]\;\otimes\;\mathbb{Q}[x_{F}\hskip 0.85358pt|\hskip 0.85358pt\text{$F$ is a proper flat of $\mathrm{M}$}].

The augmented Chow ring of M\mathrm{M} is the quotient algebra

CH(M):-SM/(IM+JM),\mathrm{CH}(\mathrm{M})\coloneq S_{\mathrm{M}}/(I_{\mathrm{M}}+J_{\mathrm{M}}),

where IMI_{\mathrm{M}} is the ideal generated by the linear forms

yiiFxF,for every element i of E,y_{i}-\sum_{i\notin F}x_{F},\ \ \text{for every element $i$ of $E$},

and JMJ_{\mathrm{M}} is the ideal generated by the quadratic monomials

xF1xF2,\displaystyle x_{F_{1}}x_{F_{2}},\ \ for every pair of incomparable proper flats F1F_{1} and F2F_{2} of M\mathrm{M}, and
yixF,\displaystyle y_{i}\hskip 1.42262ptx_{F},\ \ for every element ii of EE and every proper flat FF of M\mathrm{M} not containing ii.

The augmented Chow ring of M\mathrm{M} admits a degree map

degM:CHd(M),x:-FxF1,\deg_{\mathrm{M}}:\mathrm{CH}^{d}(\mathrm{M})\longrightarrow\mathbb{Q},\qquad x_{\mathscr{F}}\coloneq\prod_{F\in\mathscr{F}}x_{F}\longmapsto 1,

where \mathscr{F} is any complete flag of proper flats of M\mathrm{M} (Definition 2.12). For any integer kk, the degree map defines the Poincaré pairing

CHk(M)×CHdk(M),(η1,η2)degM(η1η2).\mathrm{CH}^{k}(\mathrm{M})\times\mathrm{CH}^{d-k}(\mathrm{M})\longrightarrow\mathbb{Q},\quad(\eta_{1},\eta_{2})\longmapsto\deg_{\mathrm{M}}(\eta_{1}\eta_{2}).

If M\mathrm{M} is realizable over a field, then the augmented Chow ring of M\mathrm{M} is isomorphic to the Chow ring of a smooth projective variety over the field (Remark 2.13). The augmented Chow ring contains the graded Möbius algebra H(M)\mathrm{H}(\mathrm{M}) (Proposition 2.15), and it is related to the Chow ring of M\mathrm{M} by the isomorphism

CH¯(M)CH(M)H(M).\underline{\mathrm{CH}}(\mathrm{M})\cong\mathrm{CH}(\mathrm{M})\otimes_{\mathrm{H}(\mathrm{M})}\mathbb{Q}.

The H(M)\mathrm{H}(\mathrm{M})-module structure of CH(M)\mathrm{CH}(\mathrm{M}) will be studied in detail in the forthcoming paper [BHMPW].

As before, we write Mi\mathrm{M}\setminus i for the matroid obtained from M\mathrm{M} by deleting the element ii. The augmented Chow rings of M\mathrm{M} and Mi\mathrm{M}\setminus i are related by the graded algebra homomorphism

θi=θiM:CH(Mi)CH(M),xFxF+xFi,\theta_{i}=\theta^{\mathrm{M}}_{i}:\mathrm{CH}(\mathrm{M}\setminus i)\longrightarrow\mathrm{CH}(\mathrm{M}),\qquad x_{F}\longmapsto x_{F}+x_{F\cup i},

where a variable in the target is set to zero if its label is not a flat of M\mathrm{M}. Let CH(i)\mathrm{CH}_{(i)} be the image of the homomorphism θi\theta_{i}, and let 𝒮i\mathscr{S}_{i} be the collection

𝒮i=𝒮i(M):-{F|F is a proper subset of Ei such that F(M) and Fi(M)}.\mathscr{S}_{i}=\mathscr{S}_{i}(\mathrm{M})\coloneq\big{\{}F\hskip 0.85358pt|\hskip 0.85358pt\text{$F$ is a proper subset of $E\setminus i$ such that $F\in\mathscr{L}(\mathrm{M})$ and $F\cup i\in\mathscr{L}(\mathrm{M})$}\big{\}}.
Theorem 1.2.

If ii is not a coloop of M\mathrm{M}, there is a direct sum decomposition of CH(M)\mathrm{CH}(\mathrm{M}) into indecomposable graded CH(Mi)\mathrm{CH}(\mathrm{M}\setminus i)-modules

CH(M)=CH(i)F𝒮ixFiCH(i).\mathrm{CH}(\mathrm{M})=\mathrm{CH}_{(i)}\oplus\bigoplus_{F\in\mathscr{S}_{i}}x_{F\cup i}\mathrm{CH}_{(i)}. (D1\mathrm{D}_{1})

All pairs of distinct summands are orthogonal for the Poincaré pairing of CH(M)\mathrm{CH}(\mathrm{M}). If ii is a coloop of M\mathrm{M}, there is a direct sum decomposition of CH(M)\mathrm{CH}(\mathrm{M}) into indecomposable graded CH(Mi)\mathrm{CH}(\mathrm{M}\setminus i)-modules

CH(M)=CH(i)xEiCH(i)F𝒮ixFiCH(i).\mathrm{CH}(\mathrm{M})=\mathrm{CH}_{(i)}\oplus x_{E\setminus i}\mathrm{CH}_{(i)}\oplus\bigoplus_{F\in\mathscr{S}_{i}}x_{F\cup i}\mathrm{CH}_{(i)}. (D2\mathrm{D}_{2})

All pairs of distinct summands except for the first two are orthogonal for the Poincaré pairing of CH(M)\mathrm{CH}(\mathrm{M}).

The CH(Mi)\operatorname{CH}(\mathrm{M}\setminus i)-module summands in the decompositions (D1\mathrm{D}_{1}) and (D2\mathrm{D}_{2}) admit isomorphisms

CH(i)CH(Mi)andxFiCH(i)CH¯(MFi)CH(MF)[1],\mathrm{CH}_{(i)}\cong\operatorname{CH}(\mathrm{M}\setminus i)\ \ \text{and}\ \ x_{F\cup i}\mathrm{CH}_{(i)}\cong\underline{\mathrm{CH}}(\mathrm{M}_{F\cup i})\otimes\mathrm{CH}(\mathrm{M}^{F})[-1],

where [1][-1] indicates a degree shift (Propositions 3.4 and 3.5). In addition, if ii is a coloop of M\mathrm{M},

xEiCH(i)CH(Mi)[1].x_{E\setminus i}\mathrm{CH}_{(i)}\cong\operatorname{CH}(\mathrm{M}\setminus i)[-1].

Numerically, the semi-smallness of the decomposition (D1\mathrm{D}_{1}) is reflected in the identity

dimxFiCH(i)k1=dimxFiCH(i)dk1for F𝒮i.\operatorname{dim}x_{F\cup i}\mathrm{CH}^{k-1}_{(i)}=\operatorname{dim}x_{F\cup i}\mathrm{CH}_{(i)}^{d-k-1}\ \ \text{for $F\in\mathscr{S}_{i}$}.

1.3.

Let B\mathrm{B} be the Boolean matroid on EE. By definition, every subset of EE is a flat of B\mathrm{B}. The Chow rings of B\mathrm{B} and M\mathrm{M} are related by the surjective graded algebra homomorphism

CH¯(B)CH¯(M),xSxS,\underline{\mathrm{CH}}(\mathrm{B})\longrightarrow\underline{\mathrm{CH}}(\mathrm{M}),\qquad x_{S}\longmapsto x_{S},

where a variable in the target is set to zero if its label is not a flat of M\mathrm{M}. Similarly, we have a surjective graded algebra homomorphism

CH(B)CH(M),xSxS,\mathrm{CH}(\mathrm{B})\longrightarrow\mathrm{CH}(\mathrm{M}),\qquad x_{S}\longmapsto x_{S},

where a variable in the target is set to zero if its label is not a flat of M\mathrm{M}. As in [AHK, Section 4], we may identify the Chow ring CH¯(B)\underline{\mathrm{CH}}(\mathrm{B}) with the ring of piecewise polynomial functions modulo linear functions on the normal fan Π¯B\underline{\Pi}_{\mathrm{B}} of the standard permutohedron in E\mathbb{R}^{E}. Similarly, the augmented Chow ring CH(B)\mathrm{CH}(\mathrm{B}) can be identified with the ring of piecewise polynomial functions modulo linear functions of the normal fan ΠB\Pi_{\mathrm{B}} of the stellahedron in E\mathbb{R}^{E} (Definition 2.4). A convex piecewise linear function on a complete fan is said to be strictly convex if there is a bijection between the cones in the fan and the faces of the graph of the function.

In Section 4, we use Theorems 1.1 and 1.2 to give simple proofs of Poincaré duality, the hard Lefschetz theorem, and the Hodge–Riemann relations for CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}) and CH(M)\mathrm{CH}(\mathrm{M}).

Theorem 1.3.

Let ¯\underline{\ell} be a strictly convex piecewise linear function on Π¯B\underline{\Pi}_{\mathrm{B}}, viewed as an element of CH¯1(M)\underline{\mathrm{CH}}^{1}(\mathrm{M}).

  1. (1)

    (Poincaré duality theorem) For every nonnegative integer k<d2k<\frac{d}{2}, the bilinear pairing

    CH¯k(M)×CH¯dk1(M),(η1,η2)deg¯M(η1η2)\underline{\mathrm{CH}}^{k}(\mathrm{M})\times\underline{\mathrm{CH}}^{d-k-1}(\mathrm{M})\longrightarrow\mathbb{Q},\quad(\eta_{1},\eta_{2})\longmapsto\underline{\deg}_{\mathrm{M}}(\eta_{1}\eta_{2})

    is non-degenerate.

  2. (2)

    (Hard Lefschetz theorem) For every nonnegative integer k<d2k<\frac{d}{2}, the multiplication map

    CH¯k(M)CH¯dk1(M),η¯d2k1η\underline{\mathrm{CH}}^{k}(\mathrm{M})\longrightarrow\underline{\mathrm{CH}}^{d-k-1}(\mathrm{M}),\quad\eta\longmapsto\underline{\ell}^{d-2k-1}\eta

    is an isomorphism.

  3. (3)

    (Hodge–Riemann relations) For every nonnegative integer k<d2k<\frac{d}{2}, the bilinear form

    CH¯k(M)×CH¯k(M),(η1,η2)(1)kdeg¯M(¯d2k1η1η2)\underline{\mathrm{CH}}^{k}(\mathrm{M})\times\underline{\mathrm{CH}}^{k}(\mathrm{M})\longrightarrow\mathbb{Q},\quad(\eta_{1},\eta_{2})\longmapsto(-1)^{k}\underline{\deg}_{\mathrm{M}}(\underline{\ell}^{d-2k-1}\eta_{1}\eta_{2})

    is positive definite on the kernel of multiplication by ¯d2k\underline{\ell}^{d-2k}.

Let \ell be a strictly convex piecewise linear function on ΠB\Pi_{\mathrm{B}}, viewed as an element of CH1(M)\mathrm{CH}^{1}(\mathrm{M}).

  1. (4)

    (Poincaré duality theorem) For every nonnegative integer kd2k\leq\frac{d}{2}, the bilinear pairing

    CHk(M)×CHdk(M),(η1,η2)degM(η1η2)\mathrm{CH}^{k}(\mathrm{M})\times\mathrm{CH}^{d-k}(\mathrm{M})\longrightarrow\mathbb{Q},\quad(\eta_{1},\eta_{2})\longmapsto\deg_{\mathrm{M}}(\eta_{1}\eta_{2})

    is non-degenerate.

  2. (5)

    (Hard Lefschetz theorem) For every nonnegative integer kd2k\leq\frac{d}{2}, the multiplication map

    CHk(M)CHdk(M),ηd2kη\mathrm{CH}^{k}(\mathrm{M})\longrightarrow\mathrm{CH}^{d-k}(\mathrm{M}),\quad\eta\longmapsto\ell^{d-2k}\eta

    is an isomorphism.

  3. (6)

    (Hodge–Riemann relations) For every nonnegative integer kd2k\leq\frac{d}{2}, the bilinear form

    CHk(M)×CHk(M),(η1,η2)(1)kdegM(d2kη1η2)\mathrm{CH}^{k}(\mathrm{M})\times\mathrm{CH}^{k}(\mathrm{M})\longrightarrow\mathbb{Q},\quad(\eta_{1},\eta_{2})\longmapsto(-1)^{k}\deg_{\mathrm{M}}(\ell^{d-2k}\eta_{1}\eta_{2})

    is positive definite on the kernel of multiplication by d2k+1\ell^{d-2k+1}.

Theorem 1.3 holds non-vacuously, as there are strictly convex piecewise linear functions on ΠB\Pi_{\mathrm{B}} and Π¯B\underline{\Pi}_{\mathrm{B}} (Proposition 2.6). The first part of Theorem 1.3 on CH¯(M)\underline{\operatorname{CH}}(\mathrm{M}) recovers the main result of [AHK].555Independent proofs of Poincaré duality for CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}) were given in [BES] and [BDF]. The authors of [BES] also prove the degree 11 Hodge–Riemann relations for CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}). The second part of Theorem 1.3 on CH(M)\operatorname{CH}(\mathrm{M}) is new.

1.4.

In Section 5, we use Theorems 1.1 and 1.2 to obtain decompositions of CH¯(M)\underline{\operatorname{CH}}(\mathrm{M}) and CH(M)\operatorname{CH}(\mathrm{M}) related to those appearing in [AHK, Theorem 6.18]. Let H¯α¯(M)\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M}) be the subalgebra of CH¯(M)\underline{\operatorname{CH}}(\mathrm{M}) generated by the element

α¯M:-iGxGCH¯1(M),\underline{\alpha}_{\mathrm{M}}\coloneq\sum_{i\in G}x_{G}\in\underline{\mathrm{CH}}^{1}(\mathrm{M}),

where the sum is over all nonempty proper flats GG of M\mathrm{M} containing a given element ii in EE, and let Hα(M)\operatorname{H}_{\alpha}(\mathrm{M}) be the subalgebra of CH(M)\operatorname{CH}(\mathrm{M}) generated by the element

αM:-GxGCH1(M),\alpha_{\mathrm{M}}\coloneq\sum_{G}x_{G}\in\mathrm{CH}^{1}(\mathrm{M}),

where the sum is over all proper flats GG of M\mathrm{M}. We define graded subspaces J¯α¯(M)\underline{\mathrm{J}}_{\underline{\alpha}}(\mathrm{M}) and Jα(M)\mathrm{J}_{\alpha}(\mathrm{M}) by

J¯α¯k(M):-{H¯α¯k(M)if kd1,0if k=d1,Jαk(M):-{Hαk(M)if kd,0if k=d.\underline{\mathrm{J}}_{\underline{\alpha}}^{k}(\mathrm{M})\coloneq\begin{cases}\underline{\mathrm{H}}_{\underline{\alpha}}^{k}(\mathrm{M})&\text{if $k\neq d-1$,}\\ \hfil 0&\text{if $k=d-1$,}\end{cases}\qquad\mathrm{J}_{\alpha}^{k}(\mathrm{M})\coloneq\begin{cases}\operatorname{H}_{\alpha}^{k}(\mathrm{M})&\text{if $k\neq d$,}\\ \hfil 0&\text{if $k=d$.}\end{cases}

A degree computation shows that the elements α¯Md1\underline{\alpha}_{\mathrm{M}}^{d-1} and αMd\alpha_{\mathrm{M}}^{d} are nonzero (Proposition 2.26).

Theorem 1.4.

Let 𝒞=𝒞(M)\mathscr{C}=\mathscr{C}(\mathrm{M}) be the set of all nonempty proper flats of M\mathrm{M}, and let 𝒞¯=𝒞¯(M)\underline{\mathscr{C}}=\underline{\mathscr{C}}(\mathrm{M}) be the set of all proper flats of M\mathrm{M} with rank at least two.

  1. (1)

    We have a decomposition of H¯α¯(M)\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M})-modules

    CH¯(M)=H¯α¯(M)F𝒞¯ψ¯MFCH¯(MF)J¯α¯(MF).\underline{\operatorname{CH}}(\mathrm{M})=\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M})\oplus\ \bigoplus_{F\in\underline{\mathscr{C}}}\ \underline{\psi}^{F}_{\mathrm{M}}\ \underline{\operatorname{CH}}(\mathrm{M}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F}). (D¯3\underline{\mathrm{D}}_{3})

    All pairs of distinct summands are orthogonal for the Poincaré pairing of CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}).

  2. (2)

    We have a decomposition of Hα(M)\mathrm{H}_{\alpha}(\mathrm{M})-modules

    CH(M)=Hα(M)F𝒞ψMFCH¯(MF)Jα(MF).\operatorname{CH}(\mathrm{M})=\operatorname{H}_{\alpha}(\mathrm{M})\oplus\bigoplus_{F\in\mathscr{C}}\psi^{F}_{\mathrm{M}}\ \underline{\operatorname{CH}}(\mathrm{M}_{F})\otimes{\mathrm{J}}_{\alpha}(\mathrm{M}^{F}). (D3\mathrm{D}_{3})

    All pairs of distinct summands are orthogonal for the Poincaré pairing of CH(M)\mathrm{CH}(\mathrm{M}).

Here ψ¯MF\underline{\psi}^{F}_{\mathrm{M}} is the injective CH¯(M)\underline{\operatorname{CH}}(\mathrm{M})-module homomorphism (Propositions 2.21 and 2.22)

ψ¯MF:CH¯(MF)CH¯(MF)CH¯(M),FxFFF′′xF′′xFFxFF′′xF′′,\underline{\psi}^{F}_{\mathrm{M}}:\underline{\operatorname{CH}}(\mathrm{M}_{F})\otimes\underline{\operatorname{CH}}(\mathrm{M}^{F})\longrightarrow\underline{\operatorname{CH}}(\mathrm{M}),\quad\prod_{F^{\prime}}x_{F^{\prime}\setminus F}\otimes\prod_{F^{\prime\prime}}x_{F^{\prime\prime}}\longmapsto x_{F}\prod_{F^{\prime}}x_{F^{\prime}}\prod_{F^{\prime\prime}}x_{F^{\prime\prime}},

and ψMF\psi^{F}_{\mathrm{M}} is the injective CH(M)\operatorname{CH}(\mathrm{M})-module homomorphism (Propositions 2.18 and 2.19)

ψMF:CH¯(MF)CH(MF)CH(M)FxFFF′′xF′′xFFxFF′′xF′′.\psi^{F}_{\mathrm{M}}:\underline{\operatorname{CH}}(\mathrm{M}_{F})\otimes\operatorname{CH}(\mathrm{M}^{F})\longrightarrow\operatorname{CH}(\mathrm{M})\quad\prod_{F^{\prime}}x_{F^{\prime}\setminus F}\otimes\prod_{F^{\prime\prime}}x_{F^{\prime\prime}}\longmapsto x_{F}\prod_{F^{\prime}}x_{F^{\prime}}\prod_{F^{\prime\prime}}x_{F^{\prime\prime}}.

When M\mathrm{M} is the Boolean matroid on EE, the decomposition (D¯3\underline{\mathrm{D}}_{3}) specializes to a linear recurrence relation for the Eulerian polynomials

0=1+k=0d(dk)ttdk1tsk(t),s0(t)=1.0=1+\sum_{k=0}^{d}{d\choose k}\frac{t-t^{d-k}}{1-t}s_{k}(t),\qquad s_{0}(t)=1.

When applied repeatedly, Theorem 1.4 produces bases of CH¯(M)\underline{\operatorname{CH}}(\mathrm{M}) and CH(M)\operatorname{CH}(\mathrm{M}) that are permuted by the automorphism group of M\mathrm{M}.666Different bases of CH¯(M)\underline{\operatorname{CH}}(\mathrm{M}) are given in [FY, Corollary 1] and [BES, Corollary 3.3.3].

Acknowledgments. We thank Christopher Eur and Matthew Stevens for useful discussions.

2. The Chow ring and the augmented Chow ring of a matroid

In this section, we collect the various properties of the algebras CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}) and CH(M)\mathrm{CH}(\mathrm{M}) that we will need in order to prove Theorems 1.11.4. In Section 2.1, we review the definition and basic properties of the Bergman fan and introduce the closely related augmented Bergman fan of a matroid. Section 2.2 is devoted to understanding the stars of the various rays in these two fans, while Section 2.3 is where we compute the space of balanced top-dimensional weights on each fan. Feichtner and Yuzvinsky showed that the Chow ring of a matroid coincides with the Chow ring of the toric variety associated with its Bergman fan [FY, Theorem 3], and we establish the analogous result for the augmented Chow ring in Section 2.4. Section 2.5 is where we show that the augmented Chow ring contains the graded Möbius algebra. In Section 2.6, we use the results of Section 2.2 to construct various homomorphisms that relate the Chow and augmented Chow rings of different matroids.

Remark 2.1.

It is worth noting why we need to interpret CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}) and CH(M)\mathrm{CH}(\mathrm{M}) as Chow rings of toric varieties. First, the study of balanced weights on the Bergman fan and augmented Bergman fan allow us to show that CH¯d1(M)\underline{\mathrm{CH}}^{d-1}(\mathrm{M}) and CHd(M)\mathrm{CH}^{d}(\mathrm{M}) are nonzero, which is not easy to prove directly from the definitions. The definition of the pullback and pushforward maps in Section 2.6 is made cleaner by thinking about fans, though it would also be possible to define these maps by taking Propositions 2.17, 2.18, 2.20, 2.21, 2.23, and 2.24 as definitions. Finally, and most importantly, the fan perspective will be essential for understanding the ample classes that appear in Theorem 1.3.

2.1. Fans

Let EE be a finite set, and let M\mathrm{M} be a loopless matroid of rank dd on the ground set EE. We write rkM\text{rk}_{\mathrm{M}} for the rank function of M\mathrm{M}, and write clM\text{cl}_{\mathrm{M}} for the closure operator of M\mathrm{M}, which for a set SS returns the smallest flat containing SS. The independence complex IM\mathrm{I}_{\mathrm{M}} of M\mathrm{M} is the simplicial complex of independent sets of M\mathrm{M}. A set IEI\subseteq E is independent if and only if the rank of clM(I)\text{cl}_{\mathrm{M}}(I) is |I||I|. The vertices of IM\mathrm{I}_{\mathrm{M}} are the elements of the ground set EE, and a collection of vertices is a face of IM\mathrm{I}_{\mathrm{M}} when the corresponding set of elements is an independent set of M\mathrm{M}. The Bergman complex Δ¯M\underline{\Delta}_{\mathrm{M}} of M\mathrm{M} is the order complex of the poset of nonempty proper flats of M\mathrm{M}. The vertices of Δ¯M\underline{\Delta}_{\mathrm{M}} are the nonempty proper flats of M\mathrm{M}, and a collection of vertices is a face of Δ¯M\underline{\Delta}_{\mathrm{M}} when the corresponding set of flats is a flag. The independence complex of M\mathrm{M} is pure of dimension d1d-1, and the Bergman complex of M\mathrm{M} is pure of dimension d2d-2. For a detailed study of the simplicial complexes IM\mathrm{I}_{\mathrm{M}} and Δ¯M\underline{\Delta}_{\mathrm{M}}, we refer to [Bjorner]. We introduce the augmented Bergman complex ΔM\Delta_{\mathrm{M}} of M\mathrm{M} as a simplicial complex that interpolates between the independence complex and the Bergman complex of M\mathrm{M}.

