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

A basis for the cohomology of compact models of toric arrangements

Giovanni Gaiffi Dipartimento di Matematica, Università di Pisa Largo B. Pontecorvo, 5 56127 Pisa, Italy giovanni.gaiffi@unipi.it Oscar Papini Istituto di Scienza e Tecnologie dell’Informazione “A. Faedo” Consiglio Nazionale delle Ricerche Via G. Moruzzi, 1 56124 Pisa, Italy oscar.papini@isti.cnr.it  and  Viola Siconolfi Dipartimento di Matematica, Università di Pisa Largo B. Pontecorvo, 5 56127 Pisa, Italy viola.siconolfi@dm.unipi.it
Abstract.

In this paper we find monomial bases for the integer cohomology rings of compact wonderful models of toric arrangements. In the description of the monomials various combinatorial objects come into play: building sets, nested sets, and the fan of a suitable toric variety. We provide some examples computed via a SageMath program and then we focus on the case of the toric arrangements associated with root systems of type AA. Here the combinatorial description of our basis offers a geometrical point of view on the relation between some Eulerian statistics on the symmetric group.

Key words and phrases:
Toric arrangements, compact models, configuration spaces, Eulerian numbers
2010 Mathematics Subject Classification:
14N20, 05E16, 05C30
V.S. supported by PRIN 2017 ‘Moduli and Lie Theory’, Dipartimento di Matematica, Università di Pisa.

1. Introduction

Let TT be an nn-dimensional torus and let X(T)X^{*}(T) be its group of characters; it is a lattice of rank nn and, by choosing a basis, we have isomorphisms X(T)nX^{*}(T)\simeq\mathbb{Z}^{n} and T()nT\simeq(\mathbb{C}^{*})^{n}. Given an element χX(T)\chi\in X^{*}(T), the corresponding character on TT will be denoted by xχ:Tx_{\chi}\colon T\to\mathbb{C}^{*}.

Definition 1.1.

A layer in TT is a subvariety of TT of the form

𝒦(Γ,ϕ){tTxχ(t)=ϕ(χ) for all χΓ}\mathcal{K}(\Gamma,\phi)\coloneqq\{t\in T\mid x_{\chi}(t)=\phi(\chi)\text{ for all }\chi\in\Gamma\}

where Γ<X(T)\Gamma<X^{*}(T) is a split direct summand and ϕ:Γ\phi\colon\Gamma\to\mathbb{C}^{*} is a homomorphism. A toric arrangement 𝒜\mathcal{A} is a (finite) set of layers {𝒦1,,𝒦r}\{\mathcal{K}_{1},\dotsc,\mathcal{K}_{r}\} in TT. A toric arrangement is called divisorial if every layer has codimension 1.

In [11] it is shown how to construct projective wonderful models for the complement (𝒜)=Ti𝒦i\mathcal{M}(\mathcal{A})=T\setminus\bigcup_{i}\mathcal{K}_{i}. A projective wonderful model is a smooth projective variety containing (𝒜)\mathcal{M}(\mathcal{A}) as an open set and such that the complement of (𝒜)\mathcal{M}(\mathcal{A}) is a divisor with normal crossings and smooth irreducible components. In [12] the integer cohomology ring of these projective wonderful models was described by showing generators and relations.

In this paper we describe a basis for the integer cohomology modules. This description calls into play the relevant combinatorial objects that characterize the geometrical and topological properties of these models: the fan of a suitable toric variety, the building set associated to the arrangement and its nested sets.

The construction of projective models of toric arrangements is a further step in a rich theory that was originated by De Concini and Procesi in [16, 15], where they studied wonderful models for the complement of a subspace arrangement, providing both a projective and a non-projective version of their construction.

In some cases the toric and subspace constructions provide the same variety. This happens for instance when we deal with root (hyperplane or toric) arrangements of type AA. Therefore in this case we can compare the new basis of the cohomology described in this paper with the old one coming from the subspace construction. Part of the description of these bases is similar but there are differences, that will lead us to find a bijection between two families of graphs (labeled forests) and a geometric interpretation of the equidistribution of two statistics (des and lec) on the symmetric group.

Since both subspace and toric models are involved in our results, we start providing a sketch of the history of the theory of wonderful models from both points of view.

1.1. Some history of linear and toric wonderful models

The construction of wonderful models of subspace arrangements in [16, 15] was originally motivated by the study of Drinfeld’s construction in [24] of special solutions of the Knizhnik-Zamolodchikov equations with some prescribed asymptotic behavior, then it turned out that the role of these models is crucial in several areas of mathematics. For instance in the case of a complexified root arrangement of type AnA_{n} (which we will deal with in Section 5 of this paper) the minimal model coincides with the moduli spaces of stable curves of genus 0 with n+2n+2 marked points.

In the seminal papers of De Concini and Procesi the notions of building sets and nested sets appeared for the first time in a general version. In [16] the authors showed, using a description of the cohomology rings of the projective wonderful models to give an explicit presentation of a Morgan algebra, that the mixed Hodge numbers and the rational homotopy type of the complement of a complex subspace arrangement depend only on the intersection lattice (viewed as a ranked poset). The cohomology rings of the models of subspace arrangements were also studied in [49, 30], where some integer bases were provided, and, in the real case, in [26, 44]. The arrangements associated with complex reflection groups were dealt with in [33] from the representation theoretic point of view and in [6] from the homotopical point of view.

The connections between the geometry of these models and the Chow rings of matroids were pointed out first in [28] and then in [1], where they also played a crucial role in the study of some relevant log-concavity problems. The relations with toric and tropical geometry were enlightened for instance in [27, 20, 2].

The study of toric arrangements started in [35] and received then a new impulse from several points of view. In [19] and [18] the role of toric arrangements as a link between partition functions and box splines is pointed out; interesting enumerative and combinatorial aspects have been investigated via the Tutte polynomial and arithmetics matroids in [38, 39, 9]. As for the topology of the complement (𝒜)\mathcal{M}(\mathcal{A}) of a divisorial toric arrangement, the generators of the cohomology modules over \mathbb{C} where exhibited in [17] via local nonbroken circuits sets, and in the same paper the cohomology ring structure was determined in the case of totally unimodular arrangements. By a rather general approach, Dupont in [25] proved the rational formality of (𝒜)\mathcal{M}(\mathcal{A}). In turn, in [7], it was shown, extending the results in [4, 5] and [41], that the data needed in order to state the presentation of the rational cohomology ring of (𝒜)\mathcal{M}(\mathcal{A}) is fully encoded in the poset given by all the connected components of the intersections of the layers. It follows that in the divisorial case the combinatorics of this poset determines the rational homotopy of (𝒜)\mathcal{M}(\mathcal{A}).

One of the motivations for the construction of projective wonderful models of a toric arrangement 𝒜\mathcal{A} in [11], in addition to the interest in their own geometry, was that they could be an important tool to explore the generalization of the above mentioned results to the non-divisorial case.

Indeed the presentation of the cohomology ring of these models described in [12] was used in [40] to construct a Morgan differential algebra which determines the rational homotopy type of (𝒜)\mathcal{M}(\mathcal{A}). We notice that these models, and therefore their associated Morgan algebras, depend not only on the initial combinatorial data, but also on some choices (see Section 2 for more details). In [13] a new differential graded algebra was constructed as a direct limit of the above mentioned differential Morgan algebras: it is quasi isomorphic to any of the Morgan algebras of the projective wonderful models of (𝒜)\mathcal{M}(\mathcal{A}) and it has a presentation which depends only on a set of initial discrete data extracted from 𝒜\mathcal{A}, thus proving that in the non-divisorial case the rational homotopy type of (𝒜)\mathcal{M}(\mathcal{A}) depends only on these data.

As another application of the projective wonderful models of a toric arrangement, Denham and Suciu showed in [21] that (in the divisorial case) (𝒜)\mathcal{M}(\mathcal{A}) is both a duality space and an abelian duality space.

1.2. Structure of this paper

In Section 2 we will briefly recall from [11] the construction of the projective wonderful models associated with a toric arrangement. This is done in two steps: first one embeds the torus in a suitable smooth projective toric variety XΔX_{\Delta} with fan Δ\Delta, then one considers the arrangement of subvarieties (in the sense of Li [34]) given by the closures of the layers of 𝒜\mathcal{A}. One chooses a suitable building set 𝒢\mathcal{G} of subvarieties in XΔX_{\Delta} and blowups them in a prescribed order to obtain the projective wonderful model Y(XΔ,𝒢)Y(X_{\Delta},\mathcal{G}). The 𝒢\mathcal{G}-nested sets describe the boundary of the model. The definitions of building sets and nested sets are recalled in this section. Example 2.19 provides a non trivial instance in dimension 3 of this construction, computed with the help of a SageMath program (see [14]).

In Section 3 we recall from [12] the presentation of the integer cohomology ring of Y(XΔ,𝒢)Y(X_{\Delta},\mathcal{G}) as a quotient of a polynomial ring via generators and relations.

Section 4 is devoted to our main result. We provide a description of a monomial \mathbb{Z}-basis of H(Y(XΔ,𝒢),)H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z}). Every element of this basis has two factors: one is a monomial that depends essentially on a nested set 𝒮\mathcal{S} of 𝒢\mathcal{G} with certain labels (in analogy with the case of subspace models), the other one comes from the cohomology of a toric subvariety of XΔX_{\Delta} associated with 𝒮\mathcal{S}. With the help of the above mentioned SageMath program we provide a basis for the model of Example 2.19.

Finally, we devote Section 5 to the case of a divisorial toric arrangement of type AnA_{n}. We make a canonical choice of the fan Δ\Delta, i.e. we take the fan associated with the Coxeter chambers. Therefore XΔX_{\Delta} is the toric variety of type AnA_{n} studied for instance in [43, 48, 46, 22] and the minimal toric projective model is isomorphic to the moduli space of stable curves of genus 0 with n+3n+3 marked points, i.e. to the minimal projective wonderful model of the hyperplane arrangement of type An+1A_{n+1}.

This suggests to compare the new basis described in this paper with the basis coming from [49, 30]. Both bases are described by labeled graphs. On one side we have forests on n+1n+1 leaves with labels on internal vertices, equipped by an additional label: a permutation in the symmetric group SjS_{j}, where jj is the number of trees. On the other side we have forests on n+2n+2 leaves with labels on internal vertices. We will show an explicit bijection between these two families of forests. This will also provide us with a new combinatorial proof, with a geometric interpretation, of the equidistribution of two statistics on the symmetric group: the statistic of descents des and the statistic lec introduced by Foata and Han in [29] (both give rise to the Eulerian numbers).

2. Brief description of compact models

In this section we recall the construction of a wonderful model starting from a toric arrangement 𝒜\mathcal{A}, mainly following [11] (see also [14]).

First of all, let us fix some notation that will be used throughout this paper. Given a set AA, we will use the symbol A\operatorname{\cap}A to denote the intersection of its elements, namely

A=BAB.\operatorname{\cap}A=\bigcap_{B\in A}B.

Recall from the Introduction that X(T)X^{*}(T) is the group of characters of the torus TT; likewise, we denote by X(T)X_{*}(T) the group of one-parameters subgroups of TT. Moreover we define the vector spaces V=X(T)V=X_{*}(T)\otimes_{\mathbb{Z}}\mathbb{R} and its dual V=X(T)V^{*}=X^{*}(T)\otimes_{\mathbb{Z}}\mathbb{R}. The usual pairing X(T)×X(T)X^{*}(T)\times X_{*}(T)\to\mathbb{Z} and its extension to V×VV^{*}\times V\to\mathbb{R} will both be denoted by the symbol ,\left\langle\cdot,\cdot\right\rangle. Given Γ<X(T)\Gamma<X^{*}(T), we define

(2.1) VΓ{vVχ,v=0 for all χΓ}.V_{\Gamma}\coloneqq\left\{{v}\in V\mid\left\langle\chi,{v}\right\rangle=0\text{ for all }\chi\in\Gamma\right\}.

Given a fan Δ\Delta in VV, the corresponding toric variety will be denoted by XΔX_{\Delta}.

We want to build a model following the techniques described by Li in [34]: in that paper, which is inspired by [16, 15, 36], the author describes the construction of a compact model starting from an arrangement of subvarieties.

Definition 2.1.

Let XX be a non-singular algebraic variety. A simple arrangement of subvarieties of XX is a finite set Λ\Lambda of non-singular closed connected subvarieties properly contained in XX such that

  1. (1)

    for every two Λi,ΛjΛ\Lambda_{i},\Lambda_{j}\in\Lambda, either ΛiΛjΛ\Lambda_{i}\cap\Lambda_{j}\in\Lambda or ΛiΛj=\Lambda_{i}\cap\Lambda_{j}=\emptyset;

  2. (2)

    if ΛiΛj\Lambda_{i}\cap\Lambda_{j}\neq\emptyset, the intersection is clean, i.e. it is non-singular and for every yΛiΛjy\in\Lambda_{i}\cap\Lambda_{j} we have the following conditions on the tangent spaces:

    Ty(ΛiΛj)=Ty(Λi)Ty(Λj).\mathrm{T}_{y}(\Lambda_{i}\cap\Lambda_{j})=\mathrm{T}_{y}(\Lambda_{i})\cap\mathrm{T}_{y}(\Lambda_{j}).
Definition 2.2.

Let XX be a non-singular algebraic variety. An arrangement of subvarieties of XX is a finite set Λ\Lambda of non-singular closed connected subvarieties properly contained in XX such that

  1. (1)

    for every two Λi,ΛjΛ\Lambda_{i},\Lambda_{j}\in\Lambda, either ΛiΛj\Lambda_{i}\cap\Lambda_{j} is a disjoint union of elements of Λ\Lambda or ΛiΛj=\Lambda_{i}\cap\Lambda_{j}=\emptyset;

  2. (2)

    if ΛiΛj\Lambda_{i}\cap\Lambda_{j}\neq\emptyset, the intersection is clean.

In the toric arrangements setting, the subvarieties will be given by the intersections of the layers of the arrangement, so we introduce the combinatorial object that describe them.

Definition 2.3.

The poset of layers of a toric arrangement 𝒜\mathcal{A} is the set 𝒞(𝒜)\mathcal{C}(\mathcal{A}) of the connected components of the intersections of some layers of 𝒜\mathcal{A}, partially ordered by reverse inclusion.

Remark 2.4.
  1. (1)

    The whole torus belongs to 𝒞(𝒜)\mathcal{C}(\mathcal{A}), as it can be obtained as the intersection of no layers; we define 𝒞0(𝒜)𝒞(𝒜){T}\mathcal{C}_{0}(\mathcal{A})\coloneqq\mathcal{C}(\mathcal{A})\setminus\{T\}.

  2. (2)

    The intersection of two layers 𝒦(Γ1,ϕ1)\mathcal{K}(\Gamma_{1},\phi_{1}) and 𝒦(Γ2,ϕ2)\mathcal{K}(\Gamma_{2},\phi_{2}) is the disjont union of layers of the form 𝒦(Γ,ϕi)\mathcal{K}(\Gamma,\phi_{i}), i.e. they share the same Γ\Gamma, namely the saturation of Γ1+Γ2\Gamma_{1}+\Gamma_{2}.

Given a toric arrangement 𝒜\mathcal{A} in a torus TT, we embed TT in a suitable compact toric variety. In particular we build a toric variety whose associated fan satisfies the following equal sign condition.

Definition 2.5.

Let Δ\Delta be a fan in VV. An element χX(T)\chi\in X^{*}(T) has the equal sign property with respect to Δ\Delta if, for every cone CΔC\in\Delta, either χ,c0\left\langle\chi,c\right\rangle\geq 0 for all cCc\in C or χ,c0\left\langle\chi,c\right\rangle\leq 0 for all cCc\in C.

Definition 2.6.

Let Δ\Delta be a fan in VV and let 𝒦(Γ,ϕ)\mathcal{K}(\Gamma,\phi) be a layer. A \mathbb{Z}-basis (χ1,,χm)(\chi_{1},\dotsc,\chi_{m}) for Γ\Gamma is an equal sign basis with respect to Δ\Delta if χi\chi_{i} has the equal sign property for all i=1,,mi=1,\dotsc,m.

We say that a toric variety XΔX_{\Delta} is good for an arrangement 𝒜\mathcal{A} if each layer of 𝒞(𝒜)\mathcal{C}(\mathcal{A}) has an equal sign basis with respect to the fan Δ\Delta. In fact in this situation the following Theorem holds; we present the statement from [12], which summarizes Proposition 3.1 and Theorem 3.1 from [11].

Theorem 2.7 ([12, Theorem 5.1]).

For any layer 𝒦(Γ,ϕ)𝒞(𝒜)\mathcal{K}(\Gamma,\phi)\in\mathcal{C}(\mathcal{A}) let H=H(Γ)χΓker(xχ)H=H(\Gamma)\coloneqq\cap_{\chi\in\Gamma}\ker(x_{\chi}) be the corresponding homogeneous subtorus and let VΓV_{\Gamma} as in (2.1), i.e.

VΓ{vVχ,v=0 for all χΓ}.V_{\Gamma}\coloneqq\left\{{v}\in V\mid\left\langle\chi,{v}\right\rangle=0\text{ for all }\chi\in\Gamma\right\}.
  1. (1)

    For every cone CΔC\in\Delta, its relative interior is either entirely contained in VΓV_{\Gamma} or disjoint from VΓV_{\Gamma}.

  2. (2)

    The collection of cones CΔC\in\Delta which are contained in VΓV_{\Gamma} is a smooth fan ΔH\Delta_{H}.

  3. (3)

    𝒦(Γ,ϕ)¯\overline{\mathcal{K}(\Gamma,\phi)} is a smooth HH-variety whose fan is ΔH\Delta_{H}.

  4. (4)

    Let 𝒪\mathcal{O} be an orbit of TT in XΔX_{\Delta} and let C𝒪ΔC_{\mathcal{O}}\in\Delta be the corresponding cone. Then

    1. (a)

      if C𝒪C_{\mathcal{O}} is not contained in VΓV_{\Gamma}, 𝒪¯𝒦(Γ,ϕ)¯=\overline{\mathcal{O}}\cap\overline{\mathcal{K}(\Gamma,\phi)}=\emptyset;

    2. (b)

      If C𝒪VΓC_{\mathcal{O}}\subset V_{\Gamma}, 𝒪𝒦(Γ,ϕ)¯\mathcal{O}\cap\overline{\mathcal{K}(\Gamma,\phi)} is the HH-orbit in 𝒦(Γ,ϕ)¯\overline{\mathcal{K}(\Gamma,\phi)} corresponding to C𝒪ΔHC_{\mathcal{O}}\in\Delta_{H}.

As a consequence the set of the connected components of the intersections of the closures of the layers 𝒦(Γ,ϕ)𝒜\mathcal{K}(\Gamma,\phi)\in\mathcal{A} in XΔX_{\Delta} is an arrangement of subvarieties according to Li’s definition.

Following [11] we now introduce the wonderful model associated with an arrangement Λ\Lambda of subvarieties in a generic non-singular algebraic variety XX. To do so we need to define the notion of building sets and nested sets.

Definition 2.8.

Let Λ\Lambda be a simple arrangement of subvarieties. A subset 𝒢Λ\mathcal{G}\subseteq\Lambda is a building set for Λ\Lambda if for every LΛ𝒢L\in\Lambda\setminus\mathcal{G} the minimal elements (w.r.t. the inclusion) of the set {G𝒢LG}\{G\in\mathcal{G}\mid L\subset G\} intersect transversally and their intersection is LL. These minimal elements are called the 𝒢\mathcal{G}-factors of LL.

Definition 2.9.

Let 𝒢\mathcal{G} be a building set for a simple arrangement Λ\Lambda. A subset 𝒮𝒢\mathcal{S}\subseteq\mathcal{G} is called (𝒢\mathcal{G}-)nested if for any antichain111An antichain in a poset is a set of pairwise non-comparable elements. {A1,,Ak}𝒮\{A_{1},\dotsc,A_{k}\}\subseteq\mathcal{S}, with k2k\geq 2, there is an element in Λ\Lambda of which A1,,AkA_{1},\dotsc,A_{k} are the 𝒢\mathcal{G}-factors.

Remark 2.10.

Since the empty set has no antichains of cardinality at least 22, the definition above applies vacuously for it.

Remark 2.11.