Definition 2.2.

Let II be an independent set of M\mathrm{M}, and let \mathscr{F} be a flag of proper flats of M\mathrm{M}. When II is contained in every flat in \mathscr{F}, we say that II is compatible with \mathscr{F} and write II\leq\mathscr{F}. The augmented Bergman complex ΔM\Delta_{\mathrm{M}} of M\mathrm{M} is the simplicial complex of all compatible pairs II\leq\mathscr{F}, where II is an independent set of M\mathrm{M} and \mathscr{F} is a flag of proper flats of M\mathrm{M}.

A vertex of the augmented Bergman complex ΔM\Delta_{\mathrm{M}} is either a singleton subset of EE or a proper flat of M\mathrm{M}. More precisely, the vertices of ΔM\Delta_{\mathrm{M}} are the compatible pairs either of the form {i}\{i\}\leq\varnothing or of the form {F}\varnothing\leq\{F\}, where ii is an element of EE and FF is a proper flat of M\mathrm{M}. The augmented Bergman complex contains both the independence complex IM\mathrm{I}_{\mathrm{M}} and the Bergman complex Δ¯M\underline{\Delta}_{\mathrm{M}} as subcomplexes. In fact, ΔM\Delta_{\mathrm{M}} contains the order complex of the poset of proper flats of M\mathrm{M}, which is the cone over the Bergman complex with the cone point corresponding to the empty flat. It is straightforward to check that ΔM\Delta_{\mathrm{M}} is pure of dimension d1d-1.

Proposition 2.3.

The Bergman complex and the augmented Bergman complex of M\mathrm{M} are both connected in codimension 11.

Proof.

The statement about the Bergman complex is a direct consequence of its shellability [Bjorner]. We prove the statement about the augmented Bergman complex using the statement about the Bergman complex.

The claim is that, given any two facets of ΔM\Delta_{\mathrm{M}}, one may travel from one facet to the other by passing through faces of codimension at most 11. Since the Bergman complex of M\mathrm{M} is connected in codimension 11, the subcomplex of ΔM\Delta_{\mathrm{M}} consisting of faces of the form \varnothing\leq\mathscr{F} is connected in codimension 11. Thus it suffices to show that any facet of ΔM\Delta_{\mathrm{M}} can be connected to a facet of the form \varnothing\leq\mathscr{F} through codimension 11 faces.

Let II\leq\mathscr{F} be a facet of ΔM\Delta_{\mathrm{M}}. If II is nonempty, choose any element ii of II, and consider the flag of flats 𝒢\mathscr{G} obtained by adjoining the closure of IiI\setminus i to \mathscr{F}. The independent set IiI\setminus i is compatible with the flag 𝒢\mathscr{G}, and the facet II\leq\mathscr{F} is adjacent to the facet Ii𝒢I\setminus i\leq\mathscr{G}. Repeating the procedure, we can connect the given facet to a facet of the desired form through codimension 11 faces. ∎

Let E\mathbb{R}^{E} be the vector space spanned by the standard basis vectors 𝐞i\mathbf{e}_{i} corresponding to the elements iEi\in E. For an arbitrary subset SES\subseteq E, we set

𝐞S:-iS𝐞i.\mathbf{e}_{S}\coloneq\sum_{i\in S}\mathbf{e}_{i}.

For an element iEi\in E, we write ρi\rho_{i} for the ray generated by the vector 𝐞i\mathbf{e}_{i} in E\mathbb{R}^{E}. For a subset SES\subseteq E, we write ρS\rho_{S} for the ray generated by the vector 𝐞ES-\mathbf{e}_{E\setminus S} in E\mathbb{R}^{E}, and write ρ¯S\underline{\rho}_{S} for the ray generated by the vector 𝐞S\mathbf{e}_{S} in E/𝐞S\mathbb{R}^{E}/\langle\mathbf{e}_{S}\rangle. Using these rays, we construct fan models of the Bergman complex and the augmented Bergman complex as follows.

Definition 2.4.

The Bergman fan Π¯M\underline{\Pi}_{\mathrm{M}} of M\mathrm{M} is a simplicial fan in the quotient space E/𝐞E\mathbb{R}^{E}/\langle\mathbf{e}_{E}\rangle with rays ρ¯F\underline{\rho}_{F} for nonempty proper flats FF of M\mathrm{M}. The cones of Π¯M\underline{\Pi}_{\mathrm{M}} are of the form

σ¯:-cone{𝐞F}F=cone{𝐞EF}F,\underline{\sigma}_{\mathscr{F}}\coloneq\text{cone}\{\mathbf{e}_{F}\}_{F\in\mathscr{F}}=\text{cone}\{-\mathbf{e}_{E\setminus F}\}_{F\in\mathscr{F}},

where \mathscr{F} is a flag of nonempty proper flats of M\mathrm{M}.

The augmented Bergman fan ΠM\Pi_{\mathrm{M}} of M\mathrm{M} is a simplicial fan in E\mathbb{R}^{E} with rays ρi\rho_{i} for elements ii in EE and ρF\rho_{F} for proper flats FF of M\mathrm{M}. The cones of the augmented Bergman fan are of the form

σI:-cone{𝐞i}iI+cone{𝐞EF}F,\sigma_{I\leq\mathscr{F}}\coloneq\text{cone}\{\mathbf{e}_{i}\}_{i\in I}+\text{cone}\{-\mathbf{e}_{E\setminus F}\}_{F\in\mathscr{F}},

where \mathscr{F} is a flag of proper flats of M\mathrm{M} and II is an independent set of M\mathrm{M} compatible with \mathscr{F}. We write σI\sigma_{I} for the cone σI\sigma_{I\leq\mathscr{F}} when \mathscr{F} is the empty flag of flats of M\mathrm{M}.

Remark 2.5.

If EE is nonempty, then the Bergman fan Π¯M\underline{\Pi}_{\mathrm{M}} is the star of the ray ρ\rho_{\varnothing} in the augmented Bergman fan ΠM\Pi_{\mathrm{M}}. If EE is empty, then Π¯M\underline{\Pi}_{\mathrm{M}} and ΠM\Pi_{\mathrm{M}} both consist of a single 0-dimensional cone.

\varnothing\leq\varnothing{1}\{1\}\leq\varnothing{{2}}\varnothing\leq\{\{2\}\}{2}\{2\}\leq\varnothing{{1}}\varnothing\leq\{\{1\}\}{}\varnothing\leq\{\varnothing\}{1,2}\{1,2\}\leq\varnothing{2}{{2}}\{2\}\leq\{\{2\}\}{,{2}}\varnothing\leq\{\varnothing,\{2\}\}{,{1}}\varnothing\leq\{\varnothing,\{1\}\}{1}{{1}}\{1\}\leq\{\{1\}\}
Figure 1. The augmented Bergman fan of the rank 22 Boolean matroid on {1,2}\{1,2\}.

Let N\mathrm{N} be another loopless matroid on EE. The matroid M\mathrm{M} is said to be a quotient of N\mathrm{N} if every flat of M\mathrm{M} is a flat of N\mathrm{N}. The condition implies that every independent set of M\mathrm{M} is an independent set of N\mathrm{N} [Kung, Proposition 8.1.6]. Therefore, when M\mathrm{M} is a quotient of N\mathrm{N}, the augmented Bergman fan of M\mathrm{M} is a subfan of the augmented Bergman fan of N\mathrm{N}, and the Bergman fan of M\mathrm{M} is a subfan of the Bergman fan of N\mathrm{N}. In particular, we have inclusions of fans ΠMΠB\Pi_{\mathrm{M}}\subseteq\Pi_{\mathrm{B}} and Π¯MΠ¯B\underline{\Pi}_{\mathrm{M}}\subseteq\underline{\Pi}_{\mathrm{B}}, where B\mathrm{B} is the Boolean matroid on EE defined by the condition that EE is an independent set of B\mathrm{B}.

Proposition 2.6.

The Bergman fan and the augmented Bergman fan of B\mathrm{B} are each normal fans of convex polytopes. In particular, there are strictly convex piecewise linear functions on Π¯B\underline{\Pi}_{\mathrm{B}} and ΠB\Pi_{\mathrm{B}}.

The above proposition can be used to show that the augmented Bergman fan and the Bergman fan of M\mathrm{M} are, in fact, fans.

Proof.

The statement for the Bergman fan is well-known: The Bergman fan of B\mathrm{B} is the normal fan of the standard permutohedron in 𝐞EE\mathbf{e}_{E}^{\perp}\subseteq\mathbb{R}^{E}. See, for example, [AHK, Section 2]. The statement for the augmented Bergman fan ΠB\Pi_{\mathrm{B}} follows from the fact that it is an iterated stellar subdivision of the normal fan of the simplex

conv{𝐞i,𝐞E}iEE.\text{conv}\{\mathbf{e}_{i},\mathbf{e}_{E}\}_{i\in E}\subseteq\mathbb{R}^{E}.

More precisely, ΠB\Pi_{\mathrm{B}} is isomorphic to the fan Σ𝒫\Sigma_{\mathscr{P}} in [AHK, Definition 2.3], where 𝒫\mathscr{P} is the order filter of all subsets of E0E\cup 0 containing the new element 0, via the linear isomorphism

EE0/𝐞E+𝐞0,𝐞j𝐞j.\mathbb{R}^{E}\longrightarrow\mathbb{R}^{E\cup 0}/\langle\mathbf{e}_{E}+\mathbf{e}_{0}\rangle,\quad\mathbf{e}_{j}\longmapsto\mathbf{e}_{j}.

It is shown in [AHK, Proposition 2.4] that Σ𝒫\Sigma_{\mathscr{P}} is an iterated stellar subdivision of the normal fan of the simplex.777In fact, the augmented Bergman fan ΠB\Pi_{\mathrm{B}} is the normal fan of the stellahedron in E\mathbb{R}^{E}, the graph associahedron of the star graph with |E||E| endpoints. We refer to [CD] and [Devadoss] for detailed discussions of graph associahedra and their realizations.

A direct inspection shows that ΠM\Pi_{\mathrm{M}} is a unimodular fan; that is, the set of primitive ray generators in any cone in ΠM\Pi_{\mathrm{M}} is a subset of a basis of the free abelian group E\mathbb{Z}^{E}. It follows that Π¯M\underline{\Pi}_{\mathrm{M}} is also a unimodular fan; that is, the set of primitive ray generators in any cone in ΠM\Pi_{\mathrm{M}} is a subset of a basis of the free abelian group E/𝐞E\mathbb{Z}^{E}/\langle\mathbf{e}_{E}\rangle.

2.2. Stars

For any element ii of EE, we write cl(i)\text{cl}(i) for the closure of ii in M\mathrm{M}, and write ιi\iota_{i} for the injective linear map

ιi:Ecl(i)E/𝐞i,𝐞j𝐞j.\iota_{i}:\mathbb{R}^{E\setminus\text{cl}(i)}\longrightarrow\mathbb{R}^{E}/\langle\mathbf{e}_{i}\rangle,\qquad\mathbf{e}_{j}\longmapsto\mathbf{e}_{j}.

For any proper flat FF of M\mathrm{M}, we write ιF\iota_{F} for the linear isomorphism

ιF:EF/𝐞EFFE/𝐞EF,𝐞j𝐞j.\iota_{F}:\mathbb{R}^{E\setminus F}/\langle\mathbf{e}_{E\setminus F}\rangle\ \oplus\ \mathbb{R}^{F}\longrightarrow\mathbb{R}^{E}/\langle\mathbf{e}_{E\setminus F}\rangle,\qquad\mathbf{e}_{j}\longmapsto\mathbf{e}_{j}.

For any nonempty proper flat FF of M\mathrm{M}, we write ι¯F\underline{\iota}_{F} for the linear isomorphism

ι¯F:EF/𝐞EFF/𝐞FE/𝐞E,𝐞EF,𝐞j𝐞j.\underline{\iota}_{F}:\mathbb{R}^{E\setminus F}/\langle\mathbf{e}_{E\setminus F}\rangle\ \oplus\ \mathbb{R}^{F}/\langle\mathbf{e}_{F}\rangle\longrightarrow\mathbb{R}^{E}/\langle\mathbf{e}_{E},\mathbf{e}_{E\setminus F}\rangle,\qquad\mathbf{e}_{j}\longmapsto\mathbf{e}_{j}.

Let MF\mathrm{M}^{F} be the localization of M\mathrm{M} at FF, and let MF\mathrm{M}_{F} be the contraction of M\mathrm{M} by FF.

Proposition 2.7.

The following are descriptions of the stars of the rays in ΠM\Pi_{\mathrm{M}} and Π¯M\underline{\Pi}_{\mathrm{M}} using the three linear maps above.

  1. (1)

    For any element iEi\in E, the linear map ιi\iota_{i} identifies the augmented Bergman fan of Mcl(i)\mathrm{M}_{\text{cl}(i)} with the star of the ray ρi\rho_{i} in the augmented Bergman fan of M\mathrm{M}:

    ΠMcl(i)starρiΠM.\Pi_{\mathrm{M}_{\text{cl}(i)}}\cong\text{star}_{\rho_{i}}\Pi_{\mathrm{M}}.
  2. (2)

    For any proper flat FF of M\mathrm{M}, the linear map ιF\iota_{F} identifies the product of the Bergman fan of MF\mathrm{M}_{F} and the augmented Bergman fan of MF\mathrm{M}^{F} with the star of the ray ρF\rho_{F} in the augmented Bergman fan of M\mathrm{M}:

    Π¯MF×ΠMFstarρFΠM.\underline{\Pi}_{\mathrm{M}_{F}}\times\Pi_{\mathrm{M}^{F}}\cong\text{star}_{\rho_{F}}\Pi_{\mathrm{M}}.
  3. (3)

    For any nonempty proper flat FF of M\mathrm{M}, the linear map ι¯F\underline{\iota}_{F} identifies the product of the Bergman fan of MF\mathrm{M}_{F} and the Bergman fan of MF\mathrm{M}^{F} with the star of the ray ρ¯F\underline{\rho}_{F} in the Bergman fan of M\mathrm{M}:

    Π¯MF×Π¯MFstarρ¯FΠ¯M.\underline{\Pi}_{\mathrm{M}_{F}}\times\underline{\Pi}_{\mathrm{M}^{F}}\cong\text{star}_{\underline{\rho}_{F}}\underline{\Pi}_{\mathrm{M}}.

Repeated applications of the first statement show that, for any independent set II of M\mathrm{M}, the star of the cone σI\sigma_{I} in ΠM\Pi_{\mathrm{M}} can be identified with the augmented Bergman fan of Mcl(I)\mathrm{M}_{\text{cl}(I)}, where cl(I)\text{cl}(I) is the closure of II in M\mathrm{M}.

Proof.

The first statement follows from the following facts: A flat of M\mathrm{M} contains ii if and only if it contains cl(i)\text{cl}(i), and an independent set of M\mathrm{M} containing ii does not contain any other element in cl(i)\text{cl}(i). The second and third statements follow directly from the definitions. ∎

2.3. Weights

For any simplicial fan Σ\Sigma, we write Σk\Sigma_{k} for the set of kk-dimensional cones in Σ\Sigma. If τ\tau is a codimension 11 face of a cone σ\sigma, we write

𝐞σ/τ:-the primitive generator of the unique ray in σ that is not in τ.\mathbf{e}_{\sigma/\tau}\coloneq\text{the primitive generator of the unique ray in $\sigma$ that is not in $\tau$}.

A kk-dimensional balanced weight on Σ\Sigma is a \mathbb{Q}-valued function ω\omega on Σk\Sigma_{k} that satisfies the balancing condition: For every (k1)(k-1)-dimensional cone τ\tau in Σ\Sigma,

τσω(σ)𝐞σ/τ\sum_{\tau\subset\sigma}\omega(\sigma)\mathbf{e}_{\sigma/\tau} is contained in the subspace spanned by τ\tau,

where the sum is over all kk-dimensional cones σ\sigma containing τ\tau. We write MWk(Σ)\mathrm{MW}_{k}(\Sigma) for the group of kk-dimensional balanced weights on Σ\Sigma.

Proposition 2.8.

The Bergman fan and the augmented Bergman fan of M\mathrm{M} have the following unique balancing property.

  1. (1)

    A (d1)(d-1)-dimensional weight on Π¯M\underline{\Pi}_{\mathrm{M}} is balanced if and only if it is constant.

  2. (2)

    A dd-dimensional weight on ΠM\Pi_{\mathrm{M}} is balanced if and only if it is constant.

Proof.

The first statement is [AHK, Proposition 5.2]. We prove the second statement.

Let σI\sigma_{I\leq\mathscr{F}} be a codimension 11 cone of ΠM\Pi_{\mathrm{M}}, and let FF be the smallest flat in {E}\mathscr{F}\cup\{E\}. We analyze the primitive generators of the rays in the star of the cone σI\sigma_{I\leq\mathscr{F}} in ΠM\Pi_{\mathrm{M}}. Let cl(I)\text{cl}(I) be the closure of II in M\mathrm{M}. There are two cases.

When the closure of II is not FF, the primitive ray generators in question are 𝐞Ecl(I)-\mathbf{e}_{E\setminus\text{cl}(I)} and 𝐞i\mathbf{e}_{i}, for elements ii in FF not in the closure of II. The primitive ray generators satisfy the relation

𝐞Ecl(I)+iFcl(I)𝐞i=𝐞EF,-\mathbf{e}_{E\setminus\text{cl}(I)}+\sum_{i\in F\setminus\text{cl}(I)}\mathbf{e}_{i}=-\mathbf{e}_{E\setminus F},

which is zero modulo the span of σI\sigma_{I\leq\mathscr{F}}. As the 𝐞i\mathbf{e}_{i}’s are independent modulo the span of σI\sigma_{I\leq\mathscr{F}}, any relation between the primitive generators must be a multiple of the displayed one.

When the closure of II is FF, the fact that σI\sigma_{I\leq\mathscr{F}} has codimension 11 implies that there is a unique integer kk with rkF<k<rkM\operatorname{rk}F<k<\operatorname{rk}\mathrm{M} such that \mathscr{F} does not include a flat of rank kk. Let FF_{\circ} be the unique flat in \mathscr{F} of rank k1k-1, and let FF^{\circ} be the unique flat in {E}\mathscr{F}\cup\{E\} of rank k+1k+1. The primitive ray generators in question are 𝐞EG-\mathbf{e}_{E\setminus G} for the flats GG in 𝒢\mathscr{G}, where 𝒢\mathscr{G} is the set of flats of M\mathrm{M} covering FF_{\circ} and covered by FF^{\circ}. By the flat partition property of matroids [Oxley, Section 1.4], the primitive ray generators satisfy the relation

G𝒢𝐞EG=(1)𝐞EF𝐞EF,\sum_{G\in\mathscr{G}}-\mathbf{e}_{E\setminus G}=-(\ell-1)\mathbf{e}_{E\setminus F_{\circ}}-\mathbf{e}_{E\setminus F^{\circ}},

which is zero modulo the span of σI\sigma_{I\leq\mathscr{F}}. Since any proper subset of the primitive generators 𝐞EG-\mathbf{e}_{E\setminus G} for GG in 𝒢\mathscr{G} is independent modulo the span of σI\sigma_{I\leq\mathscr{F}}, any relation between the primitive generators must be a multiple of the displayed one.

The local analysis above shows that any constant dd-dimensional weight on ΠM\Pi_{\mathrm{M}} is balanced. Since ΠM\Pi_{\mathrm{M}} is connected in codimension 11 by Proposition 2.3, it also shows that any dd-dimensional balanced weight on ΠM\Pi_{\mathrm{M}} must be constant. ∎

2.4. Chow rings

Any unimodular fan Σ\Sigma in E\mathbb{R}^{E} defines a graded commutative algebra CH(Σ)\mathrm{CH}(\Sigma), which is the Chow ring of the associated smooth toric variety XΣX_{\Sigma} over \mathbb{C} with rational coefficients. Equivalently, CH(Σ)\mathrm{CH}(\Sigma) is the ring of continuous piecewise polynomial functions on Σ\Sigma with rational coefficients modulo the ideal generated by globally linear functions [Brion, Section 3.1]. We write CHk(Σ)\mathrm{CH}^{k}(\Sigma) for the Chow group of codimension kk cycles in XΣX_{\Sigma}, so that

CH(Σ)=kCHk(Σ).\mathrm{CH}(\Sigma)=\bigoplus_{k}\mathrm{CH}^{k}(\Sigma).

The group of kk-dimensional balanced weights on Σ\Sigma is related to CHk(Σ)\mathrm{CH}^{k}(\Sigma) by the isomorphism

MWk(Σ)Hom(CHk(Σ),),ω(xσω(σ)),\mathrm{MW}_{k}(\Sigma)\longrightarrow\mathrm{Hom}_{\mathbb{Q}}(\mathrm{CH}^{k}(\Sigma),\mathbb{Q}),\quad\omega\longmapsto(x_{\sigma}\longmapsto\omega(\sigma)),

where xσx_{\sigma} is the class of the torus orbit closure in XΣX_{\Sigma} corresponding to a kk-dimensional cone σ\sigma in Σ\Sigma. See [AHK, Section 5] for a detailed discussion. For general facts on toric varieties and Chow rings, and for any undefined terms, we refer to [CLS] and [Fulton].

In Proposition 2.10 below, we show that the Chow ring of M\mathrm{M} coincides with CH(Π¯M)\operatorname{CH}(\underline{\Pi}_{\mathrm{M}}) and that the augmented Chow ring of M\mathrm{M} coincides with CH(ΠM)\operatorname{CH}(\Pi_{\mathrm{M}}).

Lemma 2.9.

The following identities hold in the augmented Chow ring CH(M)\mathrm{CH}(\mathrm{M}).

  1. (1)

    For any element ii of EE, we have yi2=0y_{i}^{2}=0.

  2. (2)

    For any two bases I1I_{1} and I2I_{2} of a flat FF of M\mathrm{M}, we have iI1yi=iI2yi\prod_{i\in I_{1}}y_{i}=\prod_{i\in I_{2}}y_{i}.