We notice that if \mathcal{H} is a subset of 𝒢\mathcal{G} whose elements have empty intersection, then it cannot be contained in any 𝒢\mathcal{G}-nested set.

In case the arrangement Λ\Lambda is not simple, the definitions above apply locally: first of all, we define the restriction of an arrangement of subvarieties Λ\Lambda to an open set UXU\subseteq X to be the set

Λ|U{ΛiUΛiΛ,ΛiU}.\Lambda|_{U}\coloneqq\{\Lambda_{i}\cap U\mid\Lambda_{i}\in\Lambda,\ \Lambda_{i}\cap U\neq\emptyset\}.
Definition 2.12.

Let Λ\Lambda be an arrangement of subvarieties of XX. A subset 𝒢Λ\mathcal{G}\subseteq\Lambda is a building set for Λ\Lambda if there is a cover 𝒰\mathcal{U} of open sets of XX such that

  1. (1)

    for every U𝒰U\in\mathcal{U}, the restriction Λ|U\Lambda|_{U} is simple;

  2. (2)

    for every U𝒰U\in\mathcal{U}, 𝒢|U\mathcal{G}|_{U} is a building set for Λ|U\Lambda|_{U}.

Definition 2.13.

Let 𝒢\mathcal{G} be a building set for an arrangement Λ\Lambda. A subset 𝒮𝒢\mathcal{S}\subseteq\mathcal{G} is called (𝒢\mathcal{G}-)nested if there is an open cover 𝒰\mathcal{U} of XX such that, for every U𝒰U\in\mathcal{U}, Λ|U\Lambda|_{U} is simple, 𝒢|U\mathcal{G}|_{U} is building for Λ|U\Lambda|_{U} and for at least one W𝒰W\in\mathcal{U}, 𝒮|W\mathcal{S}|_{W} is 𝒢|W\mathcal{G}|_{W}-nested. (In particular AWA\cap W\neq\emptyset for all A𝒮A\in\mathcal{S}.)

Instead of defining a building set in terms of a given arrangement, it is often convenient to study the notion of “building” as an intrinsic property of a set of subvarieties.

Definition 2.14.

A finite set 𝒢\mathcal{G} of connected subvarieties of XX is called a building set if the set of the connected components of all the possible intersections of collections of subvarieties from 𝒢\mathcal{G} is an arrangement of subvarieties, called the arrangement induced by 𝒢\mathcal{G} and denoted by Λ(𝒢)\Lambda(\mathcal{G}), and 𝒢\mathcal{G} is a building set for Λ(𝒢)\Lambda(\mathcal{G}) according to Definition 2.12.

From now on, Definition 2.14 applies when we refer to a set of subvarieties as “building” without specifying the arrangement.

Given an arrangement Λ\Lambda of a non-singular variety XX and a building set 𝒢\mathcal{G} for Λ\Lambda, a wonderful model Y(X,𝒢)Y(X,\mathcal{G}) can be obtained as the closure of the locally closed embedding:

(XΛiΛΛi)G𝒢BlGX\left(X\setminus\bigcup_{\Lambda_{i}\in\Lambda}\Lambda_{i}\right)\longrightarrow\prod_{G\in\mathcal{G}}\operatorname{Bl}_{G}{X}

where BlGX\operatorname{Bl}_{G}{X} is the blowup of XX along GG. Concretely we can build Y(X,𝒢)Y(X,\mathcal{G}) one step at a time, through a series of blowups, as described in the following theorem.

Theorem 2.15 (see [34, Theorem 1.3]).

Let 𝒢\mathcal{G} be a building set in a non-singular variety XX. Let us order the elements G1,,GmG_{1},\dotsc,G_{m} of 𝒢\mathcal{G} in such a way that for every 1km1\leq k\leq m the set 𝒢k{G1,,Gk}\mathcal{G}_{k}\coloneqq\{G_{1},\dotsc,G_{k}\} is building. Then if we set X0XX_{0}\coloneqq X and XkY(X,𝒢k)X_{k}\coloneqq Y(X,\mathcal{G}_{k}) for 1km1\leq k\leq m, we have

Xk=Blt(Gk)Xk1,X_{k}=\operatorname{Bl}_{t(G_{k})}{X_{k-1}},

where t(Gk)t(G_{k}) denotes the dominant transform222In the blowup of a variety MM along a centre FF the dominant transform of a subvariety ZZ coincides with the proper transform if ZFZ\nsubseteq F (and therefore it is isomorphic to the blowup of ZZ along ZFZ\cap F), and with π1(Z)\pi^{-1}(Z) if ZFZ\subseteq F, where π:BlFMM\pi\colon\operatorname{Bl}_{F}{M}\to M is the projection. We will use the same notation t(Z)t(Z) for both the proper and the dominant transform of ZZ, if no confusion arises. of GkG_{k} in Xk1X_{k-1}.

Remark 2.16.

Any total ordering of the elements of a building set 𝒢={G1,,Gm}\mathcal{G}=\{G_{1},\dotsc,G_{m}\} which refines the ordering by inclusion, that is i<ji<j if GiGjG_{i}\subset G_{j}, satisfies the condition of Theorem 2.15.

Let us denote by π:Y(X,𝒢)X\pi\colon Y(X,\mathcal{G})\to X the blowup map. The boundary of Y(X,𝒢)Y(X,\mathcal{G}) admits a description in terms of 𝒢\mathcal{G}-nested sets.

Theorem 2.17 (see [34, Theorem 1.2]).

The complement in Y(X,𝒢)Y(X,\mathcal{G}) of π1(XΛi)\pi^{-1}(X\setminus\bigcup\Lambda_{i}) is the union of the divisors t(G)t(G), where GG ranges among the elements of 𝒢\mathcal{G}. Let 𝒰\mathcal{U} be an open cover of XX such that for every U𝒰U\in\mathcal{U} Λ|U\Lambda|_{U} is simple and 𝒢|U\mathcal{G}|_{U} is building for Λ|U\Lambda|_{U}. Then, given U𝒰U\in\mathcal{U} and A1,,Ak𝒢A_{1},\dotsc,A_{k}\in\mathcal{G}, the intersection t(A1)t(Ak)π1(U)t(A_{1})\cap\dotsb\cap t(A_{k})\cap\pi^{-1}(U) is non-empty if and only if {A1|U,,Ak|U}\{A_{1}|_{U},\dotsc,A_{k}|_{U}\} is 𝒢|U\mathcal{G}|_{U}-nested; moreover, if the intersection is non-empty then it is transversal.

Remark 2.18.

We notice that Definition 2.13 and the statement of Theorem 2.17 are slightly different from the ones in the literature (see [34, 11, 12, 13, 14]), where a subset 𝒮𝒢\mathcal{S}\subseteq\mathcal{G} is considered 𝒢\mathcal{G}-nested if 𝒮|U\mathcal{S}|_{U} is 𝒢|U\mathcal{G}|_{U}-nested for every U𝒰U\in\mathcal{U}. We think that our Definition 2.13 and Theorem 2.17 are more precise and remove an ambiguity, since they point out that the intersection property depends on the property of being nested locally in the chart π1(U)\pi^{-1}(U) of Y(X,𝒢)Y(X,\mathcal{G}).

Example 2.19.

In order to compute some non-trivial examples, a series of scripts, extending the ones described in [42], were developed in the SageMath environment [45].

Let 𝒜={𝒦1,𝒦2,𝒦3}\mathcal{A}=\{\mathcal{K}_{1},\mathcal{K}_{2},\mathcal{K}_{3}\} be the arrangement in ()3(\mathbb{C}^{*})^{3}, with coordinates (x,y,z)(x,y,z), whose layers are defined by the equations

𝒦1:\displaystyle\mathcal{K}_{1}\colon xz2=1,\displaystyle xz^{2}=1,
𝒦2:\displaystyle\mathcal{K}_{2}\colon z=xy,\displaystyle z=xy,
𝒦3:\displaystyle\mathcal{K}_{3}\colon xy2=1,\displaystyle xy^{2}=1,

We can view them as 𝒦i=𝒦(Γi,ϕi)\mathcal{K}_{i}=\mathcal{K}(\Gamma_{i},\phi_{i}) where Γ1<3\Gamma_{1}<\mathbb{Z}^{3} is generated by (1,0,2)(1,0,2), Γ2\Gamma_{2} by (1,1,1)(1,1,-1) and Γ3\Gamma_{3} by (1,2,0)(1,2,0), and ϕi\phi_{i} is the constant function equal to 11 for i=1,2,3i=1,2,3. Figure 1 represents the Hasse diagram of the poset of layers 𝒞(𝒜)\mathcal{C}(\mathcal{A}).

In order to define a projective model for the arrangement, the first ingredient is a good toric variety. For this example, following the algorithm of [11], we built a toric variety XΔX_{\Delta} whose fan has 72 rays and 140 maximal cones, listed in Appendix A respectively. This is not the smallest fan associated with a good toric variety for this arrangement, but it has some addictional properties that are useful for the computation of a presentation for the cohomology ring of Y(XΔ,𝒢)Y(X_{\Delta},\mathcal{G}).

The next choice is a building set 𝒢\mathcal{G}; for this example we use the subset of the elements of 𝒞(𝒜)\mathcal{C}(\mathcal{A}) that are pictured in a double circle in Figure 1 and obtain the model Y𝒜=Y(XΔ,𝒢)Y_{\mathcal{A}}=Y(X_{\Delta},\mathcal{G}). We will study this model in later examples of this paper.

TT𝒦1\mathcal{K}_{1}𝒦2\mathcal{K}_{2}𝒦3\mathcal{K}_{3}L1L_{1}L2L_{2}L3L_{3}L4L_{4}P1P_{1}P2P_{2}P3P_{3}P4P_{4}
Figure 1. Poset of layers for the arrangement of Example 2.19, with the elements of the building set highlighted with a double circle.

3. Presentation of the cohomology ring

In this section we recall a presentation of the cohomology ring of the model Y(XΔ,𝒢)Y(X_{\Delta},\mathcal{G}). As we have seen, given a toric arrangement 𝒜\mathcal{A} and a toric variety XΔX_{\Delta} which is good for it, the set Λ={𝒦¯𝒦𝒞0(𝒜)}\Lambda=\{\overline{\mathcal{K}}\mid\mathcal{K}\in\mathcal{C}_{0}(\mathcal{A})\} is an arrangement of subvarieties of XΔX_{\Delta} according to Li.

The cohomology ring is described as a quotient of a polynomial ring with coefficients in H(XΔ,)H^{*}(X_{\Delta},\mathbb{Z}). We present here a result by Danilov that provides an explicit presentation of the cohomology ring of a toric variety.

Theorem 3.1 ([10, Theorem 10.8]).

Let XΔX_{\Delta} be a smooth complete TT-variety. Let \mathcal{R} be the set of primitive generators of the rays of Δ\Delta and define a polynomial indeterminate CrC_{r} for each rr\in\mathcal{R}. Then

H(XΔ,)[Crr]/(SR+L)H^{*}(X_{\Delta},\mathbb{Z})\simeq\mathbb{Z}[C_{r}\mid r\in\mathcal{R}]/(\mathcal{I}_{\mathrm{SR}}+\mathcal{I}_{\mathrm{L}})

where

  • SR\mathcal{I}_{\mathrm{SR}} is the Stanley-Reisner ideal

    SR(Cr1Crkr1,,rk do not belong to a cone of Δ);\mathcal{I}_{\mathrm{SR}}\coloneqq(C_{r_{1}}\dotsb C_{r_{k}}\mid r_{1},\dotsc,r_{k}\text{ do not belong to a cone of }\Delta);
  • L\mathcal{I}_{\mathrm{L}} is the linear equivalence ideal

    L(rβ,rCr|βX(T)).\mathcal{I}_{\mathrm{L}}\coloneqq\left(\sum_{r\in\mathcal{R}}\left\langle\beta,r\right\rangle C_{r}\,\middle|\,\beta\in X^{*}(T)\right).

Notice that for SR\mathcal{I}_{\mathrm{SR}} it is sufficient to take only the square-free monomials, and for L\mathcal{I}_{\mathrm{L}} it is sufficient to take only the β\beta’s belonging to a basis of X(T)X^{*}(T).

Furthermore the residue class of CrC_{r} in H2(XΔ,)H^{2}(X_{\Delta},\mathbb{Z}) is the cohomology class of the divisor DrD_{r} associated with the ray rr for each rr\in\mathcal{R}. By abuse of notation we are going to denote this residue class in H2(XΔ,)H^{2}(X_{\Delta},\mathbb{Z}) also by CrC_{r}.

Remark 3.2.

Given a layer 𝒦(Γ,ϕ)\mathcal{K}(\Gamma,\phi), the inclusion j:𝒦(Γ,ϕ)¯XΔj\colon\overline{\mathcal{K}(\Gamma,\phi)}\hookrightarrow X_{\Delta} induces a restriction map in cohomology

(3.1) j:H(XΔ,)H(𝒦(Γ,ϕ)¯,).j^{*}\colon H^{*}(X_{\Delta},\mathbb{Z})\to H^{*}(\overline{\mathcal{K}(\Gamma,\phi)},\mathbb{Z}).

As noted in [12, Proposition 5.4] this map is surjective and its kernel is generated by {Crr,rVΓ}\{C_{r}\mid r\in\mathcal{R},\ r\notin V_{\Gamma}\}. In the sequel, we identify 𝒦(Γ,ϕ)¯\overline{\mathcal{K}(\Gamma,\phi)} with XΔHX_{\Delta_{H}}, where ΔH\Delta_{H} is the same fan of Theorem 2.7 (point 3).

Example 3.3 (Example 2.19, continued).

We computed the presentation of H(XΔ,)H^{*}(X_{\Delta},\mathbb{Z}) as in Theorem 3.1 for the toric variety XΔX_{\Delta} of Example 2.19. The cohomology ring is isomorphic to a quotient of the ring [C1,,C72]\mathbb{Z}[C_{1},\dotsc,C_{72}] where each indeterminate CiC_{i} corresponds to the ray rir_{i} as listed in the table in Appendix A. We won’t report the full presentation here and give only the Betti numbers:

rk(H0(XΔ,))\displaystyle\operatorname{rk}(H^{0}(X_{\Delta},\mathbb{Z})) =1,\displaystyle{}=1,
rk(H2(XΔ,))\displaystyle\operatorname{rk}(H^{2}(X_{\Delta},\mathbb{Z})) =69,\displaystyle{}=69,
rk(H4(XΔ,))\displaystyle\operatorname{rk}(H^{4}(X_{\Delta},\mathbb{Z})) =69,\displaystyle{}=69,
rk(H6(XΔ,))\displaystyle\operatorname{rk}(H^{6}(X_{\Delta},\mathbb{Z})) =1.\displaystyle{}=1.

The presentation of the cohomology ring of Y(XΔ,𝒢)Y(X_{\Delta},\mathcal{G}) has been computed with an additional hypothesis on the building set.

Definition 3.4.

A building set 𝒢\mathcal{G} for Λ\Lambda is well-connected if for any subset 𝒢\mathcal{H}\subseteq\mathcal{G}, if the intersection \operatorname{\cap}\mathcal{H} has two or more connected components, then each of these components belongs to 𝒢\mathcal{G}.

Example 3.5 (Example 2.19, continued).

The building set 𝒢\mathcal{G} described in Example 2.19 is a well-connected building set.

Some general properties of well-connected building sets are studied and presented in [14, Section 6].

We recall here the main ingredients for the presentation of H(Y(XΔ,𝒢),)H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z}). Let ZZ be an indeterminate and let R=H(XΔ,)R=H^{*}(X_{\Delta},\mathbb{Z}) viewed as [Crr]/(SR+L)\mathbb{Z}[C_{r}\mid r\in\mathcal{R}]/(\mathcal{I}_{\mathrm{SR}}+\mathcal{I}_{\mathrm{L}}) as in Theorem 3.1. (For brevity we will use again the symbols CrC_{r} instead of the corresponding equivalence classes in the quotient.) For every GΛG\in\Lambda we denote by ΓG\Gamma_{G} the lattice such that G=𝒦(ΓG,ϕ)¯G=\overline{\mathcal{K}(\Gamma_{G},\phi)}.

Given a pair (M,G)(Λ{XΔ})×Λ(M,G)\in(\Lambda\cup\{X_{\Delta}\})\times\Lambda with GMG\subseteq M, we can choose a basis (χ1,,χs)(\chi_{1},\dotsc,\chi_{s}) for ΓG\Gamma_{G} such that it is equal sign with respect to Δ\Delta and that (χ1,,χk)(\chi_{1},\dotsc,\chi_{k}), with ksk\leq s, is a basis for ΓM\Gamma_{M} (if MM is the whole variety XΔX_{\Delta}, then we choose any equal sign basis of ΓG\Gamma_{G} and let k=0k=0). We define the polynomials PGMR[Z]P^{M}_{G}\in R[Z] as

PGM(Z)j=k+1s(Zrmin(0,χj,r)Cr).P^{M}_{G}(Z)\coloneqq\prod_{j=k+1}^{s}\left(Z-\sum_{r\in\mathcal{R}}\min(0,\left\langle\chi_{j},r\right\rangle)C_{r}\right).

If G=MG=M, we set PGG1P_{G}^{G}\coloneqq 1 since it is an empty product. As shown in [12, Proposition 6.3], different choices of the PGMP_{G}^{M}’s are possible; for example, in [40, Remark 4.4], the authors suggest the polynomials

PGM(Z)=Zsk+j=k+1s(rmin(0,χj,r)Cr).P^{M}_{G}(Z)=Z^{s-k}+\prod_{j=k+1}^{s}\left(-\sum_{r\in\mathcal{R}}\min(0,\left\langle\chi_{j},r\right\rangle)C_{r}\right).

Now we define the set

𝒲{(G,)𝒢×𝒫(𝒢)GH for all H},\mathcal{W}\coloneqq\{(G,\mathcal{H})\in\mathcal{G}\times\operatorname{\mathcal{P}}(\mathcal{G})\mid G\subsetneq H\text{ for all }H\in\mathcal{H}\},

where 𝒫(𝒢)\operatorname{\mathcal{P}}(\mathcal{G}) is the power set of 𝒢\mathcal{G}. Notice that (G,)𝒲(G,\emptyset)\in\mathcal{W} for all G𝒢G\in\mathcal{G}. For each pair (G,)𝒲(G,\mathcal{H})\in\mathcal{W} we define a relation F(G,)F(G,\mathcal{H}) in the following way: let MM be the unique connected component of \operatorname{\cap}{\mathcal{H}} that contains GG (as usual, if =\mathcal{H}=\emptyset then M=XΔM=X_{\Delta}), and for G𝒢G\in\mathcal{G} let 𝒢G{H𝒢HG}\mathcal{G}_{G}\coloneqq\{H\in\mathcal{G}\mid H\subseteq G\}; with this information we define the polynomial F(G,)R[TGG𝒢]F(G,\mathcal{H})\in R[T_{G}\mid G\in\mathcal{G}] as

F(G,)PGM(H𝒢GTH)KTK.F(G,\mathcal{H})\coloneqq P_{G}^{M}\Big{(}\sum_{H\in\mathcal{G}_{G}}-T_{H}\Big{)}\prod_{K\in\mathcal{H}}T_{K}.

Finally let 𝒲0{𝒫(𝒢)=}\mathcal{W}_{0}\coloneqq\{\mathcal{H}\in\operatorname{\mathcal{P}}(\mathcal{G})\mid\operatorname{\cap}\mathcal{H}=\emptyset\}. For each 𝒲0\mathcal{H}\in\mathcal{W}_{0} we define the polynomial F()R[TGG𝒢]F(\mathcal{H})\in R[T_{G}\mid G\in\mathcal{G}] as

F()KTK.F(\mathcal{H})\coloneqq\prod_{K\in\mathcal{H}}T_{K}.
Theorem 3.6 ([12, Theorem 7.1]).