  3. (3)

    For any dependent set JJ of M\mathrm{M}, we have jJyj=0\prod_{j\in J}y_{j}=0.

Proof.

The first identity is a straightforward consequence of the relations in IMI_{\mathrm{M}} and JMJ_{\mathrm{M}}:

yi2=yi(iFxF)=0.y_{i}^{2}=y_{i}\Big{(}\sum_{i\notin F}x_{F}\Big{)}=0.

For the second identity, we may assume that I1I2={i1}I_{1}\setminus I_{2}=\{i_{1}\} and I2I1={i2}I_{2}\setminus I_{1}=\{i_{2}\}, by the basis exchange property of matroids. Since a flat of M\mathrm{M} contains I1I_{1} if and only if it contains I2I_{2}, we have

(i1GxG)iI1I2yi=(I1GxG)iI1I2yi=(I2GxG)iI1I2yi=(i2GxG)iI1I2yi.\Big{(}\sum_{i_{1}\in G}x_{G}\Big{)}\prod_{i\in I_{1}\cap I_{2}}y_{i}=\Big{(}\sum_{I_{1}\subseteq G}x_{G}\Big{)}\prod_{i\in I_{1}\cap I_{2}}y_{i}=\Big{(}\sum_{I_{2}\subseteq G}x_{G}\Big{)}\prod_{i\in I_{1}\cap I_{2}}y_{i}=\Big{(}\sum_{i_{2}\in G}x_{G}\Big{)}\prod_{i\in I_{1}\cap I_{2}}y_{i}.

This immediately implies that we also have

(i1GxG)iI1I2yi=(i2GxG)iI1I2yi,\Big{(}\sum_{i_{1}\notin G}x_{G}\Big{)}\prod_{i\in I_{1}\cap I_{2}}y_{i}=\Big{(}\sum_{i_{2}\notin G}x_{G}\Big{)}\prod_{i\in I_{1}\cap I_{2}}y_{i},

which tells us that

iI1yi=yi1iI1I2yi=(i1GxG)iI1I2yi=(i2GxG)iI1I2yi=yi2iI1I2yi=iI2yi.\prod_{i\in I_{1}}y_{i}=y_{i_{1}}\prod_{i\in I_{1}\cap I_{2}}y_{i}=\Big{(}\sum_{i_{1}\notin G}x_{G}\Big{)}\prod_{i\in I_{1}\cap I_{2}}y_{i}=\Big{(}\sum_{i_{2}\notin G}x_{G}\Big{)}\prod_{i\in I_{1}\cap I_{2}}y_{i}=y_{i_{2}}\prod_{i\in I_{1}\cap I_{2}}y_{i}=\prod_{i\in I_{2}}y_{i}.

For the third identity, we may suppose that JJ is a circuit, that is, a minimal dependent set. Since M\mathrm{M} is a loopless matroid, we may choose distinct elements j1j_{1} and j2j_{2} from JJ. Note that the independent sets Jj1J\setminus j_{1} and Jj2J\setminus j_{2} have the same closure because JJ is a circuit. Therefore, by the second identity, we have

jJj1yj=jJj2yj.\prod_{j\in J\setminus j_{1}}y_{j}=\prod_{j\in J\setminus j_{2}}y_{j}.

Combining the above with the first identity, we get

jJyj=yj1jJj1yj=yj1jJj2yj=yj12jJ{j1,j2}yj=0.\prod_{j\in J}y_{j}=y_{j_{1}}\prod_{j\in J\setminus j_{1}}y_{j}=y_{j_{1}}\prod_{j\in J\setminus j_{2}}y_{j}=y_{j_{1}}^{2}\prod_{j\in J\setminus\{j_{1},j_{2}\}}y_{j}=0.\qed

By the second identity in Lemma 2.9, we may define

yF:-iIyiin CH(M)y_{F}\coloneq\prod_{i\in I}y_{i}\ \ \text{in $\mathrm{CH}(\mathrm{M})$}

for any flat FF of M\mathrm{M} and any basis II of FF. The element yEy_{E} will play the role of the fundamental class for the augmented Chow ring of M\mathrm{M}.

Proposition 2.10.

We have isomorphisms

CH¯(M)CH(Π¯M)CH(M)CH(ΠM).\underline{\mathrm{CH}}(\mathrm{M})\cong\mathrm{CH}(\underline{\Pi}_{\mathrm{M}})\mathrm{CH}(\mathrm{M})\cong\mathrm{CH}(\Pi_{\mathrm{M}}).
Proof.

The first isomorphism is proved in [FY, Theorem 3]; see also [AHK, Section 5.3].

Let KMK_{\mathrm{M}} be the ideal of SMS_{\mathrm{M}} generated by the monomials jJyj\prod_{j\in J}y_{j} for every dependent set JJ of M\mathrm{M}. The ring of continuous piecewise polynomial functions on ΠM\Pi_{\mathrm{M}} is isomorphic to the Stanley–Reisner ring of ΔM\Delta_{\mathrm{M}}, which is equal to

SM/(JM+KM).S_{\mathrm{M}}/(J_{\mathrm{M}}+K_{\mathrm{M}}).

The ring CH(ΠM)\mathrm{CH}(\Pi_{\mathrm{M}}) is obtained from this ring by killing the linear forms that generate the ideal IMI_{\mathrm{M}}. In other words, we have a surjective homomorphism

CH(M):-SM/(IM+JM)SM/(IM+JM+KM)CH(ΠM).\mathrm{CH}(\mathrm{M})\coloneq S_{\mathrm{M}}/(I_{\mathrm{M}}+J_{\mathrm{M}})\longrightarrow S_{\mathrm{M}}/(I_{\mathrm{M}}+J_{\mathrm{M}}+K_{\mathrm{M}})\cong\mathrm{CH}(\Pi_{\mathrm{M}}).

The fact that this is an isomorphism follows from the third part of Lemma 2.9. ∎

Remark 2.11.

By Proposition 2.10, the graded dimension of the Chow ring of the rank dd Boolean matroid CH¯(B)\underline{\operatorname{CH}}(\mathrm{B}) is given by the hh-vector of the permutohedron in E\mathbb{R}^{E}. In other words, we have

dimCH¯k(B)=the Eulerian number dk.\operatorname{dim}\underline{\operatorname{CH}}^{k}(\mathrm{B})=\text{the Eulerian number $\genfrac{<}{>}{0.0pt}{}{d}{k}$}.

See [Petersen, Section 9.1] for more on permutohedra and Eulerian numbers.

If EE is nonempty, we have the balanced weight

1MWd1(Π¯M)Hom(CH¯d1(M),),1\in\mathrm{MW}_{d-1}(\underline{\Pi}_{\mathrm{M}})\cong\text{Hom}_{\mathbb{Q}}(\underline{\mathrm{CH}}^{d-1}(\mathrm{M}),\mathbb{Q}),

which can be used to define a degree map on the Chow ring of M\mathrm{M}. Similarly, for any EE,

1MWd(ΠM)Hom(CHd(M),)1\in\mathrm{MW}_{d}(\Pi_{\mathrm{M}})\cong\text{Hom}_{\mathbb{Q}}(\mathrm{CH}^{d}(\mathrm{M}),\mathbb{Q})

can be used to define a degree map on the augmented Chow ring of M\mathrm{M}.

Definition 2.12.

Consider the following degree maps for the Chow ring and the augmented Chow ring of M\mathrm{M}.

  1. (1)

    If EE is nonempty, the degree map for CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}) is the linear map

    deg¯M:CH¯d1(M),x1,\underline{\deg}_{\mathrm{M}}:\underline{\mathrm{CH}}^{d-1}(\mathrm{M})\longrightarrow\mathbb{Q},\quad x_{\mathscr{F}}\longmapsto 1,

    where xx_{\mathscr{F}} is any monomial corresponding to a maximal cone σ¯\underline{\sigma}_{\mathscr{F}} of Π¯M\underline{\Pi}_{\mathrm{M}}.

  2. (2)

    For any EE, the degree map for CH(M)\mathrm{CH}(\mathrm{M}) is the linear map

    degM:CHd(M),xI1,\deg_{\mathrm{M}}:\mathrm{CH}^{d}(\mathrm{M})\longrightarrow\mathbb{Q},\quad x_{I\leq\mathscr{F}}\longmapsto 1,

    where xIx_{I\leq\mathscr{F}} is any monomial corresponding to a maximal cone σI\sigma_{I\leq\mathscr{F}} of ΠM\Pi_{\mathrm{M}}.

By Proposition 2.8, the degree maps are well-defined and are isomorphisms. It follows that, for any two maximal cones σ¯1\underline{\sigma}_{\mathscr{F}_{1}} and σ¯2\underline{\sigma}_{\mathscr{F}_{2}} of the Bergman fan of M\mathrm{M},

x1=x2in CH¯d1(M).x_{\mathscr{F}_{1}}=x_{\mathscr{F}_{2}}\ \ \text{in $\underline{\mathrm{CH}}^{d-1}(\mathrm{M})$}.

Similarly, for any two maximal cones σI11\sigma_{I_{1}\leq\mathscr{F}_{1}} and σI22\sigma_{I_{2}\leq\mathscr{F}_{2}} of the augmented Bergman fan of M\mathrm{M},

yF1x1=yF2x2in CHd(M),y_{F_{1}}x_{\mathscr{F}_{1}}=y_{F_{2}}x_{\mathscr{F}_{2}}\ \ \text{in $\mathrm{CH}^{d}(\mathrm{M})$},

where F1F_{1} is the closure of I1I_{1} in M\mathrm{M} and F2F_{2} is the closure of I2I_{2} in M\mathrm{M}. Proposition 2.10 shows that

CH¯k(M)=0for kdandCHk(M)=0for k>d.\underline{\mathrm{CH}}^{k}(\mathrm{M})=0\ \ \text{for $k\geq d$}\ \ \text{and}\ \ \mathrm{CH}^{k}(\mathrm{M})=0\ \ \text{for $k>d$.}
Remark 2.13.

Let 𝔽\mathbb{F} be a field, and let VV be a dd-dimensional linear subspace of 𝔽E\mathbb{F}^{E}. We suppose that the subspace VV is not contained in 𝔽S𝔽E\mathbb{F}^{S}\subseteq\mathbb{F}^{E} for any proper subset SS of EE. Let B\mathrm{B} be the Boolean matroid on EE, and let M\mathrm{M} be the loopless matroid on EE defined by

S is an independent set of Mthe restriction to V of the projection 𝔽E𝔽S is surjective.\text{$S$ is an independent set of $\mathrm{M}$}\Longleftrightarrow\text{the restriction to $V$ of the projection $\mathbb{F}^{E}\to\mathbb{F}^{S}$ is surjective}.

Let (𝔽E)\mathbb{P}(\mathbb{F}^{E}) be the projective space of lines in 𝔽E\mathbb{F}^{E}, and let 𝕋¯E\underline{\mathbb{T}}_{E} be its open torus. For any proper flat FF of M\mathrm{M}, we write HFH_{F} for the projective subspace

HF:-{p(V)|pi=0 for all iF}.H_{F}\coloneq\big{\{}p\in\mathbb{P}(V)\hskip 0.85358pt|\hskip 0.85358pt\text{$p_{i}=0$ for all $i\in F$}\big{\}}.

The wonderful variety X¯V\underline{X}_{V} is obtained from (V)\mathbb{P}(V) by first blowing up HFH_{F} for every corank 11 flat FF, then blowing up the strict transforms of HFH_{F} for every corank 22 flat FF, and so on. Equivalently,

X¯V\displaystyle\underline{X}_{V} =the closure of (V)𝕋¯E in the toric variety X¯M defined by Π¯M\displaystyle=\text{the closure of $\mathbb{P}(V)\cap\underline{\mathbb{T}}_{E}$ in the toric variety $\underline{X}_{\mathrm{M}}$ defined by $\underline{\Pi}_{\mathrm{M}}$}
=the closure of (V)𝕋¯E in the toric variety X¯B defined by Π¯B.\displaystyle=\text{the closure of $\mathbb{P}(V)\cap\underline{\mathbb{T}}_{E}$ in the toric variety $\underline{X}_{\mathrm{B}}$ defined by $\underline{\Pi}_{\mathrm{B}}$}.

When EE is nonempty, the inclusion X¯VX¯M\underline{X}_{V}\subseteq\underline{X}_{\mathrm{M}} induces an isomorphism between their Chow rings,888In general, the inclusion X¯VX¯M\underline{X}_{V}\subseteq\underline{X}_{\mathrm{M}} does not induce an isomorphism between their singular cohomology rings. and hence the Chow ring of X¯V\underline{X}_{V} is isomorphic to CH¯(M)\underline{\operatorname{CH}}(\mathrm{M}) [FY, Corollary 2].

Let (𝔽E𝔽)\mathbb{P}(\mathbb{F}^{E}\oplus\mathbb{F}) be the projective completion of 𝔽E\mathbb{F}^{E}, and let 𝕋E\mathbb{T}_{E} be its open torus. The projective completion (V𝔽)\mathbb{P}(V\oplus\mathbb{F}) contains a copy of (V)\mathbb{P}(V) as the hyperplane at infinity, and it therefore contains a copy of HFH_{F} for every nonempty proper flat FF. The augmented wonderful variety XVX_{V} is obtained from (V𝔽1)\mathbb{P}(V\oplus\mathbb{F}^{1}) by first blowing up HFH_{F} for every corank 11 flat FF, then blowing up the strict transforms of HFH_{F} for every corank 22 flat FF, and so on. Equivalently,

XV\displaystyle X_{V} =the closure of (V𝔽)𝕋E in the toric variety XM defined by ΠM\displaystyle=\text{the closure of $\mathbb{P}(V\oplus\mathbb{F})\cap\mathbb{T}_{E}$ in the toric variety $X_{\mathrm{M}}$ defined by $\Pi_{\mathrm{M}}$}
=the closure of (V𝔽)𝕋E in the toric variety XB defined by ΠB.\displaystyle=\text{the closure of $\mathbb{P}(V\oplus\mathbb{F})\cap\mathbb{T}_{E}$ in the toric variety $X_{\mathrm{B}}$ defined by $\Pi_{\mathrm{B}}$}.

The inclusion XVXMX_{V}\subseteq X_{\mathrm{M}} induces an isomorphism between their Chow rings, and hence the Chow ring of XVX_{V} is isomorphic to CH(M)\operatorname{CH}(\mathrm{M}).999This can be proved using the interpretation of CH(M)\mathrm{CH}(\mathrm{M}) in the last sentence of Remark 4.1.

2.5. The graded Möbius algebra

For any nonnegative integer kk, we define a vector space

Hk(M):-Fk(M)yF,\mathrm{H}^{k}(\mathrm{M})\coloneq\bigoplus_{F\in\mathscr{L}^{k}(\mathrm{M})}\mathbb{Q}\hskip 1.42262pty_{F},

where the direct sum is over the set k(M)\mathscr{L}^{k}(\mathrm{M}) of rank kk flats of M\mathrm{M}.

Definition 2.14.

The graded Möbius algebra of M\mathrm{M} is the graded vector space

H(M):-k0Hk(M).\mathrm{H}(\mathrm{M})\coloneq\bigoplus_{k\geq 0}\mathrm{H}^{k}(\mathrm{M}).

The multiplication in H(M)\mathrm{H}(\mathrm{M}) is defined by the rule

yF1yF2={yF1F2if rkM(F1)+rkM(F2)=rkM(F1F2),0if rkM(F1)+rkM(F2)>rkM(F1F2),y_{F_{1}}y_{F_{2}}=\begin{cases}y_{F_{1}\lor F_{2}}&\text{if $\text{rk}_{\mathrm{M}}(F_{1})+\text{rk}_{\mathrm{M}}(F_{2})=\text{rk}_{\mathrm{M}}(F_{1}\lor F_{2})$,}\\ \hfil 0&\text{if $\text{rk}_{\mathrm{M}}(F_{1})+\text{rk}_{\mathrm{M}}(F_{2})>\text{rk}_{\mathrm{M}}(F_{1}\lor F_{2})$,}\end{cases}

where \lor stands for the join operation in the lattice of flats (M)\mathscr{L}(\mathrm{M}) of M\mathrm{M}.

Our double use of the symbol yFy_{F} is justified by the following proposition.

Proposition 2.15.

The graded linear map

H(M)CH(M),yFyF\mathrm{H}(\mathrm{M})\longrightarrow\mathrm{CH}(\mathrm{M}),\quad y_{F}\longmapsto y_{F}

is an injective homomorphism of graded algebras.

Proof.

We first show that the linear map is injective. It is enough to check that the subset

{yF}Fk(M)CHk(M)\{y_{F}\}_{F\in\mathscr{L}^{k}(\mathrm{M})}\subseteq\mathrm{CH}^{k}(\mathrm{M})

is linearly independent for every nonnegative integer k<dk<d. Suppose that

Fk(M)cFyF=0for some cF.\sum_{F\in\mathscr{L}^{k}(\mathrm{M})}c_{F}y_{F}=0\ \ \text{for some $c_{F}\in\mathbb{Q}$}.

For any given rank kk flat GG, we choose a saturated flag of proper flats 𝒢\mathscr{G} whose smallest member is GG and observe that

cGyGx𝒢=(Fk(M)cFyF)x𝒢=0.c_{G}y_{G}x_{\mathscr{G}}=\Big{(}\sum_{F\in\mathscr{L}^{k}(\mathrm{M})}c_{F}y_{F}\Big{)}x_{\mathscr{G}}=0.

Since the degree of yGx𝒢y_{G}x_{\mathscr{G}} is 11, this implies that cGc_{G} must be zero.

We next check that the linear map is an algebra homomorphism using Lemma 2.9. Let I1I_{1} be a basis of a flat F1F_{1}, and let I2I_{2} be a basis of a flat F2F_{2}. If the rank of F1F2F_{1}\lor F_{2} is the sum of the ranks of F1F_{1} and F2F_{2}, then I1I_{1} and I2I_{2} are disjoint and their union is a basis of F1F2F_{1}\lor F_{2}. Therefore, in the augmented Chow ring of M\mathrm{M},

yF1yF2=iI1yiiI2yi=iI1I2yi=yF1F2.y_{F_{1}}y_{F_{2}}=\prod_{i\in I_{1}}y_{i}\prod_{i\in I_{2}}y_{i}=\prod_{i\in I_{1}\cup I_{2}}y_{i}=y_{F_{1}\lor F_{2}}.

If the rank of F1F2F_{1}\lor F_{2} is less than the sum of the ranks of F1F_{1} and F2F_{2}, then either I1I_{1} and I2I_{2} intersect or the union of I1I_{1} and I2I_{2} is dependent in M\mathrm{M}. Therefore, in the augmented Chow ring of M\mathrm{M},

yF1yF2=iI1yiiI2yi=0.y_{F_{1}}y_{F_{2}}=\prod_{i\in I_{1}}y_{i}\prod_{i\in I_{2}}y_{i}=0.\qed
Remark 2.16.

Consider the torus 𝕋E\mathbb{T}_{E}, the toric variety XBX_{\mathrm{B}}, and the augmented wonderful variety XVX_{V} in Remark 2.13. The identity of 𝕋E\mathbb{T}_{E} uniquely extends to a toric map

pB:XB(1)E.\mathrm{p}_{\mathrm{B}}:X_{\mathrm{B}}\longrightarrow(\mathbb{P}^{1})^{E}.

Let pV\mathrm{p}_{V} be the restriction of pB\mathrm{p}_{\mathrm{B}} to the augmented wonderful variety XVX_{V}. If we identify the Chow ring of XVX_{V} with CH(M)\operatorname{CH}(\mathrm{M}) as in Remark 2.13, the image of the pullback pV\mathrm{p}_{V}^{*} is the graded Möbius algebra H(M)CH(M)\mathrm{H}(\mathrm{M})\subseteq\operatorname{CH}(\mathrm{M}).

2.6. Pullback and pushforward maps

Let Σ\Sigma be a unimodular fan, and let σ\sigma be a kk-dimensional cone in Σ\Sigma. The torus orbit closure in the smooth toric variety XΣX_{\Sigma} corresponding to σ\sigma can be identified with the toric variety of the fan starσΣ\text{star}_{\sigma}\Sigma. Its class in the Chow ring of XΣX_{\Sigma} is the monomial xσx_{\sigma}, which is the product of the divisor classes xρx_{\rho} corresponding to the rays ρ\rho in σ\sigma. The inclusion ι\iota of the torus orbit closure in XΣX_{\Sigma} defines the pullback ι\iota^{*} and the pushforward ι\iota_{*} between the Chow rings, whose composition is multiplication by the monomial xσx_{\sigma}:

CH(Σ)\textstyle{\mathrm{CH}(\Sigma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}xσ\scriptstyle{x_{\sigma}}ι\scriptstyle{\iota^{*}}CH(Σ)\textstyle{\mathrm{CH}(\Sigma)}CH(starσΣ)\textstyle{\mathrm{CH}(\text{star}_{\sigma}\Sigma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι\scriptstyle{\iota_{*}}

The pullback ι\iota^{*} is a surjective graded algebra homomorphism, while the pushforward ι\iota_{*} is a degree kk homomorphism of CH(Σ)\mathrm{CH}(\Sigma)-modules.

We give an explicit description of the pullback ι\iota^{*} and the pushforward ι\iota_{*} when Σ\Sigma is the augmented Bergman fan ΠM\Pi_{\mathrm{M}} and σ\sigma is the ray ρF\rho_{F} of a proper flat FF of M\mathrm{M}. Recall from Proposition 2.7 that the star of ρF\rho_{F} admits the decomposition

starρFΠMΠ¯MF×ΠMF.\text{star}_{\rho_{F}}\Pi_{\mathrm{M}}\cong\underline{\Pi}_{\mathrm{M}_{F}}\times\Pi_{\mathrm{M}^{F}}.