Let 𝒜\mathcal{A}, XΔX_{\Delta} and Λ\Lambda be as in the beginning of this section and let 𝒢\mathcal{G} be a well-connected building set for Λ\Lambda. Let also R=H(XΔ,)R=H^{*}(X_{\Delta},\mathbb{Z}) viewed as a polynomial ring as in Theorem 3.1. The cohomology ring of the wonderful model H(Y(XΔ,𝒢),)H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z}) is isomorphic to the quotient of R[TGG𝒢]R[T_{G}\mid G\in\mathcal{G}] by the ideal 𝒢\mathcal{I}_{\mathcal{G}} generated by

  1. (1)

    the products CrTGC_{r}T_{G}, with G𝒢G\in\mathcal{G} and rr\in\mathcal{R} such that rr does not belong to VΓGV_{\Gamma_{G}};

  2. (2)

    the polynomials F(G,)F(G,\mathcal{H}) for every pair (G,)𝒲(G,\mathcal{H})\in\mathcal{W};

  3. (3)

    the polynomials F()F(\mathcal{H}) for every 𝒲0\mathcal{H}\in\mathcal{W}_{0}.

The isomorphism is given by sending TGT_{G} for G𝒢G\in\mathcal{G} to the cohomology class associated with the divisor in the boundary which is the transform of GG (t(G)t(G)). Putting all together, we have

H(Y(XΔ,𝒢),)\displaystyle H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z}) R[TGG𝒢]/𝒢\displaystyle{}\simeq R[T_{G}\mid G\in\mathcal{G}]/\mathcal{I}_{\mathcal{G}}
[Cr,TGr,G𝒢]/(SR+L+𝒢).\displaystyle{}\simeq\mathbb{Z}[C_{r},T_{G}\mid r\in\mathcal{R},G\in\mathcal{G}]/(\mathcal{I}_{\mathrm{SR}}+\mathcal{I}_{\mathrm{L}}+\mathcal{I}_{\mathcal{G}}).
Remark 3.7.

It was already noted in [11, Theorem 9.1] that the cohomology of the projective wonderful model Y(XΔ,𝒢)Y(X_{\Delta},\mathcal{G}) is a free \mathbb{Z}-module and Hi(Y(XΔ,𝒢),)=0H^{i}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z})=0 for ii odd.

Example 3.8 (Example 2.19, continued).

We computed the presentation of H(Y𝒜,)H^{*}(Y_{\mathcal{A}},\mathbb{Z}) as in Theorem 3.6 for the model Y𝒜Y_{\mathcal{A}} of Example 2.19. The cohomology ring is isomorphic to a quotient of the ring [C1,,C72,T1,,T9]\mathbb{Z}[C_{1},\dotsc,C_{72},T_{1},\dotsc,T_{9}], where the CiC_{i}’s are the same as Example 3.3 and each TjT_{j} corresponds to an element of 𝒢\mathcal{G} in the following way:

𝒦1\displaystyle\mathcal{K}_{1} T1,\displaystyle{}\rightsquigarrow T_{1}, 𝒦2\displaystyle\mathcal{K}_{2} T2,\displaystyle{}\rightsquigarrow T_{2}, 𝒦3\displaystyle\mathcal{K}_{3} T3,\displaystyle{}\rightsquigarrow T_{3}, L2\displaystyle L_{2} T4,\displaystyle{}\rightsquigarrow T_{4}, L3\displaystyle L_{3} T5,\displaystyle{}\rightsquigarrow T_{5},
P1\displaystyle P_{1} T6,\displaystyle{}\rightsquigarrow T_{6}, P3\displaystyle P_{3} T7,\displaystyle{}\rightsquigarrow T_{7}, P2\displaystyle P_{2} T8,\displaystyle{}\rightsquigarrow T_{8}, P4\displaystyle P_{4} T9.\displaystyle{}\rightsquigarrow T_{9}.

Once again we won’t report the full presentation here and give only the Betti numbers:

rk(H0(Y𝒜,))\displaystyle\operatorname{rk}(H^{0}(Y_{\mathcal{A}},\mathbb{Z})) =1,\displaystyle{}=1,
rk(H2(Y𝒜,))\displaystyle\operatorname{rk}(H^{2}(Y_{\mathcal{A}},\mathbb{Z})) =75,\displaystyle{}=75,
rk(H4(Y𝒜,))\displaystyle\operatorname{rk}(H^{4}(Y_{\mathcal{A}},\mathbb{Z})) =75,\displaystyle{}=75,
rk(H6(Y𝒜,))\displaystyle\operatorname{rk}(H^{6}(Y_{\mathcal{A}},\mathbb{Z})) =1.\displaystyle{}=1.

In the next Section we are going to give a description of a monomial basis of R[TGG𝒢]/𝒢R[T_{G}\mid G\in\mathcal{G}]/\mathcal{I}_{\mathcal{G}}.

4. Main theorem

Let 𝒜\mathcal{A} be a toric arrangement, let XΔX_{\Delta} be a good toric variety for it with associated fan Δ\Delta, let Λ\Lambda be the poset of intersections of the closures of the layers of 𝒜\mathcal{A} in XΔX_{\Delta} which is an arrangement of subvarieties, and let 𝒢\mathcal{G} be a well-connected building set for Λ\Lambda. In this Section we show a basis of the cohomology of Y(XΔ,𝒢)Y(X_{\Delta},\mathcal{G}) in terms of admissible functions (see Definition 4.4). The notion of admissible function is analogue to the one used in the linear setting case of subspace arrangements, which is introduced in [30].

We begin by proving a characterization of 𝒢\mathcal{G}-nested sets in this case, which will be useful in the proof of our main theorem.

Proposition 4.1.

Let 𝒢\mathcal{G} be a well-connected building set. A subset 𝒮𝒢\mathcal{S}\subseteq\mathcal{G} is 𝒢\mathcal{G}-nested if and only if for any antichain 𝒮\mathcal{H}\subseteq\mathcal{S} with at least two elements the intersection \operatorname{\cap}\mathcal{H} is non-empty, connected, transversal, and does not belong to 𝒢\mathcal{G}.

Proof of Proposition 4.1.

\Rightarrow Let 𝒮\mathcal{S} be a 𝒢\mathcal{G}-nested set an let ={A1,,Ak}\mathcal{H}=\{A_{1},\ldots,A_{k}\} be an antichain with k2k\geq 2. The intersection \operatorname{\cap}\mathcal{H} is not empty by Definition 2.13 and Remark 2.11. If \operatorname{\cap}\mathcal{H} is not connected then it would be the disjoint union of at least two elements in 𝒢\mathcal{G} by well-connectedness. Let us consider W𝒰W\in\mathcal{U} as in Definition 2.13, so that 𝒮|W\mathcal{S}|_{W} is 𝒢|W\mathcal{G}|_{W}-nested and for every A𝒮A\in\mathcal{S} AWA\cap W\neq\emptyset, therefore |W={A1W,,AkW}\mathcal{H}|_{W}=\{A_{1}\cap W,\dotsc,A_{k}\cap W\} is an antichain of 𝒮|W\mathcal{S}|_{W}. This implies that

(A1W)(AkW)(A_{1}\cap W)\cap\dotsb\cap(A_{k}\cap W)

would be the 𝒢|W\mathcal{G}|_{W}-decomposition of GWG\cap W, where G𝒢G\in\mathcal{G} is one of the connected components of \operatorname{\cap}\mathcal{H}. We reached a contradiction because GW𝒢|WG\cap W\in\mathcal{G}|_{W}, and we deduce that \operatorname{\cap}\mathcal{H} is connected and furthermore that it is not an element of 𝒢\mathcal{G}. The transversality of the intersection in WW also implies that \operatorname{\cap}\mathcal{H} is transversal.

\Leftarrow We now suppose that for every antichain 𝒮\mathcal{H}\subseteq\mathcal{S} with at least two elements the intersection \operatorname{\cap}\mathcal{H} is non-empty, connected, transversal and does not belong to 𝒢\mathcal{G}.

Let 𝒰\mathcal{U} be an open cover as in Definition 2.12. Let us first notice that 𝒮\operatorname{\cap}\mathcal{S} is equal to the intersection of the antichain given by the minimal elements of 𝒮\mathcal{S}, therefore it is non-empty and connected. Let U𝒰U\in\mathcal{U} such that (𝒮)U(\operatorname{\cap}\mathcal{S})\cap U is not empty. We will prove that 𝒮|U\mathcal{S}|_{U} is 𝒢|U\mathcal{G}|_{U}-nested.

Consider an antichain {A1U,,AkU}\{A_{1}\cap U,\dotsc,A_{k}\cap U\} in 𝒮|U\mathcal{S}|_{U} with k2k\geq 2 (in particular AiUA_{i}\cap U is non-empty for every ii): we will prove that A1U,,AkUA_{1}\cap U,\dotsc,A_{k}\cap U are the 𝒢|U\mathcal{G}|_{U}-factors of their intersection.

Let us put ={A1,,Ak}\mathcal{H}=\{A_{1},\dotsc,A_{k}\}. We observe that \mathcal{H} is an antichain, therefore by hypothesis we know that A=A=\operatorname{\cap}\mathcal{H} is non-empty, connected, transversal and does not belong to 𝒢\mathcal{G}. Since 𝒢|U\mathcal{G}|_{U} is building, we know that AUA\cap U is the transversal intersection of the minimal elements of 𝒢|U\mathcal{G}|_{U} among the ones containing AUA\cap U. We let those minimal elements be B1U,,BrUB_{1}\cap U,\ldots,B_{r}\cap U. For simplicity, in the rest of the proof we omit the reference to UU.

By minimality each of the AiA_{i}’s contains some of the BjB_{j}’s: in the next paragraph we partition \mathcal{H} in subsets according to this.

Up to reordering the indices of the BjB_{j}’s, we can assume that B1B_{1} is contained in some of the AiA_{i}’s, and define 1:={AiB1Ai}\mathcal{H}_{1}:=\{A_{i}\in\mathcal{H}\mid B_{1}\subseteq A_{i}\}. If 1=\mathcal{H}_{1}=\mathcal{H} we are done, otherwise there is another BjB_{j} contained in the elements of 1\mathcal{H}\setminus\mathcal{H}_{1}, and up to reordering we assume that this is B2B_{2}. We define 2:={Ai1B2Ai}\mathcal{H}_{2}:=\{A_{i}\in\mathcal{H}\setminus\mathcal{H}_{1}\mid B_{2}\subseteq A_{i}\}. If (12)\mathcal{H}\setminus(\mathcal{H}_{1}\cup\mathcal{H}_{2}) is not empty we repeat the process with B3B_{3} and obtain 3\mathcal{H}_{3}. We stop after \ell steps, where r\ell\leq r, when we have 1=\mathcal{H}_{1}\cup\dotsb\cup\mathcal{H}_{\ell}=\mathcal{H}.