Thus we may identify the Chow ring of the star of ρF\rho_{F} with CH¯(MF)CH(MF)\underline{\mathrm{CH}}(\mathrm{M}_{F})\otimes\mathrm{CH}(\mathrm{M}^{F}). We denote the pullback to the tensor product by φMF\varphi_{\mathrm{M}}^{F} and the pushforward from the tensor product by ψMF\psi_{\mathrm{M}}^{F}:

CH(M)\textstyle{\mathrm{CH}(\mathrm{M})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}xF\scriptstyle{x_{F}}φFM\scriptstyle{\varphi^{F}_{\mathrm{M}}}CH(M)\textstyle{\mathrm{CH}(\mathrm{M})}CH¯(MF)CH(MF)\textstyle{\underline{\mathrm{CH}}(\mathrm{M}_{F})\otimes\mathrm{CH}(\mathrm{M}^{F})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψFM\scriptstyle{\psi^{F}_{\mathrm{M}}}

To describe the pullback and the pushforward, we introduce Chow classes αM\alpha_{\mathrm{M}}, α¯M\underline{\alpha}_{\mathrm{M}}, and β¯M\underline{\beta}_{\mathrm{M}}. They are defined as the sums

αM:-GxGCH1(M),\alpha_{\mathrm{M}}\coloneq\sum_{G}x_{G}\in\mathrm{CH}^{1}(\mathrm{M}),

where the sum is over all proper flats GG of M\mathrm{M};

α¯M:-iGxGCH¯1(M),\underline{\alpha}_{\mathrm{M}}\coloneq\sum_{i\in G}x_{G}\in\underline{\mathrm{CH}}^{1}(\mathrm{M}),

where the sum is over all nonempty proper flats GG of M\mathrm{M} containing a given element ii in EE; and

β¯M:-iGxGCH¯1(M),\underline{\beta}_{\mathrm{M}}\coloneq\sum_{i\notin G}x_{G}\in\underline{\mathrm{CH}}^{1}(\mathrm{M}),

where the sum is over all nonempty proper flats GG of M\mathrm{M} not containing a given element ii in EE. The linear relations defining CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}) show that α¯M\underline{\alpha}_{\mathrm{M}} and β¯M\underline{\beta}_{\mathrm{M}} do not depend on the choice of ii.

The following two propositions are straightforward.

Proposition 2.17.

The pullback φFM\varphi^{F}_{\mathrm{M}} is the unique graded algebra homomorphism

CH(M)CH¯(MF)CH(MF)\mathrm{CH}(\mathrm{M})\longrightarrow\underline{\mathrm{CH}}(\mathrm{M}_{F})\otimes\mathrm{CH}(\mathrm{M}^{F})

that satisfies the following properties:

  1. \bullet

    If GG is a proper flat of M\mathrm{M} incomparable to FF, then φFM(xG)=0\varphi^{F}_{\mathrm{M}}(x_{G})=0.

  2. \bullet

    If GG is a proper flat of M\mathrm{M} properly contained in FF, then φFM(xG)=1xG\varphi^{F}_{\mathrm{M}}(x_{G})=1\otimes x_{G}.

  3. \bullet

    If GG is a proper flat of M\mathrm{M} properly containing FF, then φFM(xG)=xGF1\varphi^{F}_{\mathrm{M}}(x_{G})=x_{G\setminus F}\otimes 1.

  4. \bullet

    If ii is an element of FF, then φFM(yi)=1yi\varphi^{F}_{\mathrm{M}}(y_{i})=1\otimes y_{i}.

  5. \bullet

    If ii is an element of EFE\setminus F, then φFM(yi)=0\varphi^{F}_{\mathrm{M}}(y_{i})=0.

The above five properties imply the following additional properties of φFM\varphi^{F}_{\mathrm{M}}:

  1. \bullet

    The equality φFM(xF)=1αMFβ¯MF1\varphi^{F}_{\mathrm{M}}(x_{F})=-1\otimes\alpha_{\mathrm{M}^{F}}-\underline{\beta}_{\mathrm{M}_{F}}\otimes 1 holds.

  2. \bullet

    The equality φFM(αM)=α¯MF1\varphi^{F}_{\mathrm{M}}(\alpha_{\mathrm{M}})=\underline{\alpha}_{\mathrm{M}_{F}}\otimes 1 holds.

Proposition 2.18.

The pushforward ψFM\psi^{F}_{\mathrm{M}} is the unique CH(M)\operatorname{CH}(\mathrm{M})-module homomorphism101010We make ψFM\psi^{F}_{\mathrm{M}} into a CH(M)\mathrm{CH}(\mathrm{M})-module homomorphism via the pullback φFM\varphi^{F}_{\mathrm{M}}.

ψFM:CH¯(MF)CH(MF)CH(M)\psi^{F}_{\mathrm{M}}:\underline{\mathrm{CH}}(\mathrm{M}_{F})\otimes\mathrm{CH}(\mathrm{M}^{F})\longrightarrow\mathrm{CH}(\mathrm{M})

that satisfies, for any collection 𝒮\mathscr{S}^{\prime} of proper flats of M\mathrm{M} strictly containing FF and any collection 𝒮\mathscr{S}^{\prime\prime} of proper flats of M\mathrm{M} strictly contained in FF,

ψFM(F𝒮xFFF𝒮xF)=xFF𝒮xFF𝒮xF.\psi^{F}_{\mathrm{M}}\Bigg{(}\prod_{F^{\prime}\in\mathscr{S}^{\prime}}x_{F^{\prime}\setminus F}\otimes\prod_{F^{\prime\prime}\in\mathscr{S}^{\prime\prime}}x_{F^{\prime\prime}}\Bigg{)}=x_{F}\prod_{F^{\prime}\in\mathscr{S}^{\prime}}x_{F^{\prime}}\prod_{F^{\prime\prime}\in\mathscr{S}^{\prime\prime}}x_{F^{\prime\prime}}.

The composition ψMFφMF\psi_{\mathrm{M}}^{F}\circ\varphi_{\mathrm{M}}^{F} is multiplication by the element xFx_{F}, and the composition φMFψMF\varphi_{\mathrm{M}}^{F}\circ\psi_{\mathrm{M}}^{F} is multiplication by the element φMF(xF)\varphi_{\mathrm{M}}^{F}(x_{F}).

Proposition 2.18 shows that the pushforward ψFM\psi^{F}_{\mathrm{M}} commutes with the degree maps:

deg¯MFdegMF=degMψMF.\underline{\deg}_{\mathrm{M}_{F}}\otimes\deg_{\mathrm{M}^{F}}=\deg_{\mathrm{M}}\circ\ \psi_{\mathrm{M}}^{F}.
Proposition 2.19.

If CH¯(MF)\underline{\mathrm{CH}}(\mathrm{M}_{F}) and CH(MF)\mathrm{CH}(\mathrm{M}^{F}) satisfy the Poincaré duality part of Theorem 1.3, then ψMF\psi_{\mathrm{M}}^{F} is injective.

In other words, assuming Poincaré duality for the Chow rings, the graded CH(M)\mathrm{CH}(\mathrm{M})-module CH¯(MF)CH(MF)[1]\underline{\mathrm{CH}}(\mathrm{M}_{F})\otimes\mathrm{CH}(\mathrm{M}^{F})[-1] is isomorphic to the principal ideal of xFx_{F} in CH(M)\mathrm{CH}(\mathrm{M}).111111For a graded vector space VV, we write V[m]V[m] for the graded vector space whose degree kk piece is equal to Vk+mV^{k+m}. In particular,

CH¯(M)[1]ideal(x)CH(M).\underline{\mathrm{CH}}(\mathrm{M})[-1]\cong\text{ideal}(x_{\varnothing})\subseteq\mathrm{CH}(\mathrm{M}).
Proof.

We will use the symbol degF\deg_{F} to denote the degree function deg¯MFdegMF\underline{\deg}_{\mathrm{M}_{F}}\otimes\deg_{\mathrm{M}^{F}}. For contradiction, suppose that ψFM(η)=0\psi^{F}_{\mathrm{M}}(\eta)=0 for η0\eta\neq 0. By the two Poincaré duality statements in Theorem 1.3, there is an element ν\nu such that degF(νη)=1\deg_{F}(\nu\eta)=1. By surjectivity of the pullback φFM\varphi^{F}_{\mathrm{M}}, there is an element μ\mu such that ν=φFM(μ)\nu=\varphi^{F}_{\mathrm{M}}(\mu). Since ψFM\psi^{F}_{\mathrm{M}} is a CH(M)\mathrm{CH}(\mathrm{M})-module homomorphism that commutes with the degree maps, we have

1=degF(νη)=degM(ψFM(νη))=degM(ψFM(φFM(μ)η))=degM(μψFM(η))=degM(0)=0,1=\deg_{F}(\nu\eta)=\deg_{\mathrm{M}}(\psi^{F}_{\mathrm{M}}(\nu\eta))=\deg_{\mathrm{M}}(\psi^{F}_{\mathrm{M}}(\varphi^{F}_{\mathrm{M}}(\mu)\eta))=\deg_{\mathrm{M}}(\mu\psi^{F}_{\mathrm{M}}(\eta))=\deg_{\mathrm{M}}(0)=0,

which is a contradiction. ∎

We next give an explicit description of the pullback ι\iota^{*} and the pushforward ι\iota_{*} when Σ\Sigma is the Bergman fan Π¯M\underline{\Pi}_{\mathrm{M}} and σ\sigma is the ray ρ¯F\underline{\rho}_{F} of a nonempty proper flat FF of M\mathrm{M}. Recall from Proposition 2.7 that the star of ρ¯F\underline{\rho}_{F} admits the decomposition

starρ¯FΠ¯MΠ¯MF×Π¯MF.\text{star}_{\underline{\rho}_{F}}\underline{\Pi}_{\mathrm{M}}\cong\underline{\Pi}_{\mathrm{M}_{F}}\times\underline{\Pi}_{\mathrm{M}^{F}}.

Thus we may identify the Chow ring of the star of ρ¯F\underline{\rho}_{F} with CH¯(MF)CH¯(MF)\underline{\mathrm{CH}}(\mathrm{M}_{F})\otimes\underline{\mathrm{CH}}(\mathrm{M}^{F}). We denote the pullback to the tensor product by φ¯MF\underline{\varphi}_{\mathrm{M}}^{F} and the pushforward from the tensor product by ψ¯MF\underline{\psi}_{\mathrm{M}}^{F}:

CH¯(M)\textstyle{\underline{\mathrm{CH}}(\mathrm{M})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}xF\scriptstyle{x_{F}}φ¯FM\scriptstyle{\underline{\varphi}^{F}_{\mathrm{M}}}CH¯(M)\textstyle{\underline{\mathrm{CH}}(\mathrm{M})}CH¯(MF)CH¯(MF)\textstyle{\underline{\mathrm{CH}}(\mathrm{M}_{F})\otimes\underline{\mathrm{CH}}(\mathrm{M}^{F})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ¯FM\scriptstyle{\underline{\psi}^{F}_{\mathrm{M}}}

The following analogues of Propositions 2.17 and 2.18 are straightforward.

Proposition 2.20.

The pullback φ¯FM\underline{\varphi}^{F}_{\mathrm{M}} is the unique graded algebra homomorphism

CH¯(M)CH¯(MF)CH¯(MF)\underline{\mathrm{CH}}(\mathrm{M})\longrightarrow\underline{\mathrm{CH}}(\mathrm{M}_{F})\otimes\underline{\mathrm{CH}}(\mathrm{M}^{F})

that satisfies the following properties:

  1. \bullet

    If GG is a nonempty proper flat of M\mathrm{M} incomparable to FF, then φ¯FM(xG)=0\underline{\varphi}^{F}_{\mathrm{M}}(x_{G})=0.

  2. \bullet

    If GG is a nonempty proper flat of M\mathrm{M} properly contained in FF, then φ¯FM(xG)=1xG\underline{\varphi}^{F}_{\mathrm{M}}(x_{G})=1\otimes x_{G}.

  3. \bullet

    If GG is a nonempty proper flat of M\mathrm{M} properly containing FF, then φ¯FM(xG)=xGF1\underline{\varphi}^{F}_{\mathrm{M}}(x_{G})=x_{G\setminus F}\otimes 1.

The above three properties imply the following additional properties of φ¯FM\underline{\varphi}^{F}_{\mathrm{M}}:

  1. \bullet

    The equality φ¯FM(xF)=1α¯MFβ¯MF1\underline{\varphi}^{F}_{\mathrm{M}}(x_{F})=-1\otimes\underline{\alpha}_{\mathrm{M}^{F}}-\underline{\beta}_{\mathrm{M}_{F}}\otimes 1 holds.

  2. \bullet

    The equality φ¯FM(α¯M)=α¯MF1\underline{\varphi}^{F}_{\mathrm{M}}(\underline{\alpha}_{\mathrm{M}})=\underline{\alpha}_{\mathrm{M}_{F}}\otimes 1 holds.

  3. \bullet

    The equality φ¯FM(β¯M)=1β¯MF\underline{\varphi}^{F}_{\mathrm{M}}(\underline{\beta}_{\mathrm{M}})=1\otimes\underline{\beta}_{\mathrm{M}^{F}} holds.

Proposition 2.21.

The pushforward ψ¯FM\underline{\psi}^{F}_{\mathrm{M}} is the unique CH¯(M)\underline{\mathrm{CH}}(\mathrm{M})-module homomorphism

CH¯(MF)CH¯(MF)CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}_{F})\otimes\underline{\mathrm{CH}}(\mathrm{M}^{F})\longrightarrow\underline{\mathrm{CH}}(\mathrm{M})

that satisfies, for any collection 𝒮\mathscr{S}^{\prime} of proper flats of M\mathrm{M} strictly containing FF and any collection 𝒮\mathscr{S}^{\prime\prime} of nonempty proper flats of M\mathrm{M} strictly contained in FF,

ψ¯FM(F𝒮xFFF𝒮xF)=xFF𝒮xFF𝒮xF.\underline{\psi}^{F}_{\mathrm{M}}\Bigg{(}\prod_{F^{\prime}\in\mathscr{S}^{\prime}}x_{F^{\prime}\setminus F}\otimes\prod_{F^{\prime\prime}\in\mathscr{S}^{\prime\prime}}x_{F^{\prime\prime}}\Bigg{)}=x_{F}\prod_{F^{\prime}\in\mathscr{S}^{\prime}}x_{F^{\prime}}\prod_{F^{\prime\prime}\in\mathscr{S}^{\prime\prime}}x_{F^{\prime\prime}}.

The composition ψ¯MFφ¯MF\underline{\psi}_{\mathrm{M}}^{F}\circ\underline{\varphi}_{\mathrm{M}}^{F} is multiplication by the element xFx_{F}, and the composition φ¯MFψ¯MF\underline{\varphi}_{\mathrm{M}}^{F}\circ\underline{\psi}_{\mathrm{M}}^{F} is multiplication by the element φ¯MF(xF)\underline{\varphi}_{\mathrm{M}}^{F}(x_{F}).

Proposition 2.21 shows that the pushforward ψ¯FM\underline{\psi}^{F}_{\mathrm{M}} commutes with the degree maps:

deg¯MFdeg¯MF=deg¯Mψ¯MF.\underline{\deg}_{\mathrm{M}_{F}}\otimes\underline{\deg}_{\mathrm{M}^{F}}=\underline{\deg}_{\mathrm{M}}\circ\ \underline{\psi}_{\mathrm{M}}^{F}.
Proposition 2.22.

If CH¯(MF)\underline{\mathrm{CH}}(\mathrm{M}_{F}) and CH¯(MF)\underline{\mathrm{CH}}(\mathrm{M}^{F}) satisfy the Poincaré duality part of Theorem 1.3, then ψ¯FM\underline{\psi}^{F}_{\mathrm{M}} is injective.

In other words, assuming Poincaré duality for the Chow rings, the graded CH¯(M)\underline{\mathrm{CH}}(\mathrm{M})-module CH¯(MF)CH¯(MF)[1]\underline{\mathrm{CH}}(\mathrm{M}_{F})\otimes\underline{\mathrm{CH}}(\mathrm{M}^{F})[-1] is isomorphic to the principal ideal of xFx_{F} in CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}).

Proof.

The proof is essentially identical to that of Proposition 2.19. ∎

Last, we give an explicit description of the pullback ι\iota^{*} and the pushforward ι\iota_{*} when Σ\Sigma is the augmented Bergman fan ΠM\Pi_{\mathrm{M}} and σ\sigma is the cone σI\sigma_{I} of a independent set II of M\mathrm{M}. By Proposition 2.7, we have

starσIΠMΠMF,\text{star}_{\sigma_{I}}\Pi_{\mathrm{M}}\cong\Pi_{\mathrm{M}_{F}},

where FF is the closure of II in M\mathrm{M}. Thus we may identify the Chow ring of the star of σI\sigma_{I} with CH(MF)\mathrm{CH}(\mathrm{M}_{F}). We denote the corresponding pullback by φMF\varphi^{\mathrm{M}}_{F} and the pushforward by ψMF\psi^{\mathrm{M}}_{F}:

CH(M)\textstyle{\mathrm{CH}(\mathrm{M})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}yF\scriptstyle{y_{F}}φFM\scriptstyle{\varphi_{F}^{\mathrm{M}}}CH(M)\textstyle{\mathrm{CH}(\mathrm{M})}CH(MF)\textstyle{\mathrm{CH}(\mathrm{M}_{F})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψFM\scriptstyle{\psi_{F}^{\mathrm{M}}}

Note that the pullback and the pushforward only depend on FF and not on II.

The following analogues of Propositions 2.17 and 2.18 are straightforward.

Proposition 2.23.

The pullback φFM\varphi_{F}^{\mathrm{M}} is the unique graded algebra homomorphism

CH(M)CH(MF)\mathrm{CH}(\mathrm{M})\longrightarrow\mathrm{CH}(\mathrm{M}_{F})

that satisfies the following properties:

  1. \bullet

    If GG is a proper flat of M\mathrm{M} that contains FF, then φFM(xG)=xGF\varphi_{F}^{\mathrm{M}}(x_{G})=x_{G\setminus F}.

  2. \bullet

    If GG is a proper flat of M\mathrm{M} that does not contain FF, then φFM(xG)=0\varphi_{F}^{\mathrm{M}}(x_{G})=0.

The above two properties imply the following additional properties of φFM\varphi_{F}^{\mathrm{M}}:

  1. \bullet

    If ii is an element of FF, then φFM(yi)=0\varphi_{F}^{\mathrm{M}}(y_{i})=0.

  2. \bullet

    If ii is an element of EFE\setminus F, then φFM(yi)=yi\varphi_{F}^{\mathrm{M}}(y_{i})=y_{i}.

  3. \bullet

    The equality φFM(αM)=αMF\varphi_{F}^{\mathrm{M}}(\alpha_{\mathrm{M}})=\alpha_{\mathrm{M}_{F}} holds.

Proposition 2.24.

The pushforward ψFM\psi_{F}^{\mathrm{M}} is the unique CH(M)\mathrm{CH}(\mathrm{M})-module homomorphism

CH(MF)CH(M)\mathrm{CH}(\mathrm{M}_{F})\longrightarrow\mathrm{CH}(\mathrm{M})

that satisfies, for any collection 𝒮\mathscr{S}^{\prime} of proper flats of M\mathrm{M} containing FF,

ψFM(F𝒮xFF)=yFF𝒮xF.\psi_{F}^{\mathrm{M}}\Bigg{(}\prod_{F^{\prime}\in\mathscr{S}^{\prime}}x_{F^{\prime}\setminus F}\Bigg{)}=y_{F}\prod_{F^{\prime}\in\mathscr{S}^{\prime}}x_{F^{\prime}}.

The composition ψMFφMF\psi^{\mathrm{M}}_{F}\circ\varphi^{\mathrm{M}}_{F} is multiplication by the element yFy_{F}, and the composition φMFψMF\varphi^{\mathrm{M}}_{F}\circ\psi^{\mathrm{M}}_{F} is zero.

Proposition 2.24 shows that the pushforward ψFM\psi_{F}^{\mathrm{M}} commutes with the degree maps:

degMF=degMψFM.\deg_{\mathrm{M}_{F}}=\deg_{\mathrm{M}}\circ\ \psi_{F}^{\mathrm{M}}.
Proposition 2.25.

If CH(MF)\mathrm{CH}(\mathrm{M}_{F}) satisfies the Poincaré duality part of Theorem 1.3, then ψFM\psi_{F}^{\mathrm{M}} is injective.

In other words, assuming Poincaré duality for the Chow rings, the graded CH(M)\mathrm{CH}(\mathrm{M})-module CH(MF)[rkM(F)]\mathrm{CH}(\mathrm{M}_{F})[-\text{rk}_{\mathrm{M}}(F)] is isomorphic to the principal ideal of yFy_{F} in CH(M)\mathrm{CH}(\mathrm{M}).

Proof.

The proof is essentially identical to that of Proposition 2.19. ∎

The basic properties of the pullback and the pushforward maps can be used to describe the fundamental classes of CH¯(M)\underline{\operatorname{CH}}(\mathrm{M}) and CH(M)\operatorname{CH}(\mathrm{M}) in terms of α¯M\underline{\alpha}_{\mathrm{M}} and αM\alpha_{\mathrm{M}}.

Proposition 2.26.

The degree of α¯d1M\underline{\alpha}^{d-1}_{\mathrm{M}} is 11, and the degree of αdM\alpha^{d}_{\mathrm{M}} is 11.

Proof.

We prove the first statement by induction on d1d\geq 1. Note that, for any nonempty proper flat FF of rank kk, we have

xFα¯Mdk=ψ¯FM(φ¯FM(α¯Mdk))=ψ¯FM(α¯MFdk1)=0,x_{F}\hskip 0.85358pt\underline{\alpha}_{\mathrm{M}}^{d-k}=\underline{\psi}^{F}_{\mathrm{M}}\big{(}\underline{\varphi}^{F}_{\mathrm{M}}(\underline{\alpha}_{\mathrm{M}}^{d-k})\big{)}=\underline{\psi}^{F}_{\mathrm{M}}\big{(}\underline{\alpha}_{\mathrm{M_{F}}}^{d-k}\otimes 1\big{)}=0,

since CH¯dk(MF)=0\underline{\mathrm{\operatorname{CH}}}^{d-k}(\mathrm{M}_{F})=0. Therefore, for any proper flat aa of rank 11 and any element ii in aa, we have

α¯Md1=(iFxF)α¯Md2=xaα¯Md2.\underline{\alpha}_{\mathrm{M}}^{d-1}=\Big{(}\sum_{i\in F}x_{F}\Big{)}\underline{\alpha}_{\mathrm{M}}^{d-2}=x_{a}\hskip 0.85358pt\underline{\alpha}_{\mathrm{M}}^{d-2}.

Now, using the induction hypothesis applied to the matroid Ma\mathrm{M}_{a} of rank d1d-1, we get

α¯Md1=xaα¯Md2=ψ¯aM(φ¯aM(α¯Md2))=ψ¯aM(α¯Mad21)=x,\underline{\alpha}_{\mathrm{M}}^{d-1}=x_{a}\hskip 0.85358pt\underline{\alpha}_{\mathrm{M}}^{d-2}=\underline{\psi}^{a}_{\mathrm{M}}\big{(}\underline{\varphi}^{a}_{\mathrm{M}}(\underline{\alpha}_{\mathrm{M}}^{d-2})\big{)}=\underline{\psi}^{a}_{\mathrm{M}}\big{(}\underline{\alpha}_{\mathrm{M}_{a}}^{d-2}\otimes 1\big{)}=x_{\mathscr{F}},

where \mathscr{F} is any maximal flag of nonempty proper flats of M\mathrm{M} that starts from aa.

For the second statement, note that, for any proper flat FF of rank kk,

xFαMdk=ψFM(φFM(αMdk))=ψFM(α¯MFdk1)=0.x_{F}\hskip 0.85358pt\alpha_{\mathrm{M}}^{d-k}=\psi^{F}_{\mathrm{M}}\big{(}\varphi^{F}_{\mathrm{M}}(\alpha_{\mathrm{M}}^{d-k})\big{)}=\psi^{F}_{\mathrm{M}}\big{(}\underline{\alpha}_{\mathrm{M_{F}}}^{d-k}\otimes 1\big{)}=0.

Using the first statement, we get the conclusion from the identity

αMd=(FxF)αMd1=xαMd1=ψM(φM(αMd1))=ψM(α¯Md1).\alpha_{\mathrm{M}}^{d}=\Big{(}\sum_{F}x_{F}\Big{)}\alpha_{\mathrm{M}}^{d-1}=x_{\varnothing}\hskip 0.85358pt\alpha_{\mathrm{M}}^{d-1}=\psi^{\varnothing}_{\mathrm{M}}\big{(}\varphi^{\varnothing}_{\mathrm{M}}(\alpha_{\mathrm{M}}^{d-1})\big{)}=\psi^{\varnothing}_{\mathrm{M}}\big{(}\underline{\alpha}_{\mathrm{M}}^{d-1}\big{)}.\qed

More generally, the degree of α¯Mdkβ¯Mk\underline{\alpha}_{\mathrm{M}}^{d-k}\underline{\beta}_{\mathrm{M}}^{k} is the kk-th coefficient of the reduced characteristic polynomial of M\mathrm{M} [AHK, Proposition 9.5].

Remark 2.27.

In the setting of Remark 2.13, the element αM\alpha_{\mathrm{M}}, viewed as an element of the Chow ring of the augmented wonderful variety XVX_{V}, is the class of the pullback of the hyperplane (V)(V𝔽)\mathbb{P}(V)\subseteq\mathbb{P}(V\oplus\mathbb{F}).

3. Proofs of the semi-small decompositions and the Poincaré duality theorems

In this section, we prove Theorems 1.1 and 1.2 together with the two Poincaré duality statements in Theorem 1.3. For an element ii of EE, we write πi\pi_{i} and π¯i\underline{\pi}_{i} for the coordinate projections

πi:EEiandπ¯i:E/𝐞EEi/𝐞Ei.\pi_{i}:\mathbb{R}^{E}\longrightarrow\mathbb{R}^{E\setminus i}\ \ \text{and}\ \ \underline{\pi}_{i}:\mathbb{R}^{E}/\langle\mathbf{e}_{E}\rangle\longrightarrow\mathbb{R}^{E\setminus i}/\langle\mathbf{e}_{E\setminus i}\rangle.

Note that πi(ρi)=0\pi_{i}(\rho_{i})=0 and π¯i(ρ¯{i})=0\underline{\pi}_{i}(\underline{\rho}_{\{i\}})=0. In addition, πi(ρS)=ρSi\pi_{i}(\rho_{S})=\rho_{S\setminus i} and π¯i(ρ¯S)=ρ¯Si\underline{\pi}_{i}(\underline{\rho}_{S})=\underline{\rho}_{S\setminus i} for SES\subseteq E.

Proposition 3.1.

Let M\mathrm{M} be a loopless matroid on EE, and let ii be an element of EE.

  1. (1)

    The projection πi\pi_{i} maps any cone of ΠM\Pi_{\mathrm{M}} onto a cone of ΠMi\Pi_{\mathrm{M}\setminus i}.

  2. (2)

    The projection π¯i\underline{\pi}_{i} maps any cone of Π¯M\underline{\Pi}_{\mathrm{M}} onto a cone of Π¯Mi\underline{\Pi}_{\mathrm{M}\setminus i}.

Recall that a linear map defines a morphism of fans Σ1Σ2\Sigma_{1}\to\Sigma_{2} if it maps any cone of Σ1\Sigma_{1} into a cone of Σ2\Sigma_{2} [CLS, Chapter 3]. Thus the above proposition is stronger than the statement that πi\pi_{i} and π¯i\underline{\pi}_{i} induce morphisms of fans.

Proof.

The projection πi\pi_{i} maps σI\sigma_{I\leq\mathscr{F}} onto σIii\sigma_{I\setminus i\leq\mathscr{F}\setminus i}, where i\mathscr{F}\setminus i is the flag of flats of Mi\mathrm{M}\setminus i obtained by removing ii from the members of \mathscr{F}. Similarly, π¯i\underline{\pi}_{i} maps σ¯\underline{\sigma}_{\mathscr{F}} onto σ¯i\underline{\sigma}_{\mathscr{F}\setminus i}. ∎

By Proposition 3.1, the projection πi\pi_{i} defines a map from the toric variety XMX_{\mathrm{M}} of ΠM\Pi_{\mathrm{M}} to the toric variety XMiX_{\mathrm{M}\setminus i} of ΠMi\Pi_{{\mathrm{M}\setminus i}}, and hence the pullback homomorphism CH(Mi)CH(M)\mathrm{CH}(\mathrm{M}\setminus i)\to\mathrm{CH}(\mathrm{M}). Explicitly, the pullback is the graded algebra homomorphism

θi=θMi:CH(Mi)CH(M),xFxF+xFi,\theta_{i}=\theta^{\mathrm{M}}_{i}:\mathrm{CH}(\mathrm{M}\setminus i)\longrightarrow\mathrm{CH}(\mathrm{M}),\qquad x_{F}\longmapsto x_{F}+x_{F\cup i},

where a variable in the target is set to zero if its label is not a flat of M\mathrm{M}. Similarly, π¯i\underline{\pi}_{i} defines a map from the toric variety X¯M\underline{X}_{\mathrm{M}} of Π¯M\underline{\Pi}_{\mathrm{M}} to the toric variety X¯Mi\underline{X}_{\mathrm{M}\setminus i} of Π¯Mi\underline{\Pi}_{{\mathrm{M}\setminus i}}, and hence an algebra homomorphism

θ¯i=θ¯Mi:CH¯(Mi)CH¯(M),xFxF+xFi,\underline{\theta}_{i}=\underline{\theta}^{\mathrm{M}}_{i}:\underline{\mathrm{CH}}(\mathrm{M}\setminus i)\longrightarrow\underline{\mathrm{CH}}(\mathrm{M}),\qquad x_{F}\longmapsto x_{F}+x_{F\cup i},

where a variable in the target is set to zero if its label is not a flat of M\mathrm{M}.

Remark 3.2.

We use the notations introduced in Remark 2.13. Let ViV\setminus i be the image of VV under the ii-th projection 𝔽E𝔽Ei\mathbb{F}^{E}\to\mathbb{F}^{E\setminus i}. We have the commutative diagrams of wonderful varieties and their Chow rings

X¯B{\underline{X}_{\mathrm{B}}}X¯V{\underline{X}_{V}}X¯Bi{\underline{X}_{\mathrm{B}\setminus i}}X¯Vi,{\underline{X}_{V\setminus i},}p¯iB\scriptstyle{\underline{p}_{i}^{\mathrm{B}}}p¯iV\scriptstyle{\underline{p}_{i}^{V}}            CH¯(B){\underline{\operatorname{CH}}(\mathrm{B})}CH¯(M){\underline{\operatorname{CH}}(\mathrm{M})}CH¯(Bi){\underline{\operatorname{CH}}({\mathrm{B}\setminus i})}CH¯(Mi).{\underline{\operatorname{CH}}(\mathrm{M}\setminus i).}

The map p¯iV\underline{p}_{i}^{V} is birational if and only if ii is not a coloop of M\mathrm{M}. By Proposition 3.1, the fibers of p¯iB\underline{p}_{i}^{\mathrm{B}} are at most one-dimensional, and hence the fibers of p¯iV\underline{p}_{i}^{V} are at most one-dimensional. It follows that p¯iV\underline{p}_{i}^{V} is semi-small in the sense of Goresky–MacPherson when ii is not a coloop of M\mathrm{M}.

Similarly, we have the diagrams of augmented wonderful varieties and their Chow rings

XB{X_{\mathrm{B}}}XV{X_{V}}XBi{X_{\mathrm{B}\setminus i}}XVi,{X_{V\setminus i},}piB\scriptstyle{p_{i}^{\mathrm{B}}}piV\scriptstyle{p_{i}^{V}}            CH(B){\operatorname{CH}(\mathrm{B})}CH(M){\operatorname{CH}(\mathrm{M})}CH(Bi){\operatorname{CH}({\mathrm{B}\setminus i})}CH(Mi).{\operatorname{CH}(\mathrm{M}\setminus i).}

The map piVp_{i}^{V} is birational if and only if ii is not a coloop of M\mathrm{M}. By Proposition 3.1, the fibers of piBp_{i}^{\mathrm{B}} are at most one-dimensional, and hence piVp_{i}^{V} is semi-small when ii is not a coloop of M\mathrm{M}.

Numerically, the semi-smallness of p¯iV\underline{p}_{i}^{V} is reflected in the identity

dimxFiCH¯k1(i)=dimxFiCH¯(i)dk2.\operatorname{dim}x_{F\cup i}\underline{\mathrm{CH}}^{k-1}_{(i)}=\operatorname{dim}x_{F\cup i}\underline{\mathrm{CH}}_{(i)}^{d-k-2}.

Similarly, the semi-smallness of piVp_{i}^{V} is reflected in the identity121212The displayed identities follow from Proposition 3.5 and the Poincaré duality parts of Theorem 1.3.

dimxFiCHk1(i)=dimxFiCH(i)dk1.\operatorname{dim}x_{F\cup i}\mathrm{CH}^{k-1}_{(i)}=\operatorname{dim}x_{F\cup i}\mathrm{CH}_{(i)}^{d-k-1}.

For a detailed discussion of semi-small maps in the context of Hodge theory and the decomposition theorem, see [dCM2] and [dCM1].

The element ii is said to be a coloop of M\mathrm{M} if the ranks of M\mathrm{M} and Mi\mathrm{M}\setminus i are not equal. We show that the pullbacks θi\theta_{i} and θ¯i\underline{\theta}_{i} are compatible with the degree maps of M\mathrm{M} and Mi\mathrm{M}\setminus i.

Lemma 3.3.

Suppose that EiE\setminus i is nonempty.

  1. (1)

    If ii is not a coloop of M\mathrm{M}, then θi\theta_{i} commutes with the degree maps:

    degMi=degMθi.\deg_{\mathrm{M}\setminus i}=\deg_{\mathrm{M}}\circ\ \theta_{i}.
  2. (2)

    If ii is not a coloop of M\mathrm{M}, then θ¯i\underline{\theta}_{i} commutes with the degree maps:

    deg¯Mi=deg¯Mθ¯i.\underline{\deg}_{\mathrm{M}\setminus i}=\underline{\deg}_{\mathrm{M}}\circ\ \underline{\theta}_{i}.
  3. (3)

    If ii is a coloop of M\mathrm{M}, we have

    degMi=degMxEiθi=degMαMθi,\deg_{\mathrm{M}\setminus i}=\deg_{\mathrm{M}}\circ\ x_{E\setminus i}\circ\ \theta_{i}=\deg_{\mathrm{M}}\circ\ \alpha_{\mathrm{M}}\circ\ \theta_{i},

    where the middle maps are multiplications by the elements xEix_{E\setminus i} and αM\alpha_{\mathrm{M}}.

  4. (4)

    If ii is a coloop of M\mathrm{M}, we have

    deg¯Mi=deg¯MxEiθ¯i=deg¯Mα¯Mθ¯i,\underline{\deg}_{\mathrm{M}\setminus i}=\underline{\deg}_{\mathrm{M}}\circ\ x_{E\setminus i}\circ\ \underline{\theta}_{i}=\underline{\deg}_{\mathrm{M}}\circ\ \underline{\alpha}_{\mathrm{M}}\circ\ \underline{\theta}_{i},

    where the middle maps are multiplications by the elements xEix_{E\setminus i} and α¯M\underline{\alpha}_{\mathrm{M}}.

Proof.

If ii is not a coloop of M\mathrm{M}, we may choose a basis BB of Mi\mathrm{M}\setminus i that is also a basis of M\mathrm{M}. We have

CHd(Mi)=span(yB)andCHd(M)=span(yB).\mathrm{CH}^{d}(\mathrm{M}\setminus i)=\text{span}(y_{B})\ \ \text{and}\ \ \mathrm{CH}^{d}(\mathrm{M})=\text{span}(y_{B}).

Since θi(yj)=yj\theta_{i}(y_{j})=y_{j} for all jj, the first identity follows. Similarly, by Proposition 2.26,

CH¯d1(Mi)=span(α¯Mid1)andCH¯d1(M)=span(α¯Md1).\underline{\mathrm{CH}}^{d-1}(\mathrm{M}\setminus i)=\text{span}(\underline{\alpha}_{\mathrm{M}\setminus i}^{d-1})\ \ \text{and}\ \ \underline{\mathrm{CH}}^{d-1}(\mathrm{M})=\text{span}(\underline{\alpha}_{\mathrm{M}}^{d-1}).

Since θ¯i(α¯Mi)=α¯M\underline{\theta}_{i}(\underline{\alpha}_{\mathrm{M}\setminus i})=\underline{\alpha}_{\mathrm{M}} when ii is not a coloop, the second identity follows.

Suppose now that ii is a coloop of M\mathrm{M}. In this case, Mi=MEi\mathrm{M}\setminus i=\mathrm{M}^{E\setminus i}, and hence

φEiMθi=identity of CH(Mi)andφ¯EiMθ¯i=identity of CH¯(Mi).\varphi^{E\setminus i}_{\mathrm{M}}\circ\ \theta_{i}=\text{identity of $\operatorname{CH}(\mathrm{M}\setminus i)$}\ \ \text{and}\ \ \underline{\varphi}^{E\setminus i}_{\mathrm{M}}\circ\ \underline{\theta}_{i}=\text{identity of $\underline{\operatorname{CH}}(\mathrm{M}\setminus i)$}.

Using the compatibility of the pushforward ψEiM\psi^{E\setminus i}_{\mathrm{M}} with the degree maps, we have

degMi=degMψEiM=degMψEiMφEiMθi=degMxEiθi.\deg_{\mathrm{M}\setminus i}=\deg_{\mathrm{M}}\circ\ \psi^{E\setminus i}_{\mathrm{M}}=\deg_{\mathrm{M}}\circ\ \psi^{E\setminus i}_{\mathrm{M}}\circ\ \varphi^{E\setminus i}_{\mathrm{M}}\circ\ \theta_{i}=\deg_{\mathrm{M}}\circ\ x_{E\setminus i}\circ\ \theta_{i}.

Since θi(αMi)=αMxEi\theta_{i}(\alpha_{\mathrm{M}\setminus i})=\alpha_{\mathrm{M}}-x_{E\setminus i} when ii is a coloop of M\mathrm{M}, the above implies

degMi=degMxEiθi=degM(αMθi(αMi))θi=degMαMθi,\deg_{\mathrm{M}\setminus i}=\deg_{\mathrm{M}}\circ\ x_{E\setminus i}\circ\ \theta_{i}=\deg_{\mathrm{M}}\circ\ \big{(}\alpha_{\mathrm{M}}-\theta_{i}(\alpha_{\mathrm{M}\setminus i})\big{)}\circ\ \theta_{i}=\deg_{\mathrm{M}}\circ\ \alpha_{\mathrm{M}}\circ\ \theta_{i},

The identities for deg¯Mi\underline{\deg}_{\mathrm{M}\setminus i} can be obtained in a similar way. ∎

Proposition 3.4.

If CH(Mi)\mathrm{CH}(\mathrm{M}\setminus i) satisfies the Poincaré duality part of Theorem 1.3, then θi\theta_{i} is injective. Also, if CH¯(Mi)\underline{\mathrm{CH}}(\mathrm{M}\setminus i) satisfies the Poincaré duality part of Theorem 1.3, then θ¯i\underline{\theta}_{i} is injective.

Proof.

The proof is essentially identical to that of Proposition 2.19. ∎

For a flat FF in 𝒮i\mathscr{S}_{i}, we write θiFi\theta_{i}^{F\cup i} for the pullback map between the augmented Chow rings obtained from the deletion of ii from the localization MFi\mathrm{M}^{F\cup i}:

θiFi:CH(MF)CH(MFi).\theta_{i}^{F\cup i}:\mathrm{CH}(\mathrm{M}^{F})\to\mathrm{CH}(\mathrm{M}^{F\cup i}).

Similarly, for a flat FF in 𝒮¯i\underline{\mathscr{S}}_{i}, we write θ¯Fii\underline{\theta}^{F\cup i}_{i} for the pullback map between the Chow rings obtained from the deletion of ii from the localization MFi\mathrm{M}^{F\cup i}:

θ¯iFi:CH¯(MF)CH¯(MFi).\underline{\theta}_{i}^{F\cup i}:\underline{\mathrm{CH}}(\mathrm{M}^{F})\to\underline{\mathrm{CH}}(\mathrm{M}^{F\cup i}).

Note that ii is a coloop of MFi\mathrm{M}^{F\cup i} in these cases.

Proposition 3.5.

The summands appearing in Theorems 1.1 and 1.2 can be described as follows.

  1. (1)

    If F𝒮iF\in\mathscr{S}_{i}, then xFiCH(i)=ψMFi(CH¯(MFi)θiFiCH(MF))x_{F\cup i}\mathrm{CH}_{(i)}=\psi_{\mathrm{M}}^{F\cup i}\big{(}\underline{\mathrm{CH}}(\mathrm{M}_{F\cup i})\otimes\theta_{i}^{F\cup i}\mathrm{CH}(\mathrm{M}^{F})\big{)}.

  2. (2)

    If F𝒮¯iF\in\underline{\mathscr{S}}_{i}, then xFiCH¯(i)=ψ¯MFi(CH¯(MFi)θ¯iFiCH¯(MF))x_{F\cup i}\underline{\mathrm{CH}}_{(i)}=\underline{\psi}_{\mathrm{M}}^{F\cup i}\big{(}\underline{\mathrm{CH}}(\mathrm{M}_{F\cup i})\otimes\underline{\theta}_{i}^{F\cup i}\underline{\mathrm{CH}}(\mathrm{M}^{F})\big{)}.

  3. (3)

    If ii is a coloop of M\mathrm{M}, then xEiCH(i)=ψEiMCH(Mi)x_{E\setminus i}\operatorname{CH}_{(i)}=\psi^{E\setminus i}_{\mathrm{M}}\operatorname{CH}(\mathrm{M}\setminus i) and xEiCH¯(i)=ψ¯EiMCH¯(Mi)x_{E\setminus i}\underline{\operatorname{CH}}_{(i)}=\underline{\psi}^{E\setminus i}_{\mathrm{M}}\underline{\operatorname{CH}}(\mathrm{M}\setminus i).

It follows, assuming Poincaré duality for the Chow rings,131313We need Poincaré duality for CH¯(MF)\underline{\operatorname{CH}}(\mathrm{M}^{F}), CH(MF)\operatorname{CH}(\mathrm{M}^{F}), CH¯(MFi)\underline{\operatorname{CH}}(\mathrm{M}^{F\cup i}), CH(MFi)\operatorname{CH}(\mathrm{M}^{F\cup i}), and CH¯(MFi)\underline{\operatorname{CH}}(\mathrm{M}_{F\cup i}). that

xFiCH(i)CH¯(MFi)CH(MF)[1]andxFiCH¯(i)CH¯(MFi)CH¯(MF)[1].x_{F\cup i}\mathrm{CH}_{(i)}\cong\underline{\mathrm{CH}}(\mathrm{M}_{F\cup i})\otimes\mathrm{CH}(\mathrm{M}^{F})[-1]\ \ \text{and}\ \ x_{F\cup i}\underline{\mathrm{CH}}_{(i)}\cong\underline{\mathrm{CH}}(\mathrm{M}_{F\cup i})\otimes\underline{\mathrm{CH}}(\mathrm{M}^{F})[-1].

Therefore, again assuming Poincaré duality for the Chow rings, we have

dimxFiCH¯k1(i)=dimxFiCH¯(i)dk2anddimxFiCHk1(i)=dimxFiCH(i)dk1.\operatorname{dim}x_{F\cup i}\underline{\mathrm{CH}}^{k-1}_{(i)}=\operatorname{dim}x_{F\cup i}\underline{\mathrm{CH}}_{(i)}^{d-k-2}\ \ \text{and}\ \ \operatorname{dim}x_{F\cup i}\mathrm{CH}^{k-1}_{(i)}=\operatorname{dim}x_{F\cup i}\mathrm{CH}_{(i)}^{d-k-1}.
Proof.

We prove the first statement. The proof of the second statement is essentially identical. The third statement is a straightforward consequence of the fact that φEiMθi\varphi^{E\setminus i}_{\mathrm{M}}\circ\theta_{i} and φ¯EiMθ¯i\underline{\varphi}^{E\setminus i}_{\mathrm{M}}\circ\underline{\theta}_{i} are the identity maps when ii is a coloop.

Let FF be a flat in 𝒮i\mathscr{S}_{i}. It is enough to show that

φFiM(CH(i))=CH¯(MFi)θiFiCH(MF),\varphi^{F\cup i}_{\mathrm{M}}\big{(}\mathrm{CH}_{(i)}\big{)}=\underline{\mathrm{CH}}(\mathrm{M}_{F\cup i})\otimes\theta_{i}^{F\cup i}\mathrm{CH}(\mathrm{M}^{F}),

since the result will then follow by applying ψMFi\psi_{\mathrm{M}}^{F\cup i}. The projection πi\pi_{i} maps the ray ρFi\rho_{F\cup i} to the ray ρF\rho_{F}, and hence πi\pi_{i} defines morphisms of fans

starρFiΠM{\text{star}_{\rho_{F\cup i}}\Pi_{\mathrm{M}}}Π¯MFi×ΠMFi{\underline{\Pi}_{\mathrm{M}_{F\cup i}}\times\Pi_{\mathrm{M}^{F\cup i}}}Π¯(M/i)F×ΠMFi{\underline{\Pi}_{(\mathrm{M}/i)_{F}}\times\Pi_{\mathrm{M}^{F\cup i}}}starρFΠMi{\text{star}_{\rho_{F}}\Pi_{\mathrm{M}\setminus i}}Π¯(Mi)F×Π(Mi)F{\underline{\Pi}_{(\mathrm{M}\setminus i)_{F}}\times\Pi_{(\mathrm{M}\setminus i)^{F}}}Π¯(Mi)F×ΠMF,{\underline{\Pi}_{(\mathrm{M}\setminus i)_{F}}\times\Pi_{\mathrm{M}^{F}},}πi\scriptstyle{\pi^{\prime}_{i}}ιFi\scriptstyle{\iota_{F\cup i}}πi\scriptstyle{\pi^{\prime\prime}_{i}}πi\scriptstyle{\pi^{\prime\prime\prime}_{i}}ιF\scriptstyle{\iota_{F}}

where ιFi\iota_{F\cup i} and ιF\iota_{F} are the isomorphisms in Proposition 2.7. The main point is that the matroid (M/i)F(\mathrm{M}/i)_{F} is a quotient of (Mi)F(\mathrm{M}\setminus i)_{F}. In other words, we have the inclusion of Bergman fans

Π¯(M/i)FΠ¯(Mi)F.\underline{\Pi}_{(\mathrm{M}/i)_{F}}\subseteq\underline{\Pi}_{(\mathrm{M}\setminus i)_{F}}.