By construction we have:

(4.1) B11,,B,B_{1}\subseteq\operatorname{\cap}\mathcal{H}_{1},\ldots,B_{\ell}\subseteq\operatorname{\cap}\mathcal{H}_{\ell},

so

(1)()==AB1B(1)()(\operatorname{\cap}\mathcal{H}_{1})\cap\dotsb\cap(\operatorname{\cap}\mathcal{H}_{\ell})=\operatorname{\cap}\mathcal{H}=A\subseteq B_{1}\cap\dotsb\cap B_{\ell}\subseteq(\operatorname{\cap}\mathcal{H}_{1})\cap\dotsb\cap(\operatorname{\cap}\mathcal{H}_{\ell})

and this implies that A=B1BA=B_{1}\cap\dotsb\cap B_{\ell}, that is to say, r=r=\ell.

Recall that B1BB_{1}\cap\dotsb\cap B_{\ell} is transversal and so is \operatorname{\cap}\mathcal{H}; since each j\mathcal{H}_{j} is an antichain, j\operatorname{\cap}\mathcal{H}_{j} is transversal too, and it follows that (1)()(\operatorname{\cap}\mathcal{H}_{1})\cap\dotsb\cap(\operatorname{\cap}\mathcal{H}_{\ell}) is transversal. From this information and from (4.1) we deduce that dim(Bj)=dim(j)\dim(B_{j})=\dim(\operatorname{\cap}\mathcal{H}_{j}) for all j=1,,j=1,\ldots,\ell, therefore Bj=jB_{j}=\operatorname{\cap}\mathcal{H}_{j}.

This implies that |j|=1\left|\mathcal{H}_{j}\right|=1 for any jj, otherwise if |j|>1\left|\mathcal{H}_{j}\right|>1 that would contradict the fact that the intersection of an antichain does not belong to 𝒢\mathcal{G}.

Since 1=={A1,,Ak}\mathcal{H}_{1}\cup\dotsb\cup\mathcal{H}_{\ell}=\mathcal{H}=\{A_{1},\ldots,A_{k}\} we deduce that k=k=\ell and, up to a relabeling, we can assume that i={Ai}\mathcal{H}_{i}=\{A_{i}\}. We conclude that Aj=BjA_{j}=B_{j} and this shows that A1,,AkA_{1},\ldots,A_{k} are the 𝒢\mathcal{G}-factors of AA. ∎

Remark 4.2.

Proposition 4.1 implies that when 𝒢\mathcal{G} is well-connected the property of being nested can be expressed in global terms (without charts).

Example 4.3 (Example 2.19, continued).

The building set 𝒢\mathcal{G} of Example 2.19 has 48 nested sets, namely

,{𝒦1},{𝒦2},{𝒦3},{L2},{L3},{P1},{P2},{P3},{P4},{𝒦1,𝒦2},{𝒦1,L2},{𝒦1,L3},{𝒦1,P1},{𝒦1,P2},{𝒦1,P3},{𝒦1,P4},{𝒦2,𝒦3},{𝒦2,P1},{𝒦2,P2},{𝒦2,P3},{𝒦2,P4},{𝒦3,L2},{𝒦3,L3},{𝒦3,P1},{𝒦3,P2},{𝒦3,P3},{𝒦3,P4},{L2,P1},{L2,P3},{L3,P2},{L3,P4},{𝒦1,𝒦2,P1},{𝒦1,𝒦2,P2},{𝒦1,𝒦2,P3},{𝒦1,𝒦2,P4},{𝒦1,L2,P1},{𝒦1,L2,P3},{𝒦1,L3,P2},{𝒦1,L3,P4},{𝒦2,𝒦3,P1},{𝒦2,𝒦3,P2},{𝒦2,𝒦3,P3},{𝒦2,𝒦3,P4},{𝒦3,L2,P1},{𝒦3,L2,P3},{𝒦3,L3,P2},{𝒦3,L3,P4}.\begin{array}[]{cccc}\emptyset,&\{\mathcal{K}_{1}\},&\{\mathcal{K}_{2}\},&\{\mathcal{K}_{3}\},\\ \{L_{2}\},&\{L_{3}\},&\{P_{1}\},&\{P_{2}\},\\ \{P_{3}\},&\{P_{4}\},&\{\mathcal{K}_{1},\mathcal{K}_{2}\},&\{\mathcal{K}_{1},L_{2}\},\\ \{\mathcal{K}_{1},L_{3}\},&\{\mathcal{K}_{1},P_{1}\},&\{\mathcal{K}_{1},P_{2}\},&\{\mathcal{K}_{1},P_{3}\},\\ \{\mathcal{K}_{1},P_{4}\},&\{\mathcal{K}_{2},\mathcal{K}_{3}\},&\{\mathcal{K}_{2},P_{1}\},&\{\mathcal{K}_{2},P_{2}\},\\ \{\mathcal{K}_{2},P_{3}\},&\{\mathcal{K}_{2},P_{4}\},&\{\mathcal{K}_{3},L_{2}\},&\{\mathcal{K}_{3},L_{3}\},\\ \{\mathcal{K}_{3},P_{1}\},&\{\mathcal{K}_{3},P_{2}\},&\{\mathcal{K}_{3},P_{3}\},&\{\mathcal{K}_{3},P_{4}\},\\ \{L_{2},P_{1}\},&\{L_{2},P_{3}\},&\{L_{3},P_{2}\},&\{L_{3},P_{4}\},\\ \{\mathcal{K}_{1},\mathcal{K}_{2},P_{1}\},&\{\mathcal{K}_{1},\mathcal{K}_{2},P_{2}\},&\{\mathcal{K}_{1},\mathcal{K}_{2},P_{3}\},&\{\mathcal{K}_{1},\mathcal{K}_{2},P_{4}\},\\ \{\mathcal{K}_{1},L_{2},P_{1}\},&\{\mathcal{K}_{1},L_{2},P_{3}\},&\{\mathcal{K}_{1},L_{3},P_{2}\},&\{\mathcal{K}_{1},L_{3},P_{4}\},\\ \{\mathcal{K}_{2},\mathcal{K}_{3},P_{1}\},&\{\mathcal{K}_{2},\mathcal{K}_{3},P_{2}\},&\{\mathcal{K}_{2},\mathcal{K}_{3},P_{3}\},&\{\mathcal{K}_{2},\mathcal{K}_{3},P_{4}\},\\ \{\mathcal{K}_{3},L_{2},P_{1}\},&\{\mathcal{K}_{3},L_{2},P_{3}\},&\{\mathcal{K}_{3},L_{3},P_{2}\},&\{\mathcal{K}_{3},L_{3},P_{4}\}.\\ \end{array}

Let 𝒢={G1,G2,,Gm}\mathcal{G}=\{G_{1},G_{2},\ldots,G_{m}\} be a well-connected building set and let 𝒮\mathcal{S} be a 𝒢\mathcal{G}-nested set. Given A𝒮A\in\mathcal{S}, we define 𝒮A:={B𝒮AB}\mathcal{S}_{A}:=\{B\in\mathcal{S}\mid A\subsetneq B\} and for every A𝒮A\in\mathcal{S} we denote by M𝒮(A)M_{\mathcal{S}}(A) the (connected) intersection 𝒮A\operatorname{\cap}\mathcal{S}_{A}. We will omit the nested set 𝒮\mathcal{S} and write just M(A)M(A) for brevity when it is clear from the context which is the involved nested set.

Definition 4.4.

A function f:𝒢f\colon\mathcal{G}\to\mathbb{N} is (𝒢\mathcal{G}-)admissible if it has both the following properties:

  1. (1)

    suppf\operatorname{supp}f is 𝒢\mathcal{G}-nested;

  2. (2)

    for every AsuppfA\in\operatorname{supp}f we have f(A)<dimMsuppf(A)dimAf(A)<\dim M_{\operatorname{supp}f}(A)-\dim A.

Notice that the zero function, i.e. the function such that f(A)=0f(A)=0 for every A𝒢A\in\mathcal{G}, is admissible since its support is the empty set.

Example 4.5 (Example 2.19, continued).

For each nested set 𝒮\mathcal{S} listed in Example 4.3, we test if it can be the support of an admissible function by computing the maximum value that the candidate function can assume on the elements of 𝒮\mathcal{S} (see Definition 4.4). It turns out that only 7 of the 48 nested sets give rise to admissible functions, namely

(4.2) ,{L2},{L3},{P1},{P2},{P3},{P4}.\emptyset,\ \{L_{2}\},\ \{L_{3}\},\ \{P_{1}\},\ \{P_{2}\},\ \{P_{3}\},\ \{P_{4}\}.

In particular we find 11 admissible functions:

Support Values
\emptyset f(G)=0f(G)=0 for all G𝒢G\in\mathcal{G}
{L2}\{L_{2}\} f(L2)=1f(L_{2})=1
{L3}\{L_{3}\} f(L3)=1f(L_{3})=1
{P1}\{P_{1}\} f(P1)=1f(P_{1})=1
{P1}\{P_{1}\} f(P1)=2f(P_{1})=2
{P2}\{P_{2}\} f(P2)=1f(P_{2})=1
{P2}\{P_{2}\} f(P2)=2f(P_{2})=2
{P3}\{P_{3}\} f(P3)=1f(P_{3})=1
{P3}\{P_{3}\} f(P3)=2f(P_{3})=2
{P4}\{P_{4}\} f(P4)=1f(P_{4})=1
{P4}\{P_{4}\} f(P4)=2f(P_{4})=2

Let us fix some notation. Given a nested set 𝒮\mathcal{S} we know that the intersection 𝒮\operatorname{\cap}\mathcal{S} is non-empty and it is a layer of type 𝒦(Γ(𝒮),ϕ𝒮)¯\overline{\mathcal{K}(\Gamma(\mathcal{S}),\phi_{\mathcal{S}})} for some Γ(𝒮)\Gamma(\mathcal{S}) and ϕ𝒮\phi_{\mathcal{S}}. Let H(𝒮)H(\mathcal{S}) be the subtorus associated with Γ(𝒮)\Gamma(\mathcal{S}) as in Theorem 2.7, namely H(𝒮)=χΓ(𝒮)ker(xχ)H(\mathcal{S})=\cap_{\chi\in\Gamma(\mathcal{S})}\ker(x_{\chi}). By Theorem 2.7, given a fan Δ\Delta, ΔH(𝒮)ΔVΓ(𝒮)\Delta_{H(\mathcal{S})}\coloneqq\Delta\cap V_{\Gamma(\mathcal{S})} is a smooth fan which we denote by Δ(𝒮)\Delta(\mathcal{S}). Let XΔ(𝒮)X_{\Delta(\mathcal{S})} be the corresponding toric variety and let π𝒮\pi_{\mathcal{S}} be the projection π𝒮:H(XΔ,)H(XΔ(𝒮),)\pi_{\mathcal{S}}\colon H^{*}(X_{\Delta},\mathbb{Z})\to H^{*}(X_{\Delta(\mathcal{S})},\mathbb{Z}), which is the restriction map induced by the inclusion, i.e. the one of (3.1). Let Θ(𝒮)\Theta(\mathcal{S}) be a minimal set of elements of H(XΔ,)H^{*}(X_{\Delta},\mathbb{Z}) such that their image via π𝒮\pi_{\mathcal{S}} is a basis of H(XΔ(𝒮),)H^{*}(X_{\Delta(\mathcal{S})},\mathbb{Z}).