Therefore, the morphism πi\pi_{i}^{\prime\prime\prime} admits the factorization

Π¯(M/i)F×ΠMFi{\underline{\Pi}_{(\mathrm{M}/i)_{F}}\times\Pi_{\mathrm{M}^{F\cup i}}}Π¯(M/i)F×ΠMF{\underline{\Pi}_{(\mathrm{M}/i)_{F}}\times\Pi_{\mathrm{M}^{F}}}Π¯(Mi)F×ΠMF,{\underline{\Pi}_{(\mathrm{M}\setminus i)_{F}}\times\Pi_{\mathrm{M}^{F}},}

where the second map induces a surjective pullback map qq between the Chow rings. By the equality (M/i)F=MFi(\mathrm{M}/i)_{F}=\mathrm{M}_{F\cup i}, we have the commutative diagram of pullback maps between the Chow rings

CH(Mi){\mathrm{CH}(\mathrm{M}\setminus i)}CH(M){\mathrm{CH}(\mathrm{M})}CH¯((Mi)F)CH((Mi)F){\underline{\mathrm{CH}}((\mathrm{M}\setminus i)_{F})\otimes\mathrm{CH}((\mathrm{M}\setminus i)^{F})}CH¯(MFi)CH(MF){\underline{\mathrm{CH}}(\mathrm{M}_{F\cup i})\otimes\mathrm{CH}(\mathrm{M}^{F})}CH¯(MFi)CH(MFi).{\underline{\mathrm{CH}}(\mathrm{M}_{F\cup i})\otimes\mathrm{CH}(\mathrm{M}^{F\cup i}).}θi\scriptstyle{\theta_{i}}φFMi\scriptstyle{\varphi^{F}_{\mathrm{M}\setminus i}}φFiM\scriptstyle{\varphi^{F\cup i}_{\mathrm{M}}}q\scriptstyle{q}1θFii\scriptstyle{1\otimes\theta^{F\cup i}_{i}}

The conclusion follows from the surjectivity of the pullback maps φFMi\varphi^{F}_{\mathrm{M}\setminus i} and qq. ∎

Remark 3.6.

Since ii is a coloop in MFi\mathrm{M}^{F\cup i} when F𝒮iF\in\mathscr{S}_{i} or F𝒮¯iF\in\underline{\mathscr{S}}_{i}, Proposition 3.5 implies that

xFiCH(i)d1=0for F𝒮iandxFiCH¯(i)d2=0for F𝒮¯i.x_{F\cup i}\mathrm{CH}_{(i)}^{d-1}=0\;\text{for $F\in\mathscr{S}_{i}$}\ \ \text{and}\ \ x_{F\cup i}\underline{\mathrm{CH}}_{(i)}^{d-2}=0\;\text{for $F\in\underline{\mathscr{S}}_{i}$}.
Proposition 3.7.

The Poincaré pairing on the summands appearing in Theorems 1.1 and 1.2 can be described as follows.

  1. (1)

    If F𝒮iF\in\mathscr{S}_{i}, then for any μ1,μ2CH¯(MFi)CH(MF)\mu_{1},\mu_{2}\in\underline{\mathrm{CH}}(\mathrm{M}_{F\cup i})\otimes{\mathrm{CH}}(\mathrm{M}^{F}) of complementary degrees,

    degM(ψFiM(1θiFi(μ1))ψFiM(1θiFi(μ2)))=deg¯MFidegMF(μ1μ2).\deg_{\mathrm{M}}\big{(}\psi^{F\cup i}_{\mathrm{M}}\big{(}1\otimes\theta_{i}^{F\cup i}(\mu_{1})\big{)}\cdot\psi^{F\cup i}_{\mathrm{M}}\big{(}1\otimes\theta_{i}^{F\cup i}(\mu_{2})\big{)}\big{)}=-\underline{\deg}_{\mathrm{M}_{F\cup i}}\otimes\deg_{\mathrm{M}^{F}}(\mu_{1}\mu_{2}).
  2. (2)

    If F𝒮¯iF\in\underline{\mathscr{S}}_{i}, then for any ν1,ν2CH¯(MFi)CH¯(MF)\nu_{1},\nu_{2}\in\underline{\mathrm{CH}}(\mathrm{M}_{F\cup i})\otimes\underline{\mathrm{CH}}(\mathrm{M}^{F}) of complementary degrees,

    deg¯M(ψ¯FiM(1θ¯iFi(ν1))ψ¯FiM(1θ¯iFi(ν2)))=deg¯MFideg¯MF(ν1ν2).\underline{\deg}_{\mathrm{M}}\big{(}\underline{\psi}^{F\cup i}_{\mathrm{M}}\big{(}1\otimes\underline{\theta}_{i}^{F\cup i}(\nu_{1})\big{)}\cdot\underline{\psi}^{F\cup i}_{\mathrm{M}}\big{(}1\otimes\underline{\theta}_{i}^{F\cup i}(\nu_{2})\big{)}\big{)}=-\underline{\deg}_{\mathrm{M}_{F\cup i}}\otimes\underline{\deg}_{\mathrm{M}^{F}}(\nu_{1}\nu_{2}).

It follows, assuming Poincaré duality for the Chow rings,141414We need Poincaré duality for CH¯(MF)\underline{\operatorname{CH}}(\mathrm{M}^{F}), CH(MF)\operatorname{CH}(\mathrm{M}^{F}), CH¯(MFi)\underline{\operatorname{CH}}(\mathrm{M}^{F\cup i}), CH(MFi)\operatorname{CH}(\mathrm{M}^{F\cup i}), and CH¯(MFi)\underline{\operatorname{CH}}(\mathrm{M}_{F\cup i}). that the restriction of the Poincaré pairing of CH(M)\operatorname{CH}(\mathrm{M}) to the subspace xFiCH(i)x_{F\cup i}\operatorname{CH}_{(i)} is non-degenerate, and the restriction of the Poincaré pairing of CH¯(M)\underline{\operatorname{CH}}(\mathrm{M}) to the subspace xFiCH¯(i)x_{F\cup i}\underline{\operatorname{CH}}_{(i)} is non-degenerate.

Proof.

We prove the first identity. The second identity can be proved in the same way.

Since the pushforward ψFiM\psi^{F\cup i}_{\mathrm{M}} is a CH(M)\mathrm{CH}(\mathrm{M})-module homomorphism, the left-hand side is

degM(ψFiM(φFiMψFiM(1θiFi(μ1))(1θiFi(μ2)))).\deg_{\mathrm{M}}\big{(}\psi^{F\cup i}_{\mathrm{M}}\big{(}\varphi^{F\cup i}_{\mathrm{M}}\psi^{F\cup i}_{\mathrm{M}}\big{(}1\otimes\theta_{i}^{F\cup i}(\mu_{1})\big{)}\cdot\big{(}1\otimes\theta_{i}^{F\cup i}(\mu_{2})\big{)}\big{)}\big{)}.

The pushforward commutes with the degree maps, so the above is equal to

deg¯MFidegMFi(φFiMψFiM(1θiFi(μ1))(1θiFi(μ2))).\underline{\deg}_{\mathrm{M}_{F\cup i}}\otimes\deg_{\mathrm{M}^{F\cup i}}\big{(}\varphi^{F\cup i}_{\mathrm{M}}\psi^{F\cup i}_{\mathrm{M}}\big{(}1\otimes\theta_{i}^{F\cup i}(\mu_{1})\big{)}\cdot\big{(}1\otimes\theta_{i}^{F\cup i}(\mu_{2})\big{)}\big{)}.

Using that the composition φFiMψFiM\varphi^{F\cup i}_{\mathrm{M}}\psi^{F\cup i}_{\mathrm{M}} is multiplication by φFiM(xFi)\varphi^{F\cup i}_{\mathrm{M}}(x_{F\cup i}), we get

deg¯MFidegMFi((1αMFi+β¯MFi1)(1θiFi(μ1))(1θiFi(μ2))).-\underline{\deg}_{\mathrm{M}_{F\cup i}}\otimes\deg_{\mathrm{M}^{F\cup i}}\big{(}\big{(}1\otimes\alpha_{\mathrm{M}^{F\cup i}}+\underline{\beta}_{\mathrm{M}_{F\cup i}}\otimes 1\big{)}\cdot\big{(}1\otimes\theta_{i}^{F\cup i}(\mu_{1})\big{)}\cdot\big{(}1\otimes\theta_{i}^{F\cup i}(\mu_{2})\big{)}\big{)}.

Since ii is a coloop of MFi\mathrm{M}^{F\cup i}, the expression simplifies to

deg¯MFidegMFi((1αMFi)(1θiFi(μ1))(1θiFi(μ2))).-\underline{\deg}_{\mathrm{M}_{F\cup i}}\otimes\deg_{\mathrm{M}^{F\cup i}}\big{(}\big{(}1\otimes\alpha_{\mathrm{M}^{F\cup i}}\big{)}\cdot\big{(}1\otimes\theta_{i}^{F\cup i}(\mu_{1})\big{)}\cdot\big{(}1\otimes\theta_{i}^{F\cup i}(\mu_{2})\big{)}\big{)}.

Now the third part of Lemma 3.3 shows that the above quantity is the right-hand side of the formula in statement (1). ∎

Lemma 3.8.

If flats F1F_{1}, F2F_{2} are in 𝒮i\mathscr{S}_{i} and F1F_{1} is a proper subset of F2F_{2}, then

xF1ixF2ixF1iCH(i).x_{F_{1}\cup i}\,x_{F_{2}\cup i}\in x_{F_{1}\cup i}\mathrm{CH}_{(i)}.

Similarly, if F1F_{1}, F2F_{2} are in 𝒮¯i\underline{\mathscr{S}}_{i} and F1F_{1} is a proper subset of F2F_{2}, then

xF1ixF2ixF1iCH¯(i).x_{F_{1}\cup i}\,x_{F_{2}\cup i}\in x_{F_{1}\cup i}\underline{\mathrm{CH}}_{(i)}.
Proof.

Since F1iF_{1}\cup i is not comparable to F2F_{2}, we have

xF1ixF2i=xF1i(xF2+xF2i)=xF1iθi(xF2).x_{F_{1}\cup i}\,x_{F_{2}\cup i}=x_{F_{1}\cup i}(x_{F_{2}}+x_{F_{2}\cup i})=x_{F_{1}\cup i}\theta_{i}(x_{F_{2}}).

The second part follows from the same argument. ∎

Proof of Theorem 1.1, Theorem 1.2, and parts (1) and (4) of Theorem 1.3.

All the summands in the proposed decompositions are cyclic, and therefore indecomposable in the category of graded modules.151515By [CF, Corollary 2] or [GG, Theorem 3.2], the indecomposability of the summands in the category of graded modules implies the indecomposability of the summands in the category of modules. We prove the decompositions by induction on the cardinality of the ground set EE. If EE is empty, then Theorem 1.1, Theorem 1.2, and part (1) of Theorem 1.3 are vacuous, while part (4) of Theorem 1.3 is trivial. Furthermore, all of these results are trivial when EE is a singleton. Thus, we may assume that ii is an element of EE, that EiE\setminus i is nonempty, and that all the results hold for loopless matroids whose ground set is a proper subset of EE.

First we assume that ii is not a coloop. Let us show that the terms in the right-hand side of the decomposition (D1\mathrm{D}_{1}) are orthogonal. Multiplying CH(i)\mathrm{CH}_{(i)} and xFiCH(i)x_{F\cup i}\mathrm{CH}_{(i)} lands in xFiCH(i)x_{F\cup i}\mathrm{CH}_{(i)}, and this ideal vanishes in degree dd by Remark 3.6, so they are orthogonal. On the other hand, the product of xF1iCH(i)x_{F_{1}\cup i}\mathrm{CH}_{(i)} and xF2iCH(i)x_{F_{2}\cup i}\mathrm{CH}_{(i)} vanishes if F1,F2𝒮iF_{1},F_{2}\in\mathscr{S}_{i} are not comparable, while if F1<F2F_{1}<F_{2} or F2<F1F_{2}<F_{1}, the product is contained in xF1iCH(i)x_{F_{1}\cup i}\mathrm{CH}_{(i)} or xF2iCH(i)x_{F_{2}\cup i}\mathrm{CH}_{(i)} respectively, by Lemma 3.8. So these terms are also orthogonal.

It follows from the induction hypothesis and Lemma 3.3 that the restriction of the Poincaré pairing of CH(M)\mathrm{CH}(\mathrm{M}) to CH(i)\mathrm{CH}_{(i)} is non-degenerate. By Proposition 3.5, Proposition 3.7, and the induction hypothesis, the restriction of the Poincaré pairing of CH(M)\mathrm{CH}(\mathrm{M}) to any other summand xFiCH(i)x_{F\cup i}\mathrm{CH}_{(i)} is also non-degenerate. Therefore, we can conclude that the sum on the right-hand side of (D1\mathrm{D}_{1}) is a direct sum with a non-degenerate Poincaré pairing.

To complete the proof of the decomposition (D1\mathrm{D}_{1}) and the Poincaré duality theorem for CH(M)\mathrm{CH}(\mathrm{M}), we must show that the direct sum

CH(i)F𝒮ixFiCH(i)\mathrm{CH}_{(i)}\oplus\bigoplus_{F\in\mathscr{S}_{i}}x_{F\cup i}\mathrm{CH}_{(i)}

is equal to all of CH(M)\mathrm{CH}(\mathrm{M}). This is obvious in degree 0. To see that it holds in degree 11, it is enough to check that xGx_{G} is contained in the direct sum for any proper flat GG of M\mathrm{M}. If GiG\setminus i is a not flat of M\mathrm{M}, then xG=θi(xGi).x_{G}=\theta_{i}(x_{G\setminus i}). If GiG\setminus i is a flat of M\mathrm{M}, then either Gi𝒮iG\setminus i\in\mathscr{S}_{i} or G𝒮iG\in\mathscr{S}_{i}. In the first case, xGx_{G} is an element of the summand indexed by GiG\setminus i. In the second case, xG=θi(xG)xGiCH(i)+xGiCH(i).x_{G}=\theta_{i}(x_{G})-x_{G\cup i}\in\mathrm{CH}_{(i)}+x_{G\cup i}\mathrm{CH}_{(i)}.

Since our direct sum is a sum of CH(Mi)\mathrm{CH}(\mathrm{M}\setminus i)-modules and it includes the degree 0 and 11 parts of CH(M)\mathrm{CH}(\mathrm{M}), it will suffice to show that CH(M)\mathrm{CH}(\mathrm{M}) is generated in degrees 0 and 11 as a graded CH(Mi)\mathrm{CH}(\mathrm{M}\setminus i)-module. In other words, we need to show that

CH1(i)CHk(M)=CHk+1(M)for any k1.\mathrm{CH}^{1}_{(i)}\cdot\mathrm{CH}^{k}(\mathrm{M})=\mathrm{CH}^{k+1}(\mathrm{M})\ \ \text{for any $k\geq 1$.}

We first prove the equality when k=1k=1. Since we have proved that the decomposition (D1\mathrm{D}_{1}) holds in degree 11, we know that

CH2(M)=CH1(M)CH1(M)=(CH1(i)F𝒮ixFi)(CH1(i)F𝒮ixFi).\mathrm{CH}^{2}(\mathrm{M})=\mathrm{CH}^{1}(\mathrm{M})\cdot\mathrm{CH}^{1}(\mathrm{M})=\left(\mathrm{CH}^{1}_{(i)}\oplus\bigoplus_{F\in\mathscr{S}_{i}}\mathbb{Q}x_{F\cup i}\right)\cdot\left(\mathrm{CH}^{1}_{(i)}\oplus\bigoplus_{F\in\mathscr{S}_{i}}\mathbb{Q}x_{F\cup i}\right).

Using Lemma 3.8, we may reduce the problem to showing that

xFi2CH1(i)CH1(M)for any F𝒮i.x_{F\cup i}^{2}\in\mathrm{CH}^{1}_{(i)}\cdot\mathrm{CH}^{1}(\mathrm{M})\ \ \text{for any $F\in\mathscr{S}_{i}$.}

We can rewrite the relation 0=xFyi0=x_{F}y_{i} in the augmented Chow ring of M\mathrm{M} as

0\displaystyle 0 =(θi(xF)xFi)iGxG\displaystyle=(\theta_{i}(x_{F})-x_{F\cup i})\sum_{i\notin G}x_{G}
=θi(xF)(iGxG)xFi(GFxG),\displaystyle=\theta_{i}(x_{F})\Big{(}\sum_{i\notin G}x_{G}\Big{)}-x_{F\cup i}\Big{(}\sum_{G\leq F}x_{G}\Big{)},
=θi(xF)(iGxG)(θi(xF)xF)(G<FxG)xFixF\displaystyle=\theta_{i}(x_{F})\Big{(}\sum_{i\notin G}x_{G}\Big{)}-(\theta_{i}(x_{F})-x_{F})\Big{(}\sum_{G<F}x_{G}\Big{)}-x_{F\cup i}x_{F}
=θi(xF)(iGxGG<FxG)+xF(G<FxG)xFiθi(xF)+xFi2,\displaystyle=\theta_{i}(x_{F})\Big{(}\sum_{i\notin G}x_{G}-\sum_{G<F}x_{G}\Big{)}+x_{F}\Big{(}\sum_{G<F}x_{G}\Big{)}-x_{F\cup i}\theta_{i}(x_{F})+x_{F\cup i}^{2},

thus reducing the problem to showing that

xFxGCH1(i)CH1(M)for any G<F𝒮i.x_{F}x_{G}\in\mathrm{CH}^{1}_{(i)}\cdot\mathrm{CH}^{1}(\mathrm{M})\ \ \text{for any $G<F\in\mathscr{S}_{i}$.}

The collection 𝒮i\mathscr{S}_{i} is downward closed, meaning that if G<F𝒮iG<F\in\mathscr{S}_{i}, then G𝒮iG\in\mathscr{S}_{i}; therefore,

xFxG=(θi(xF)xFi)(θi(xG)xGi).x_{F}x_{G}=(\theta_{i}(x_{F})-x_{F\cup i})(\theta_{i}(x_{G})-x_{G\cup i}).

Lemma 3.8 tells us that xFixGiCH1(i)CH1(M)x_{F\cup i}x_{G\cup i}\in\mathrm{CH}^{1}_{(i)}\cdot\mathrm{CH}^{1}(\mathrm{M}), thus so is xFxGx_{F}x_{G}.

We next prove the equality when k2k\geq 2. In this case, we use the result for k=1k=1 along with the fact that the algebra CH(M)\mathrm{CH}(\mathrm{M}) is generated in degree 11 to conclude that

CH1(i)CHk(M)=CH1(i)CH1(M)CHk1(M)=CH2(M)CHk1(M)=CHk+1(M).\mathrm{CH}^{1}_{(i)}\cdot\mathrm{CH}^{k}(\mathrm{M})=\mathrm{CH}^{1}_{(i)}\cdot\mathrm{CH}^{1}(\mathrm{M})\cdot\mathrm{CH}^{k-1}(\mathrm{M})=\mathrm{CH}^{2}(\mathrm{M})\cdot\mathrm{CH}^{k-1}(\mathrm{M})=\mathrm{CH}^{k+1}(\mathrm{M}).

This completes the proof of the decomposition (D1\mathrm{D}_{1}) and the Poincaré duality theorem for CH(M)\mathrm{CH}(\mathrm{M}) when there is an element ii that is not a coloop of M\mathrm{M}.

The proof when ii is a coloop is almost the same; we explain the places where something different must be said. The orthogonality of xEiCH(i)x_{E\setminus i}\mathrm{CH}_{(i)} and xFiCH(i)x_{F\cup i}\mathrm{CH}_{(i)} for F𝒮iF\in\mathscr{S}_{i} follows because EiE\setminus i and FiF\cup i are incomparable. To show that the right-hand side of (D2\mathrm{D}_{2}) spans CH(M)\mathrm{CH}(\mathrm{M}), one extra statement we need to check is that

xEi2CH1(i)CH1(M).x_{E\setminus i}^{2}\in\mathrm{CH}^{1}_{(i)}\cdot\mathrm{CH}^{1}(\mathrm{M}).

Since ii is a coloop, 𝒮i\mathscr{S}_{i} is the set of all flats properly contained in EiE\setminus i, and we have

0=xEiyi=iFxFxEi=xEi2+F𝒮ixEixF=xEi2+F𝒮ixEiθi(xF),0=x_{E\setminus i}y_{i}=\sum_{i\notin F}x_{F}x_{E\setminus i}=x_{E\setminus i}^{2}+\sum_{F\in\mathscr{S}_{i}}x_{E\setminus i}x_{F}=x_{E\setminus i}^{2}+\sum_{F\in\mathscr{S}_{i}}x_{E\setminus i}\theta_{i}(x_{F}),

where the last equality follows because EiE\setminus i and FiF\cup i are not comparable. Thus

xEi2=F𝒮ixEiθi(xF)CH1(i)CH1(M).x_{E\setminus i}^{2}=-\sum_{F\in\mathscr{S}_{i}}x_{E\setminus i}\theta_{i}(x_{F})\in\mathrm{CH}^{1}_{(i)}\cdot\mathrm{CH}^{1}(\mathrm{M}).

By the induction hypothesis, we know CH(Mi)\mathrm{CH}(\mathrm{M}\setminus i) satisfies the Poincaré duality theorem. By the coloop case of Lemma 3.3, the Poincaré pairing on CH(M)\mathrm{CH}(\mathrm{M}) restricts to a perfect pairing between CH(i)\mathrm{CH}_{(i)} and xEiCH(i)x_{E\setminus i}\mathrm{CH}_{(i)}. Since CH(i)\mathrm{CH}_{(i)} is a subring of CH(M)\mathrm{CH}(\mathrm{M}) and is zero in degree dd, the restriction of the Poincaré pairing on CH(M)\mathrm{CH}(\mathrm{M}) to CH(i)\mathrm{CH}_{(i)} is zero. Therefore, the subspaces CH(i)\mathrm{CH}_{(i)} and xEiCH(i)x_{E\setminus i}\mathrm{CH}_{(i)} intersect trivially, and the restriction of the Poincaré pairing on CH(M)\mathrm{CH}(\mathrm{M}) to CH(i)xEiCH(i)\mathrm{CH}_{(i)}\oplus x_{E\setminus i}\mathrm{CH}_{(i)} is non-degenerate. This completes the proof of the theorems about CH(M)\mathrm{CH}(\mathrm{M}) when ii is a coloop.

We observe that the surjectivity of the pullback φM\varphi^{\varnothing}_{\mathrm{M}} gives the equality

CH¯1(i)CH¯k(M)=CH¯k+1(M)for any k1.\underline{\mathrm{CH}}^{1}_{(i)}\cdot\underline{\mathrm{CH}}^{k}(\mathrm{M})=\underline{\mathrm{CH}}^{k+1}(\mathrm{M})\ \ \text{for any $k\geq 1$.}

The proof of the theorems about CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}) then follows by an argument identical to the one used for CH(M)\mathrm{CH}(\mathrm{M}). ∎

4. Proofs of the hard Lefschetz theorems and the Hodge–Riemann relations

In this section, we prove Theorem 1.3. Parts (1) and (4) have already been proved in the previous section. We will first prove parts (2) and (3) by induction on the cardinality of EE. The proof of parts (5) and (6) is nearly identical to the proof of parts (2) and (3), with the added nuance that we use parts (2) and (3) for the matroid M\mathrm{M} in the proof of parts (5) and (6) for the matroid M\mathrm{M}.

For any fan Σ\Sigma, we will say that Σ\Sigma satisfies the hard Lefschetz theorem or the Hodge–Riemann relations with respect to some piecewise linear function on Σ\Sigma if the ring CH(Σ)\mathrm{CH}(\Sigma) satisfies the hard Lefschetz theorem or the Hodge–Riemann relations with respect to the corresponding element of CH1(Σ)\mathrm{CH}^{1}(\Sigma).

Proof of Theorem 1.3, parts (2) and (3).

The statements are trivial when the cardinality of EE is 0 or 11, so we will assume throughout the proof that the cardinality of EE is at least 22.

Let B\mathrm{B} be the Boolean matroid on EE. By the induction hypothesis, we know that for every nonempty proper flat FF of M\mathrm{M}, the fans Π¯MF\underline{\Pi}_{\mathrm{M}_{F}} and Π¯MF\underline{\Pi}_{\mathrm{M}^{F}} satisfy the hard Lefschetz theorem and the Hodge–Riemann relations with respect to any strictly convex piecewise linear functions on Π¯BF\underline{\Pi}_{\mathrm{B}_{F}} and Π¯BF\underline{\Pi}_{\mathrm{B}^{F}}, respectively. By [AHK, Proposition 7.7], this implies that for every nonempty proper flat FF of M\mathrm{M}, the product Π¯MF×Π¯MF\underline{\Pi}_{\mathrm{M}_{F}}\times\underline{\Pi}_{\mathrm{M}^{F}} satisfies the hard Lefschetz theorem and the Hodge–Riemann relations with respect to any strictly convex piecewise linear function on Π¯BF×Π¯BF\underline{\Pi}_{\mathrm{B}_{F}}\times\underline{\Pi}_{\mathrm{B}^{F}}. In other words, Π¯M\underline{\Pi}_{\mathrm{M}} satisfies the local Hodge–Riemann relations [AHK, Definition 7.14]:

The star of any ray in Π¯M\underline{\Pi}_{\mathrm{M}} satisfies the Hodge–Riemann relations.

This in turn implies that Π¯M\underline{\Pi}_{\mathrm{M}} satisfies the hard Lefschetz theorem with respect to any strictly convex piecewise linear function on Π¯B\underline{\Pi}_{\mathrm{B}} [AHK, Proposition 7.15]. It remains to prove only that Π¯M\underline{\Pi}_{\mathrm{M}} satisfies the Hodge–Riemann relations with respect to any strictly convex piecewise linear function on Π¯B\underline{\Pi}_{\mathrm{B}}.

Let \ell be a piecewise linear function on Π¯B\underline{\Pi}_{\mathrm{B}}, and let HR¯k(M)\underline{\mathrm{HR}}_{\ell}^{k}(\mathrm{M}) be the Hodge–Riemann form

HR¯k(M):CH¯k(M)×CH¯k(M),(η1,η2)(1)kdeg¯M(d2k1η1η2).\underline{\mathrm{HR}}_{\ell}^{k}(\mathrm{M}):\underline{\mathrm{CH}}^{k}(\mathrm{M})\times\underline{\mathrm{CH}}^{k}(\mathrm{M})\longrightarrow\mathbb{Q},\qquad(\eta_{1},\eta_{2})\longmapsto(-1)^{k}\underline{\deg}_{\mathrm{M}}(\ell^{d-2k-1}\eta_{1}\eta_{2}).

By [AHK, Proposition 7.6], the fan Π¯M\underline{\Pi}_{\mathrm{M}} satisfies the Hodge–Riemann relations with respect to \ell if and only if, for all k<d2k<\frac{d}{2}, the Hodge–Riemann form HR¯k(M)\underline{\mathrm{HR}}_{\ell}^{k}(\mathrm{M}) is non-degenerate and has the signature

j=0k(1)kj(dimCH¯j(M)dimCH¯j1(M)).\sum_{j=0}^{k}(-1)^{k-j}\Big{(}\operatorname{dim}\underline{\mathrm{CH}}^{j}(\mathrm{M})-\operatorname{dim}\underline{\mathrm{CH}}^{j-1}(\mathrm{M})\Big{)}.

Since Π¯M\underline{\Pi}_{\mathrm{M}} satisfies the hard Lefschetz theorem with respect to any strictly convex piecewise linear function on Π¯B\underline{\Pi}_{\mathrm{B}} and signature is a locally constant function on the space of nonsingular forms, the following statements are equivalent:

  1. (i)

    The fan Π¯M\underline{\Pi}_{\mathrm{M}} satisfies the Hodge–Riemann relations with respect to any strictly convex piecewise linear function on Π¯B\underline{\Pi}_{\mathrm{B}}.

  2. (ii)

    The fan Π¯M\underline{\Pi}_{\mathrm{M}} satisfies the Hodge–Riemann relations with respect to some strictly convex piecewise linear function on Π¯B\underline{\Pi}_{\mathrm{B}}.

Furthermore, since satisfying the Hodge–Riemann relations with respect to a given piecewise linear function is an open condition on the function, statement (ii) is equivalent to the following:

  1. (iii)

    The fan Π¯M\underline{\Pi}_{\mathrm{M}} satisfies the Hodge–Riemann relations with respect to some convex piecewise linear function on Π¯B\underline{\Pi}_{\mathrm{B}}.

We show that statement (iii) holds using the semi-small decomposition in Theorem 1.1.

If M\mathrm{M} is the Boolean matroid B\mathrm{B}, then CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}) can be identified with the cohomology ring of the smooth complex projective toric variety XΠ¯BX_{\underline{\Pi}_{\mathrm{B}}}. Therefore, in this case, Theorem 1.3 is a special case of the usual hard Lefschetz theorem and the Hodge–Riemann relations for smooth complex projective varieties.161616It is not difficult to directly prove the hard Lefschetz theorem and the Hodge–Riemann relations for CH¯(B)\underline{\mathrm{CH}}(\mathrm{B}) using the coloop case of Theorem 1.1. Alternatively, we may apply McMullen’s hard Lefschetz theorem and Hodge–Riemann relations for polytope algebras [McMullen] to the standard permutohedron in E\mathbb{R}^{E}.

If M\mathrm{M} is not the Boolean matroid B\mathrm{B}, choose an element ii that is not a coloop in M\mathrm{M}, and consider the morphism of fans

π¯i:Π¯MΠ¯Mi.\underline{\pi}_{i}:\underline{\Pi}_{\mathrm{M}}\longrightarrow\underline{\Pi}_{\mathrm{M}\setminus i}.

By induction, we know that Π¯Mi\underline{\Pi}_{\mathrm{M}\setminus i} satisfies the Hodge–Riemann relations with respect to any strictly convex piecewise linear function \ell on Π¯Bi\underline{\Pi}_{\mathrm{B}\setminus i}. We will show that Π¯M\underline{\Pi}_{\mathrm{M}} satisfies the Hodge–Riemann relations with respect to the pullback i:-πi\ell_{i}\coloneq\ell\circ\pi_{i}, which is a piecewise linear function on Π¯B\underline{\Pi}_{\mathrm{B}} that is convex but not necessarily strictly convex.

By Theorem 1.1, we have the orthogonal decomposition of CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}) into CH¯(Mi)\underline{\mathrm{CH}}(\mathrm{M}\setminus i)-modules

CH¯(M)=CH¯(i)F𝒮¯ixFiCH¯(i).\underline{\mathrm{CH}}(\mathrm{M})=\underline{\mathrm{CH}}_{(i)}\oplus\bigoplus_{F\in\underline{\mathscr{S}}_{i}}x_{F\cup i}\underline{\mathrm{CH}}_{(i)}.

By orthogonality, it is enough to show that each summand of CH¯(M)\underline{\mathrm{CH}}(\mathrm{M}) satisfies the Hodge–Riemann relations with respect to i\ell_{i}:

  1. (iv)

    For every nonnegative integer k<d2k<\frac{d}{2}, the bilinear form

    CH¯k(i)×CH¯k(i),(η1,η2)(1)kdeg¯M(id2k1η1η2)\underline{\mathrm{CH}}^{k}_{(i)}\times\underline{\mathrm{CH}}^{k}_{(i)}\longrightarrow\mathbb{Q},\qquad(\eta_{1},\eta_{2})\longmapsto(-1)^{k}\underline{\deg}_{\mathrm{M}}(\ell_{i}^{d-2k-1}\eta_{1}\eta_{2})

    is positive definite on the kernel of multiplication by id2k\ell_{i}^{d-2k}.

  2. (v)

    For every nonnegative integer k<d2k<\frac{d}{2}, the bilinear form

    xFiCH¯k1(i)×xFiCH¯k1(i),(η1,η2)(1)kdeg¯M(id2k1η1η2)x_{F\cup i}\underline{\mathrm{CH}}^{k-1}_{(i)}\times x_{F\cup i}\underline{\mathrm{CH}}^{k-1}_{(i)}\longrightarrow\mathbb{Q},\qquad(\eta_{1},\eta_{2})\longmapsto(-1)^{k}\underline{\deg}_{\mathrm{M}}(\ell_{i}^{d-2k-1}\eta_{1}\eta_{2})

    is positive definite on the kernel of multiplication by id2k\ell_{i}^{d-2k}.

By Proposition 3.4, the homomorphism θ¯i\underline{\theta}_{i} restricts to an isomorphism of CH¯(Mi)\underline{\mathrm{CH}}(\mathrm{M}\setminus i)-modules

CH¯(Mi)CH¯(i).\underline{\mathrm{CH}}(\mathrm{M}\setminus i)\cong\underline{\mathrm{CH}}_{(i)}.

Thus, statement (iv) follows from Lemma 3.3 and the induction hypothesis applied to Mi\mathrm{M}\setminus i. By Propositions 2.22, 3.4, and 3.5, the homomorphisms θ¯iFi\underline{\theta}_{i}^{F\cup i} and ψ¯FiM\underline{\psi}^{F\cup i}_{\mathrm{M}} give a CH¯(Mi)\underline{\mathrm{CH}}(\mathrm{M}\setminus i)-module isomorphism

CH¯(MFi)CH¯(MF)CH¯(MFi)θ¯iFiCH¯(MF)xFiCH¯(i)[1].\underline{\mathrm{CH}}(\mathrm{M}_{F\cup i})\otimes\underline{\mathrm{CH}}(\mathrm{M}^{F})\cong\underline{\mathrm{CH}}(\mathrm{M}_{F\cup i})\otimes\underline{\theta}_{i}^{F\cup i}\underline{\mathrm{CH}}(\mathrm{M}^{F})\cong x_{F\cup i}\underline{\mathrm{CH}}_{(i)}[1].

Note that the pullback of a strictly convex piecewise linear function on Π¯Bi\underline{\Pi}_{\mathrm{B}\setminus i} to the star

Π¯(Bi)F×Π¯(Bi)F=Π¯BFi×Π¯BF\underline{\Pi}_{(\mathrm{B}\setminus i)_{F}}\times\underline{\Pi}_{(\mathrm{B}\setminus i)^{F}}=\underline{\Pi}_{\mathrm{B}_{F\cup i}}\times\underline{\Pi}_{\mathrm{B}^{F}}

is the class of a strictly convex piecewise linear function. Therefore, statement (v) follows from Proposition 3.7 and the induction applied to MFi\mathrm{M}_{F\cup i} and MF\mathrm{M}^{F}. ∎

Proof of Theorem 1.3, parts (5) and (6).

This proof is nearly identical to the proof of parts (2) and (3). In that argument, we used the fact that rays of Π¯M\underline{\Pi}_{\mathrm{M}} are indexed by nonempty proper flats of M\mathrm{M} and the star of the ray ρ¯F\underline{\rho}_{F} is isomorphic to Π¯MF×Π¯MF\underline{\Pi}_{\mathrm{M}_{F}}\times\underline{\Pi}_{\mathrm{M}^{F}}, which we can show satisfies the hard Lefschetz theorem and the Hodge–Riemann relations using the induction hypothesis. When dealing instead with the augmented Bergman fan ΠM\Pi_{\mathrm{M}}, we have rays indexed by elements of EE and rays indexed by proper flats of M\mathrm{M}, with

starρiΠMΠMcl(i)starρFΠMΠ¯MF×ΠMF.\text{star}_{\rho_{i}}\Pi_{\mathrm{M}}\cong\Pi_{\mathrm{M}_{\text{cl}(i)}}\text{star}_{\rho_{F}}\Pi_{\mathrm{M}}\cong\underline{\Pi}_{\mathrm{M}_{F}}\times\Pi_{\mathrm{M}^{F}}.

Thus the stars of ρi\rho_{i} and ρF\rho_{F} for nonempty FF can be shown to satisfy the hard Lefschetz theorem and the Hodge–Riemann relations using the induction hypothesis. However, the star of ρ\rho_{\varnothing} is isomorphic to Π¯M\underline{\Pi}_{\mathrm{M}}, so we need to use parts (2) and (3) of Theorem 1.3 for M\mathrm{M} itself. ∎

Remark 4.1.

It is possible to deduce Poincaré duality, the hard Lefschetz theorem, and the Hodge–Riemann relations for CH(M)\mathrm{CH}(\mathrm{M}) using [AHK, Theorem 6.19 and Theorem 8.8], where the three properties are proved for generalized Bergman fans ΣN,𝒫\Sigma_{\mathrm{N},\mathscr{P}} in [AHK, Definition 3.2]. We sketch the argument here, leaving details to the interested readers. Consider the direct sum M0\mathrm{M}\oplus 0 of M\mathrm{M} and the rank 11 matroid on the singleton {0}\{0\} and the order filter 𝒫(M)\mathscr{P}(\mathrm{M}) of all proper flats of M0\mathrm{M}\oplus 0 that contain 0. The symbols B0\mathrm{B}\oplus 0 and 𝒫(B)\mathscr{P}(\mathrm{B}) are defined in the same way for the Boolean matroid B\mathrm{B} on EE. It is straightforward to check that the linear isomorphism

EE0/𝐞E+𝐞0,𝐞j𝐞j\mathbb{R}^{E}\longrightarrow\mathbb{R}^{E\cup 0}/\langle\mathbf{e}_{E}+\mathbf{e}_{0}\rangle,\quad\mathbf{e}_{j}\longmapsto\mathbf{e}_{j}

identifies the complete fan ΠB\Pi_{\mathrm{B}} with the complete fan ΣB0,𝒫(B)\Sigma_{\mathrm{B}\hskip 0.85358pt\oplus\hskip 0.85358pt0,\mathscr{P}(\mathrm{B})}, and the augmented Bergman fan ΠM\Pi_{\mathrm{M}} with a subfan of ΣM0,𝒫(M)\Sigma_{\mathrm{M}\hskip 0.85358pt\oplus\hskip 0.85358pt0,\mathscr{P}(\mathrm{M})}. The third identity in Lemma 2.9 shows that the inclusion of the augmented Bergman fan ΠM\Pi_{\mathrm{M}} into the generalized Bergman fan ΣM0,𝒫(M)\Sigma_{\mathrm{M}\hskip 0.85358pt\oplus\hskip 0.85358pt0,\mathscr{P}(\mathrm{M})} induces an isomorphism between their Chow rings.

5. Proof of Theorem 1.4

In this section, we prove the decomposition (D¯3\underline{\mathrm{D}}_{3}) by induction on the cardinality of EE. The decomposition (D3\mathrm{D}_{3}) can be proved using the same argument. The results are trivial when EE has at most one element. Thus, we may assume that ii is an element of EE, that EiE\setminus i is nonempty, and that all the results hold for loopless matroids whose ground set is a proper subset of EE.

We first prove that the summands appearing in the right-hand side of (D¯3\underline{\mathrm{D}}_{3}) are orthogonal to each other.

Lemma 5.1.

Let FF and GG be distinct nonempty proper flats of M\mathrm{M}.

  1. (1)

    The spaces ψ¯FMCH¯(MF)J¯α¯(MF)\underline{\psi}^{F}_{\mathrm{M}}\ \underline{\operatorname{CH}}(\mathrm{M}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F}) and H¯α¯(M)\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M}) are orthogonal in CH¯(M)\underline{\operatorname{CH}}(\mathrm{M}).

  2. (2)

    The spaces ψ¯FMCH¯(MF)J¯α¯(MF)\underline{\psi}^{F}_{\mathrm{M}}\underline{\operatorname{CH}}(\mathrm{M}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F}) and ψ¯GMCH¯(MG)J¯α¯(MG)\underline{\psi}^{G}_{\mathrm{M}}\underline{\operatorname{CH}}(\mathrm{M}_{G})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{G}) are orthogonal in CH¯(M)\underline{\operatorname{CH}}(\mathrm{M}).

Proof.

The fifth bullet point in Proposition 2.20, together with the fact that ψ¯MF\underline{\psi}_{\mathrm{M}}^{F} is a CH¯(M)\underline{\operatorname{CH}}(\mathrm{M})-module homomorphism via φ¯MF\underline{\varphi}_{\mathrm{M}}^{F}, implies that every summand in the right-hand side of (D¯3\underline{\mathrm{D}}_{3}) is an H¯α¯(M)\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M})-submodule. Thus the first orthogonality follows from the vanishing of ψ¯FMCH¯(MF)J¯α¯(MF)\underline{\psi}^{F}_{\mathrm{M}}\ \underline{\operatorname{CH}}(\mathrm{M}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F}) in degree d1d-1.

For the second orthogonality, we may suppose that FF is a proper subset of GG. Since ψ¯GM\underline{\psi}^{G}_{\mathrm{M}} is a CH¯(M)\underline{\operatorname{CH}}(\mathrm{M})-module homomorphism commuting with the degree maps, it is enough to show that

φ¯GMψ¯FMCH¯(MF)J¯α¯(MF)\underline{\varphi}^{G}_{\mathrm{M}}\underline{\psi}^{F}_{\mathrm{M}}\underline{\operatorname{CH}}(\mathrm{M}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F}) and CH¯(MG)J¯α¯(MG)\underline{\operatorname{CH}}(\mathrm{M}_{G})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{G}) are orthogonal in CH¯(MG)CH¯(MG)\underline{\operatorname{CH}}(\mathrm{M}_{G})\otimes\underline{\operatorname{CH}}(\mathrm{M}^{G}).

For this, we use the commutative diagram of pullback and pushforward maps

CH¯(MF)CH¯(MF)\textstyle{\underline{\operatorname{CH}}(\mathrm{M}_{F})\otimes\underline{\operatorname{CH}}(\mathrm{M}^{F})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ¯FM\scriptstyle{\underline{\psi}^{F}_{\mathrm{M}}}φ¯GFMF 1\scriptstyle{\underline{\varphi}^{G\setminus F}_{\mathrm{M}_{F}}\otimes\ 1}CH¯(M)\textstyle{\underline{\operatorname{CH}}(\mathrm{M})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ¯GM\scriptstyle{\underline{\varphi}^{G}_{\mathrm{M}}}CH¯(MG)CH¯(MGF)CH¯(MF)\textstyle{\underline{\operatorname{CH}}(\mathrm{M}_{G})\otimes\underline{\operatorname{CH}}(\mathrm{M}^{G}_{F})\otimes\underline{\operatorname{CH}}(\mathrm{M}^{F})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1ψ¯FMG\scriptstyle{\quad 1\ \otimes\ \underline{\psi}^{F}_{\mathrm{M}^{G}}}CH¯(MG)CH¯(MG),\textstyle{\underline{\operatorname{CH}}(\mathrm{M}_{G})\otimes\underline{\operatorname{CH}}(\mathrm{M}^{G}),}

which further reduces to the assertion that

ψ¯FMGCH¯(MFG)J¯α¯(MF)\underline{\psi}^{F}_{\mathrm{M}^{G}}\underline{\operatorname{CH}}(\mathrm{M}_{F}^{G})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F}) and J¯α¯(MG)\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{G}) are orthogonal in CH¯(MG)\underline{\operatorname{CH}}(\mathrm{M}^{G}).

Since J¯α¯(MG)H¯α¯(MG)\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{G})\subseteq\underline{\operatorname{H}}_{\underline{\alpha}}(\mathrm{M}^{G}), the above follows from the first orthogonality for MG\mathrm{M}^{G}. ∎

We next show that the restriction of the Poincaré pairing of CH¯(M)\underline{\operatorname{CH}}(\mathrm{M}) to each summand appearing in the right-hand side of (D¯3\underline{\mathrm{D}}_{3}) is non-degenerate.

Lemma 5.2.

Let FF be a nonempty proper flat of M\mathrm{M}, and let k=rkM(F)k=\text{rk}_{\mathrm{M}}(F).

  1. (1)

    The restriction of the Poincaré pairing of CH¯(M)\underline{\operatorname{CH}}(\mathrm{M}) to H¯α¯(M)\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M}) is non-degenerate.

  2. (2)

    The restriction of the Poincaré pairing of CH¯(M)\underline{\operatorname{CH}}(\mathrm{M}) to ψ¯FMCH¯(MF)J¯α¯(MF)\underline{\psi}^{F}_{\mathrm{M}}\underline{\operatorname{CH}}(\mathrm{M}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F}) is non-degenerate.

Proof.

The first statement follows from Proposition 2.26. We prove the second statement.