Given a (non necessarily admissible) function f:𝒢f\colon\mathcal{G}\to\mathbb{N} we define the monomial in H(Y(XΔ,𝒢),)H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z}) viewed as R[T1,,Tm]/𝒢R[T_{1},\ldots,T_{m}]/\mathcal{I}_{\mathcal{G}}

mf=Gi𝒢Tif(Gi)m_{f}=\prod_{G_{i}\in\mathcal{G}}T_{i}^{f(G_{i})}

where R=H(XΔ,)R=H^{*}(X_{\Delta},\mathbb{Z}) and TiT_{i} is the (class of the) variable associated with GiG_{i}, and denote by 𝒢\mathcal{B}_{\mathcal{G}} the following set of elements of H(Y(XΔ,𝒢),)H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z}):

(4.3) 𝒢={bmff is admissible,bΘ(suppf)}.\mathcal{B}_{\mathcal{G}}=\{b\,m_{f}\mid f\text{ is admissible},b\in\Theta({\operatorname{supp}f})\}.
Remark 4.6.

For 𝒮=\mathcal{S}=\emptyset, the only admissible function is the zero function. In this case the associated monomial is 11, and Θ()\Theta(\emptyset) is a basis of H(XΔ,)H^{*}(X_{\Delta},\mathbb{Z}). In fact 𝒮=XΔ\operatorname{\cap}\mathcal{S}=X_{\Delta} by the usual convention, which is the closure of TT that, as a layer, has Γ={0}\Gamma=\{0\} and VΓ=VV_{\Gamma}=V. So Δ()=Δ\Delta(\emptyset)=\Delta and π\pi_{\emptyset} is the identity function. In particular 𝒢\mathcal{B}_{\mathcal{G}} contains the set {a1aΘ()}\{a\cdot 1\mid a\in\Theta(\emptyset)\}.

Theorem 4.7.

Let 𝒜\mathcal{A}, XΔX_{\Delta}, Λ\Lambda and 𝒢\mathcal{G} be as in the beginning of this section. The set 𝒢\mathcal{B}_{\mathcal{G}} defined as in (4.3) is a \mathbb{Z}-basis of H(Y(XΔ,𝒢),)H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z}).

Proof.

This proof is divided in two parts: we first show that the elements in 𝒢\mathcal{B}_{\mathcal{G}} generate H(Y(XΔ,𝒢),)H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z}) as a \mathbb{Z}-module; then we see that they are independent by counting them.

𝒢\mathcal{B}_{\mathcal{G}} generates H(Y(XΔ,𝒢),)H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z}) as a \mathbb{Z}-module. First of all notice that the relations in Theorem 3.6 imply that, given an admissible function ff and dH(XΔ,)d\in H^{*}(X_{\Delta},\mathbb{Z}) such that πsuppf(d)=0\pi_{\operatorname{supp}f}(d)=0 in H(XΔ(suppf),)H^{*}(X_{\Delta(\operatorname{supp}f)},\mathbb{Z}) then dmf=0d\cdot m_{f}=0 in H(Y(XΔ,𝒢),)H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z}). In fact dd is a polynomial in terms of variables CrC_{r} where rr does not belong to VΓ(suppf)V_{\Gamma(\operatorname{supp}f)}—this follows from Theorem 2.7. Therefore we can prove that 𝒢\mathcal{B}_{\mathcal{G}} generates H(Y(XΔ,𝒢),)H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z}) by showing that the set

𝒢={mff is admissible}\mathcal{M}_{\mathcal{G}}=\{m_{f}\mid f\text{ is admissible}\}

generates H(Y(XΔ,𝒢),)H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z}) as a H(XΔ,)H^{*}(X_{\Delta},\mathbb{Z})-module. To show this, let us consider a function g:𝒢g\colon\mathcal{G}\rightarrow\mathbb{N} which is not admissible: we will prove that mgm_{g} can be obtained as a H(XΔ,)H^{*}(X_{\Delta},\mathbb{Z})-linear combination of monomials in 𝒢\mathcal{M}_{\mathcal{G}}.

If suppg\operatorname{supp}g is non-nested then the intersection of the elements in suppg\operatorname{supp}g is empty by Theorem 2.2 of [11], so mg=0m_{g}=0 in H(Y(XΔ,𝒢),)H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z}). From now on we can assume that suppg\operatorname{supp}g is nested, and following the notation introduced before Definition 4.4 we will write M(B)=Msuppg(B)M(B)=M_{\operatorname{supp}g}(B).

Let us consider 𝒢{}\mathcal{G}\cup\{\emptyset\} as a set partially ordered by inclusion. On this poset we define a “pseudo-rank” function rr as follows: r()=0r(\emptyset)=0 and for G𝒢G\in\mathcal{G} let r(G)r(G) be the maximal length of a chain between \emptyset and GG in the Hasse diagram of the poset.

We say that BsuppgB\in\operatorname{supp}g is a bad component for gg if g(B)dimM(B)dimBg(B)\geq\dim M(B)-\dim B. To every bad component BB we assign the pair

(r(B),g(B)(dimM(B)dimB))×\big{(}r(B),g(B)-(\dim M(B)-\dim B)\big{)}\in\mathbb{N}\times\mathbb{N}

and we put on ×\mathbb{N}\times\mathbb{N} the lexicographic order.333i.e. (a,b)<(c,d)(a,b)<(c,d) if a<ca<c or if a=ca=c and b<db<d. Since we assumed that suppg\operatorname{supp}g is nested but that gg is not admissible, the set of bad components for gg is not empty, so we can define the evaluation of gg as the maximal pair associated with the bad components of gg. We will proceed by induction on the evaluation.

Base step. Let us consider a non-admissible function gg whose evaluation is (1,a)(1,a), for some aa\in\mathbb{N}. This means that the maximal bad components of gg are minimal elements in 𝒢\mathcal{G} w.r.t.  inclusion. Let BB be such a bad component. We can partition suppg\operatorname{supp}g as 𝒮𝒮{B}\mathcal{S}\cup\mathcal{S}^{\prime}\cup\{B\} where 𝒮:=(suppg)B\mathcal{S}:=(\operatorname{supp}g)_{B} and 𝒮:=suppg(𝒮{B})\mathcal{S}^{\prime}:=\operatorname{supp}g\setminus(\mathcal{S}\cup\{B\}). The monomial mgm_{g} associated with gg is of the following type:

mg=K𝒮TKg(K)K𝒮TKg(K)TBg(B).m_{g}=\prod_{K\in\mathcal{S}^{\prime}}T_{K}^{g(K)}\prod_{K\in\mathcal{S}}T_{K}^{g(K)}\cdot T_{B}^{g(B)}.

From Theorem 3.6 we know that in the ideal 𝒢\mathcal{I}_{\mathcal{G}} there is the element

F(B,𝒮)=PBM(TB)K𝒮TKF(B,\mathcal{S})=P^{M}_{B}(-T_{B})\prod_{K\in\mathcal{S}}T_{K}

where M=M𝒮(B)=Msuppg(B)=M(B)M=M_{\mathcal{S}}(B)=M_{\operatorname{supp}g}(B)=M(B).

Now we notice that PBM(TB)P^{M}_{B}(-T_{B}) is a polynomial in H(XΔ,)[TB]H^{*}(X_{\Delta},\mathbb{Z})[T_{B}] of the following form:

±(TB)dimMdimB+lower order terms in TB.\pm(T_{B})^{\dim M-\dim B}+\textrm{lower order terms in }T_{B}.

By writing mgm_{g} as

mg=(K𝒮TKg(K)K𝒮TKg(K)1TBg(B)(dimMdimB))(K𝒮TKTBdimMdimB),m_{g}=\Big{(}\prod_{K\in\mathcal{S}^{\prime}}T_{K}^{g(K)}\prod_{K\in\mathcal{S}}T_{K}^{g(K)-1}\cdot T_{B}^{g(B)-(\dim M-\dim B)}\Big{)}\Big{(}\prod_{K\in\mathcal{S}}T_{K}\cdot T_{B}^{\dim M-\dim B}\Big{)},

we see that the leading term of F(B,𝒮)F(B,\mathcal{S}), namely TBdimMdimBK𝒮TKT_{B}^{\dim M-\dim B}\prod_{K\in\mathcal{S}}T_{K}, divides mgm_{g}, so when we reduce mgm_{g} modulo 𝒢\mathcal{I}_{\mathcal{G}} we obtain a polynomial of the following form:

(K𝒮TKg(K)K𝒮TKg(K)1TBg(B)(dimMdimB))(K𝒮TKk=0dimMdimB1ckTBk)\displaystyle\Big{(}\prod_{K\in\mathcal{S}^{\prime}}T_{K}^{g(K)}\prod_{K\in\mathcal{S}}T_{K}^{g(K)-1}\cdot T_{B}^{g(B)-(\dim M-\dim B)}\Big{)}\Big{(}\prod_{K\in\mathcal{S}}T_{K}\cdot\sum_{k=0}^{\dim M-\dim B-1}c_{k}\cdot T_{B}^{k}\Big{)}
=\displaystyle= k=0dimMdimB1ck(Ksuppg{B}TKg(K))TBg(B)(dimMdimB)+k=k=0dimMdimB1ckmgk\displaystyle\sum_{k=0}^{\dim M-\dim B-1}c_{k}\left(\prod_{K\in\operatorname{supp}g\setminus\{B\}}T_{K}^{g(K)}\right)\cdot T_{B}^{g(B)-(\dim M-\dim B)+k}=\sum_{k=0}^{\dim M-\dim B-1}c_{k}m_{g_{k}}

for some ckH(XΔ,)c_{k}\in H^{*}(X_{\Delta},\mathbb{Z}) and some suitable gk:𝒢g_{k}\colon\mathcal{G}\to\mathbb{N}. Notice that suppgk=suppg\operatorname{supp}g_{k}=\operatorname{supp}g (eventually without BB) and that the gkg_{k}’s coincide with gg in suppg{B}\operatorname{supp}g\setminus\{B\}, whereas

gk(B)=g(B)(dimMdimB)+k.g_{k}(B)=g(B)-(\dim M-\dim B)+k.