Since the Poincaré pairing of CH¯(MF)\underline{\operatorname{CH}}(\mathrm{M}_{F}) is non-degenerate, it is enough to show that the restriction of the Poincaré pairing satisfies

deg¯M(ψ¯FM(μ1ν1)ψ¯FM(μ2ν2))=deg¯MF(μ1μ2)deg¯MF(α¯MFν1ν2).\underline{\deg}_{\mathrm{M}}\big{(}\underline{\psi}^{F}_{\mathrm{M}}(\mu_{1}\otimes\nu_{1})\cdot\underline{\psi}^{F}_{\mathrm{M}}(\mu_{2}\otimes\nu_{2})\big{)}=-\underline{\deg}_{\mathrm{M}_{F}}(\mu_{1}\mu_{2})\ \underline{\deg}_{\mathrm{M}^{F}}(\underline{\alpha}_{\mathrm{M}^{F}}\nu_{1}\nu_{2}).

The proof of the identity is nearly identical to that of Proposition 3.7. The left-hand side is

deg¯M(ψ¯FM(φ¯FMψ¯FM(μ1ν1)(μ2ν2)))=deg¯MFdeg¯MF(φ¯FMψ¯FM(μ1ν1)(μ2ν2))\underline{\deg}_{\mathrm{M}}\big{(}\underline{\psi}^{F}_{\mathrm{M}}\big{(}\underline{\varphi}^{F}_{\mathrm{M}}\underline{\psi}^{F}_{\mathrm{M}}(\mu_{1}\otimes\nu_{1})\cdot(\mu_{2}\otimes\nu_{2})\big{)}\big{)}=\underline{\deg}_{\mathrm{M}_{F}}\otimes\underline{\deg}_{\mathrm{M}^{F}}\big{(}\underline{\varphi}^{F}_{\mathrm{M}}\underline{\psi}^{F}_{\mathrm{M}}(\mu_{1}\otimes\nu_{1})\cdot(\mu_{2}\otimes\nu_{2})\big{)}

because ψ¯FM\underline{\psi}^{F}_{\mathrm{M}} is a CH¯(M)\underline{\operatorname{CH}}(\mathrm{M})-module homomorphism commuting with the degree maps. Since the composition φ¯FMψ¯FM\underline{\varphi}^{F}_{\mathrm{M}}\underline{\psi}^{F}_{\mathrm{M}} is multiplication by φ¯FM(xF)\underline{\varphi}^{F}_{\mathrm{M}}(x_{F}), the above becomes

deg¯MFdeg¯MF((1α¯MF+β¯MF1)(μ1ν1)(μ2ν2)).-\underline{\deg}_{\mathrm{M}_{F}}\otimes\underline{\deg}_{\mathrm{M}^{F}}\big{(}(1\otimes\underline{\alpha}_{\mathrm{M}^{F}}+\underline{\beta}_{\mathrm{M}_{F}}\otimes 1)\cdot(\mu_{1}\otimes\nu_{1})\cdot(\mu_{2}\otimes\nu_{2})\big{)}.

The vanishing of J¯α¯(MF)\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F}) in degree k1k-1 further simplifies the above to the desired expression

deg¯MFdeg¯MF((1α¯MF)(μ1ν1)(μ2ν2))=deg¯MF(μ1μ2)deg¯MF(α¯MFν1ν2).-\underline{\deg}_{\mathrm{M}_{F}}\otimes\underline{\deg}_{\mathrm{M}^{F}}\big{(}(1\otimes\underline{\alpha}_{\mathrm{M}^{F}})\cdot(\mu_{1}\otimes\nu_{1})\cdot(\mu_{2}\otimes\nu_{2})\big{)}=-\underline{\deg}_{\mathrm{M}_{F}}(\mu_{1}\mu_{2})\ \underline{\deg}_{\mathrm{M}^{F}}(\underline{\alpha}_{\mathrm{M}^{F}}\nu_{1}\nu_{2}).\qed

To complete the proof, we only need to show that the graded vector spaces on both sides of (D¯3\underline{\mathrm{D}}_{3}) have the same dimension, which is the next proposition.

Proposition 5.3.

As graded vector spaces, there exists an isomorphism

CH¯(M)H¯α¯(M)F𝒞¯(M)CH¯(MF)J¯α¯(MF)[1],\underline{\operatorname{CH}}(\mathrm{M})\cong\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M})\oplus\ \bigoplus_{F\in\underline{\mathscr{C}}(\mathrm{M})}\ \underline{\operatorname{CH}}(\mathrm{M}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F})[-1], (D¯3\underline{\mathrm{D}}_{3}^{\prime})

where the sum is over the set 𝒞¯(M)\underline{\mathscr{C}}(\mathrm{M}) of proper flats of M\mathrm{M} with rank at least two.

Proof.

We prove the proposition using induction on the cardinality of EE. Suppose the proposition holds for any matroid whose ground set is a proper subset of EE. Suppose that there exists an element iEi\in E that is not a coloop. Then the decomposition (D¯1\underline{\mathrm{D}}_{1}) implies

CH¯(M)CH¯(Mi)G𝒮¯i(M)CH¯(MGi)CH¯(MG)[1],\underline{\operatorname{CH}}(\mathrm{M})\cong\underline{\operatorname{CH}}(\mathrm{M}\setminus i)\oplus\bigoplus_{G\in\underline{\mathscr{S}}_{i}(\mathrm{M})}\underline{\mathrm{CH}}(\mathrm{M}_{G\cup i})\otimes\underline{\mathrm{CH}}(\mathrm{M}^{G})[-1],

since the maps θ¯i\underline{\theta}_{i}, θ¯iGi\underline{\theta}_{i}^{G\cup i}, and ψ¯MGi\underline{\psi}_{\mathrm{M}}^{G\cup i} are injective via the Poincaré duality part of Theorem 1.3. By applying the induction hypothesis to the matroids Mi\mathrm{M}\setminus i and MG\mathrm{M}^{G}, we see that the left-hand side of (D¯3\underline{\mathrm{D}}_{3}^{\prime}) is isomorphic to the graded vector space

H¯α¯(Mi)\displaystyle\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M}\setminus i) G𝒞¯(Mi)CH¯((Mi)G)J¯α¯((Mi)G)[1]\displaystyle\oplus\bigoplus_{G\in\underline{\mathscr{C}}({\mathrm{M}\setminus i})}\underline{\operatorname{CH}}\big{(}(\mathrm{M}\setminus i)_{G}\big{)}\otimes\underline{\operatorname{J}}_{\underline{\alpha}}\big{(}(\mathrm{M}\setminus i)^{G}\big{)}[-1]
G𝒮¯i(M)CH¯(MGi)H¯α¯(MG)[1]\displaystyle\oplus\bigoplus_{G\in\underline{\mathscr{S}}_{i}({\mathrm{M}})}\ \underline{\mathrm{CH}}(\mathrm{M}_{G\cup i})\otimes\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M}^{G})[-1]
G𝒮¯i(M)F𝒞¯(MG)CH¯(MGi)CH¯(MGF)J¯α¯(MF)[2].\displaystyle\oplus\bigoplus_{G\in\underline{\mathscr{S}}_{i}({\mathrm{M}})}\bigoplus_{F\in\underline{\mathscr{C}}({\mathrm{M}^{G}})}\underline{\mathrm{CH}}(\mathrm{M}_{G\cup i})\otimes\underline{\operatorname{CH}}(\mathrm{M}^{G}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F})[-2].

Since ii is not a coloop, we may replace H¯α¯(Mi)\underline{\operatorname{H}}_{\underline{\alpha}}(\mathrm{M}\setminus i) by H¯α¯(M)\underline{\operatorname{H}}_{\underline{\alpha}}(\mathrm{M}).

Now, we further decompose the right-hand side of (D¯3\underline{\mathrm{D}}_{3}^{\prime}) to match the displayed expression. For this, we split the index set 𝒞¯(M)\underline{\mathscr{C}}(\mathrm{M}) into three groups:

  1. (1)

    F𝒞¯(M),iF,Fi𝒮¯i(M)F\in\underline{\mathscr{C}}({\mathrm{M}}),i\in F,F\setminus i\in\underline{\mathscr{S}}_{i}({\mathrm{M}}),

  2. (2)

    F𝒞¯(M),iF,Fi𝒮¯i(M)F\in\underline{\mathscr{C}}({\mathrm{M}}),i\in F,F\setminus i\notin\underline{\mathscr{S}}_{i}({\mathrm{M}}), and

  3. (3)

    F𝒞¯(M),iFF\in\underline{\mathscr{C}}({\mathrm{M}}),i\notin F.

Suppose FF belongs to the first group. In this case, we have J¯α¯(MF)H¯α¯(MFi)\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F})\cong\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M}^{F\setminus i}) as graded vector spaces. Therefore, we have

F𝒞¯(M)iF,Fi𝒮¯i(M)CH¯(MF)J¯α¯(MF)[1]G𝒮¯i(M)CH¯(MGi)H¯α¯(MG)[1].\bigoplus_{\begin{subarray}{c}F\in\underline{\mathscr{C}}(\mathrm{M})\\ i\in F,F\setminus i\in\underline{\mathscr{S}}_{i}(\mathrm{M})\end{subarray}}\underline{\mathrm{CH}}(\mathrm{M}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F})[-1]\cong\bigoplus_{G\in\underline{\mathscr{S}}_{i}(\mathrm{M})}\underline{\mathrm{CH}}(\mathrm{M}_{G\cup i})\otimes\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M}^{G})[-1].

Suppose FF belongs to the second group. In this case, MF=(Mi)Fi\mathrm{M}_{F}=(\mathrm{M}\setminus i)_{F\setminus i}, and the matroids MF\mathrm{M}^{F} and (Mi)Fi(\mathrm{M}\setminus i)^{F\setminus i} have the same rank. Therefore, we have

F𝒞¯(M)iF,Fi𝒮¯i(M)CH¯(MF)J¯α¯(MF)[1]G𝒞¯(Mi)𝒞¯(M)CH¯((Mi)G)J¯α¯((Mi)G)[1].\bigoplus_{\begin{subarray}{c}F\in\underline{\mathscr{C}}(\mathrm{M})\\ i\in F,F\setminus i\notin\underline{\mathscr{S}}_{i}(\mathrm{M})\end{subarray}}\underline{\mathrm{CH}}(\mathrm{M}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F})[-1]\cong\bigoplus_{G\in\underline{\mathscr{C}}({\mathrm{M}\setminus i})\setminus\underline{\mathscr{C}}(\mathrm{M})}\ \underline{\operatorname{CH}}\big{(}(\mathrm{M}\setminus i)_{G}\big{)}\otimes\underline{\operatorname{J}}_{\underline{\alpha}}\big{(}(\mathrm{M}\setminus i)^{G}\big{)}[-1].

Suppose FF belongs to the third group. In this case, we apply (D¯1\underline{\mathrm{D}}_{1}) to MF\mathrm{M}_{F} and get

F𝒞¯(M),iFCH¯(MF)J¯α¯(MF)[1]\displaystyle\bigoplus_{F\in\underline{\mathscr{C}}(\mathrm{M}),i\notin F}\underline{\operatorname{CH}}(\mathrm{M}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F})[-1]
\displaystyle\cong F𝒞¯(M),iF(CH¯(MFi)G𝒮¯i(MF)CH¯(MGi)CH¯(MGF)[1])J¯α¯(MF)[1]\displaystyle\bigoplus_{F\in\underline{\mathscr{C}}(\mathrm{M}),i\notin F}\Big{(}\underline{\operatorname{CH}}\big{(}\mathrm{M}_{F}\setminus i\big{)}\oplus\bigoplus_{G\in\underline{\mathscr{S}}_{i}(\mathrm{M}_{F})}\underline{\operatorname{CH}}(\mathrm{M}_{G\cup i})\otimes\underline{\operatorname{CH}}(\mathrm{M}^{G}_{F})[-1]\Big{)}\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F})[-1]
\displaystyle\cong F𝒞¯(M),iFCH¯(MFi)J¯α¯(MF)[1]G𝒮¯i(M)F𝒞¯(MG)CH¯(MGi)CH¯(MGF)J¯α¯(MF)[2]\displaystyle\bigoplus_{F\in\underline{\mathscr{C}}(\mathrm{M}),i\notin F}\underline{\operatorname{CH}}\big{(}\mathrm{M}_{F}\setminus i\big{)}\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F})[-1]\oplus\bigoplus_{\begin{subarray}{c}G\in\underline{\mathscr{S}}_{i}(\mathrm{M})\\ F\in\underline{\mathscr{C}}({\mathrm{M}^{G}})\end{subarray}}\underline{\operatorname{CH}}(\mathrm{M}_{G\cup i})\otimes\underline{\operatorname{CH}}(\mathrm{M}^{G}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F})[-2]
\displaystyle\cong G𝒞¯(Mi)𝒞¯(M)CH¯((Mi)G)J¯α¯((Mi)G)[1]G𝒮¯i(M)F𝒞¯(MG)CH¯(MGi)CH¯(MGF)J¯α¯(MF)[2].\displaystyle\bigoplus_{G\in\underline{\mathscr{C}}({\mathrm{M}\setminus i})\cap\underline{\mathscr{C}}({\mathrm{M}})}\underline{\operatorname{CH}}\big{(}(\mathrm{M}\setminus i)_{G}\big{)}\otimes\underline{\operatorname{J}}_{\underline{\alpha}}\big{(}(\mathrm{M}\setminus i)^{G}\big{)}[-1]\oplus\bigoplus_{\begin{subarray}{c}G\in\underline{\mathscr{S}}_{i}(\mathrm{M})\\ F\in\underline{\mathscr{C}}({\mathrm{M}^{G}})\end{subarray}}\underline{\operatorname{CH}}(\mathrm{M}_{G\cup i})\otimes\underline{\operatorname{CH}}(\mathrm{M}^{G}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F})[-2].

The decomposition (D¯3\underline{\mathrm{D}}_{3}^{\prime}) follows.

Suppose now that every element of EE is a coloop of M\mathrm{M}; that is, M\mathrm{M} is a Boolean matroid. We fix an element iEi\in E. The decomposition (D¯2\underline{\mathrm{D}}_{2}) and the Poincaré duality part of Theorem 1.3 imply

CH¯(M)CH¯(Mi)CH¯(Mi)[1]G𝒮¯i(M)CH¯(MGi)CH¯(MG)[1].\underline{\operatorname{CH}}(\mathrm{M})\cong\underline{\operatorname{CH}}(\mathrm{M}\setminus i)\oplus\underline{\operatorname{CH}}(\mathrm{M}\setminus i)[-1]\oplus\bigoplus_{G\in\underline{\mathscr{S}}_{i}(\mathrm{M})}\underline{\mathrm{CH}}(\mathrm{M}_{G\cup i})\otimes\underline{\mathrm{CH}}(\mathrm{M}^{G})[-1].

The assumption that ii is a coloop implies that 𝒮¯i(M)𝒞¯(M)=𝒞¯(Mi)\underline{\mathscr{S}}_{i}(\mathrm{M})\cap\underline{\mathscr{C}}(\mathrm{M})=\underline{\mathscr{C}}(\mathrm{M}\setminus i). The induction hypothesis applies to the matroids Mi\mathrm{M}\setminus i and MG\mathrm{M}^{G}, and hence the left-hand side of (D¯3\underline{\mathrm{D}}_{3}^{\prime}) is isomorphic to

H¯α¯(Mi)G𝒞¯(Mi)CH¯((Mi)G)J¯α¯((Mi)G)[1]\displaystyle\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M}\setminus i)\oplus\ \bigoplus_{G\in\underline{\mathscr{C}}(\mathrm{M}\setminus i)}\ \underline{\operatorname{CH}}\big{(}(\mathrm{M}\setminus i)_{G}\big{)}\otimes\underline{\operatorname{J}}_{\underline{\alpha}}\big{(}(\mathrm{M}\setminus i)^{G}\big{)}[-1]
H¯α¯(Mi)[1]G𝒞¯(Mi)CH¯((Mi)G)J¯α¯((Mi)G)[2]\displaystyle\oplus\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M}\setminus i)[-1]\oplus\ \bigoplus_{G\in\underline{\mathscr{C}}(\mathrm{M}\setminus i)}\ \underline{\operatorname{CH}}\big{(}(\mathrm{M}\setminus i)_{G}\big{)}\otimes\underline{\operatorname{J}}_{\underline{\alpha}}\big{(}(\mathrm{M}\setminus i)^{G}\big{)}[-2]
G𝒮¯i(M)CH¯(MGi)(H¯α¯(MG)F𝒞¯(MG)CH¯(MGF)J¯α¯(MF)[1])[1].\displaystyle\oplus\bigoplus_{G\in\underline{\mathscr{S}}_{i}(\mathrm{M})}\underline{\mathrm{CH}}(\mathrm{M}_{G\cup i})\otimes\Big{(}\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M}^{G})\oplus\ \bigoplus_{F\in\underline{\mathscr{C}}(\mathrm{M}^{G})}\ \underline{\operatorname{CH}}(\mathrm{M}^{G}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F})[-1]\Big{)}[-1].

Now, we further decompose the right-hand side of (D¯3\underline{\mathrm{D}}_{3}^{\prime}) to match the displayed expression. For this, we split the index set 𝒞¯(M)\underline{\mathscr{C}}(\mathrm{M}) into three groups:

  1. (1)

    F𝒞¯(M),iFF\in\underline{\mathscr{C}}({\mathrm{M}}),i\in F,

  2. (2)

    F𝒞¯(M),F=EiF\in\underline{\mathscr{C}}({\mathrm{M}}),F=E\setminus i, and

  3. (3)

    F𝒞¯(M),F𝒮¯i(M)F\in\underline{\mathscr{C}}({\mathrm{M}}),F\in\underline{\mathscr{S}}_{i}({\mathrm{M}}).

If FF belongs to the first group, then J¯α¯(MF)H¯α¯(MFi)\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F})\cong\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M}^{F\setminus i}), and hence

F𝒞¯(M),iFCH¯(MF)J¯α¯(MF)[1]G𝒮¯i(M)CH¯(MGi)H¯α¯(MG)[1].\bigoplus_{F\in\underline{\mathscr{C}}(\mathrm{M}),i\in F}\ \underline{\operatorname{CH}}(\mathrm{M}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F})[-1]\cong\bigoplus_{G\in\underline{\mathscr{S}}_{i}(\mathrm{M})}\underline{\mathrm{CH}}(\mathrm{M}_{G\cup i})\otimes\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M}^{G})[-1].

If FF is the flat EiE\setminus i, we have

H¯α¯(M)CH¯(MEi)J¯α¯(MEi)[1]H¯α¯(Mi)H¯α¯(Mi)[1].\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M})\oplus\underline{\operatorname{CH}}(\mathrm{M}_{E\setminus i})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{E\setminus i})[-1]\cong\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M}\setminus i)\oplus\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M}\setminus i)[-1].

If FF belongs to the third group, we apply (D¯2\underline{\mathrm{D}}_{2}) to MF\mathrm{M}_{F} and get

F𝒞¯(M)F𝒮¯i(M)CH¯(MF)J¯α¯(MF)[1]\displaystyle\bigoplus_{\begin{subarray}{c}F\in\underline{\mathscr{C}}(\mathrm{M})\\ F\in\underline{\mathscr{S}}_{i}(\mathrm{M})\end{subarray}}\ \underline{\operatorname{CH}}(\mathrm{M}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F})[-1]
\displaystyle\cong F𝒞¯(M)F𝒮¯i(M)(CH¯(MFi)CH¯(MFi)[1]G𝒮¯i(MF)CH¯(MGi)CH¯(MGF)[1])J¯α¯(MF)[1]\displaystyle\bigoplus_{\begin{subarray}{c}F\in\underline{\mathscr{C}}(\mathrm{M})\\ F\in\underline{\mathscr{S}}_{i}(\mathrm{M})\end{subarray}}\ \Big{(}\underline{\operatorname{CH}}(\mathrm{M}_{F}\setminus i)\oplus\underline{\operatorname{CH}}(\mathrm{M}_{F}\setminus i)[-1]\oplus\bigoplus_{G\in\underline{\mathscr{S}}_{i}(\mathrm{M}_{F})}\underline{\operatorname{CH}}(\mathrm{M}_{G\cup i})\otimes\underline{\operatorname{CH}}(\mathrm{M}^{G}_{F})[-1]\Big{)}\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F})[-1]
\displaystyle\cong G𝒞¯(M)G𝒮¯i(M)CH¯(MGi)J¯α¯(MG)[1]G𝒞¯(M)G𝒮¯i(M)CH¯(MGi)J¯α¯(MG)[2]\displaystyle\bigoplus_{\begin{subarray}{c}G\in\underline{\mathscr{C}}(\mathrm{M})\\ G\in\underline{\mathscr{S}}_{i}(\mathrm{M})\end{subarray}}\ \underline{\operatorname{CH}}(\mathrm{M}_{G}\setminus i)\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{G})[-1]\oplus\bigoplus_{\begin{subarray}{c}G\in\underline{\mathscr{C}}(\mathrm{M})\\ G\in\underline{\mathscr{S}}_{i}(\mathrm{M})\end{subarray}}\ \underline{\operatorname{CH}}(\mathrm{M}_{G}\setminus i)\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{G})[-2]\ \oplus
G𝒮¯i(M)F𝒞¯(MG)CH¯(MGi)CH¯(MGF)J¯α¯(MF)[2].\displaystyle\bigoplus_{\begin{subarray}{c}G\in\underline{\mathscr{S}}_{i}(\mathrm{M})\\ F\in\underline{\mathscr{C}}(\mathrm{M}^{G})\end{subarray}}\underline{\operatorname{CH}}(\mathrm{M}_{G\cup i})\otimes\underline{\operatorname{CH}}(\mathrm{M}^{G}_{F})\otimes\underline{\operatorname{J}}_{\underline{\alpha}}(\mathrm{M}^{F})[-2].

The decomposition (D¯3\underline{\mathrm{D}}_{3}^{\prime}) follows. ∎

Remark 5.4.

The decomposition of graded vector spaces appearing in [AHK, Theorem 6.18] specializes to decompositions of CH¯(M)\underline{\operatorname{CH}}(\mathrm{M}) and of CH(M)\operatorname{CH}(\mathrm{M}), where the latter goes through Remark 4.1. At the level of Poincaré polynomials, these decompositions coincide with those of Theorem 1.4. However, the subspaces appearing in the decompositions are not the same. In particular, the decompositions in [AHK, Theorem 6.18] are not orthogonal, and they are not compatible with the H¯α¯(M)\underline{\mathrm{H}}_{\underline{\alpha}}(\mathrm{M})-module structure on CH¯(M)\underline{\operatorname{CH}}(\mathrm{M}) or the Hα(M)\mathrm{H}_{\alpha}(\mathrm{M})-module structure on CH(M)\operatorname{CH}(\mathrm{M}).

References