Therefore every monomial mgkm_{g_{k}} appearing in this formula either is in 𝒢\mathcal{B}_{\mathcal{G}} or its evaluation is (1,b)(1,b) with b=a+k(dimMdimB)<ab=a+k-(\dim M-\dim B)<a. We can apply the same argument to the latter monomials until we get a linear combination of monomials in 𝒢\mathcal{B}_{\mathcal{G}}.

Inductive step. Let us suppose that our claim is true for non-admissible functions (with nested support) whose evaluation is (k,c)(k,c) with k>1k>1 and cc\in\mathbb{N}. Let us consider a non-admissible function gg with evaluation (k+1,a)(k+1,a). This means that there is at least one bad component BB whose associated pair is (k+1,a)(k+1,a).

As before we consider the element of 𝒢\mathcal{I}_{\mathcal{G}}:

F(B,𝒮)=PBM(D𝒢,DBTD)K𝒮TKF(B,\mathcal{S})=P^{M}_{B}(\sum_{D\in\mathcal{G},D\subseteq B}-T_{D})\prod_{K\in\mathcal{S}}T_{K}

where 𝒮=(suppg)B\mathcal{S}=(\operatorname{supp}g)_{B} and M=M𝒮(B)=Msuppg(B)=M(B)M=M_{\mathcal{S}}(B)=M_{\operatorname{supp}g}(B)=M(B). The polynomial PBM(TD)P^{M}_{B}(\sum-T_{D}) is of type

±(TB)dimMdimB+q,\pm(T_{B})^{\dim M-\dim B}+q,

where qq is a polynomial in H(XΔ,)[TDD𝒢,DB]H^{*}(X_{\Delta},\mathbb{Z})[T_{D}\mid D\in\mathcal{G},D\subseteq B] with degree in TBT_{B} strictly less than dimMdimB\dim M-\dim B.

As in the base step, we notice that the leading term of F(B,𝒮)F(B,\mathcal{S}) divides mgm_{g} and this allows us to write mgm_{g} modulo 𝒢\mathcal{I}_{\mathcal{G}} as a H(XΔ,)H^{*}(X_{\Delta},\mathbb{Z})-linear combination of monomials. If these are not 0 modulo 𝒢\mathcal{I}_{\mathcal{G}} either they are in 𝒢\mathcal{B}_{\mathcal{G}} or their evaluation is strictly less than (k+1,a)(k+1,a). If some of them have evaluation (k+1,b)(k+1,b) with b<ab<a we can again use the relations in 𝒢\mathcal{I}_{\mathcal{G}} and in a finite number of steps we get a H(XΔ,)H^{*}(X_{\Delta},\mathbb{Z})-linear combination of monomials that are in 𝒢\mathcal{B}_{\mathcal{G}} or have evaluation (s,t)(s,t) with sks\leq k. To these latter monomials we can apply the inductive hypothesis.

𝒢\mathcal{B}_{\mathcal{G}} is a set of independent elements. To show that the elements in 𝒢\mathcal{B}_{\mathcal{G}} are linearly independent over \mathbb{Z} it suffices to show that |𝒢|\left|\mathcal{B}_{\mathcal{G}}\right| is equal to the rank of H(Y(XΔ,𝒢),)H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z}) which is a free \mathbb{Z}-module (see Theorem 3.6). We proceed by induction on mm, the cardinality of 𝒢{\mathcal{G}}.

Base step. If m=1m=1 then 𝒢={G1}\mathcal{G}=\{G_{1}\} and Y(XΔ,𝒢)Y(X_{\Delta},\mathcal{G}) is the blowup of XΔX_{\Delta} along G1G_{1}. As it is well known (see for example [32, Chapter 4, Section 6]) we have the following isomorphism of graded \mathbb{Z}-modules:

H(Y(XΔ,𝒢),)H(XΔ,)J=1codimG11H(G1,)ζJH^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z})\cong H^{*}(X_{\Delta},\mathbb{Z})\oplus\bigoplus_{J=1}^{\operatorname{codim}G_{1}-1}H^{*}(G_{1},\mathbb{Z})\zeta^{J}

where ζJ\zeta^{J} is a symbol that shifts the degrees of +2J+2J. We now split 𝒢\mathcal{B}_{\mathcal{G}} into two disjoint subsets:

𝒢1:={a1aΘ()}\mathcal{B}_{\mathcal{G}}^{1}:=\{a\cdot 1\mid a\in\Theta({\emptyset})\}
𝒢2:={bT1rr=1,,codimG11 and bΘ({G1})}\mathcal{B}_{\mathcal{G}}^{2}:=\{b\,T_{1}^{r}\mid r=1,\ldots,\operatorname{codim}G_{1}-1\text{ and }b\in\Theta(\{G_{1}\})\}

We observe that by Theorem 2.7, point 3, G1G_{1} is isomorphic to XΔ({G1})X_{\Delta(\{G_{1}\})}. Therefore there is a grade-preserving bijection between 𝒢2\mathcal{B}_{\mathcal{G}}^{2} and a basis of H(G1,)ζJ\bigoplus H^{*}(G_{1},\mathbb{Z})\zeta^{J} which, together with the known bijection between 𝒢1\mathcal{B}_{\mathcal{G}}^{1} and a basis of H(XΔ,)H^{*}(X_{\Delta},\mathbb{Z}), proves our claim in this case.

Inductive step. We assume that the claim holds for every toric model associated with a building set of cardinality less than or equal to m1m-1. Consider 𝒢={G1,,Gm}\mathcal{G}=\{G_{1},\ldots,G_{m}\} and assume that the labelling is a refinement of the ordering in 𝒢\mathcal{G} by inclusion. We put 𝒢i:={G1,,Gi}\mathcal{G}_{i}:=\{G_{1},\ldots,G_{i}\} for every imi\leq m and Z:=GmZ:=G_{m}; we denote then t(Z)t(Z) the proper transform of ZZ in the variety Y(XΔ,𝒢m1)Y(X_{\Delta},\mathcal{G}_{m-1}). Then Y(XΔ,𝒢)Y(X_{\Delta},\mathcal{G}) is obtained as the blow up of Y(XΔ,𝒢m1)Y(X_{\Delta},\mathcal{G}_{m-1}) along t(Z)t(Z).

We can use again the result from [32] and obtain the following graded isomorphism of \mathbb{Z}-modules:

H(Y(XΔ,𝒢),)H(Y(XΔ,𝒢m1),)J=1codimGm1H(t(Z),)ζJ.H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z})\cong H^{*}(Y(X_{\Delta},\mathcal{G}_{m-1}),\mathbb{Z})\oplus\bigoplus_{J=1}^{\operatorname{codim}G_{m}-1}H^{*}(t(Z),\mathbb{Z})\zeta^{J}.

Following the idea from the base step, we write 𝒢\mathcal{B}_{\mathcal{G}} as the union of the disjoint sets:

𝒢1\displaystyle\mathcal{B}_{\mathcal{G}}^{1} ={bmff admissible, f(Gm)=0,bΘ(suppf)},\displaystyle{}=\{b\,m_{f}\mid f\text{ admissible, }f(G_{m})=0,\ b\in\Theta(\operatorname{supp}f)\},
𝒢2\displaystyle\mathcal{B}_{\mathcal{G}}^{2} ={bmff admissible, f(Gm)0,bΘ(suppf)}.\displaystyle{}=\{b\,m_{f}\mid f\text{ admissible, }f(G_{m})\neq 0,\ b\in\Theta(\operatorname{supp}f)\}.

There is a bijective correspondence, provided by the restriction, between the set {f:𝒢f admissible, f(Gm)=0}\{f\colon\mathcal{G}\to\mathbb{N}\mid f\text{ admissible, }f(G_{m})=0\} and the set {f:𝒢m1f admissible}\{f\colon\mathcal{G}_{m-1}\to\mathbb{N}\mid f\text{ admissible}\}. By the inductive hypothesis, 𝒢1\mathcal{B}_{\mathcal{G}}^{1} is in bijection with the basis 𝒢m1\mathcal{B}_{\mathcal{G}_{m-1}} of H(Y(XΔ,𝒢m1),)H^{*}(Y(X_{\Delta},\mathcal{G}_{m-1}),\mathbb{Z}) and this correspondence is grade-preserving.

Notice that because GmG_{m} is maximal in 𝒢\mathcal{G}, then GmG_{m} is maximal in suppf\operatorname{supp}f for every admissible function ff such that f(Gm)0f(G_{m})\neq 0. So the possible values for f(Gm)f(G_{m}) are 1,,codimGm11,\ldots,\operatorname{codim}G_{m}-1.

Now we observe that, given an element bmf𝒢2b\,m_{f}\in\mathcal{B}_{\mathcal{G}}^{2} (so that f(Gm)0f(G_{m})\neq 0), we also have in 𝒢2\mathcal{B}_{\mathcal{G}}^{2} the monomials bmgb\,m_{g} for all the admissible functions gg that coincide with ff on 𝒢{Gm}\mathcal{G}\setminus\{G_{m}\} and such that g(Gm){1,,codimGm1}{f(Gm)}g(G_{m})\in\{1,\ldots,\operatorname{codim}G_{m}-1\}\setminus\{f(G_{m})\}. As a consequence the sets {bmf𝒢2f(Gm)=i}\{b\,m_{f}\in\mathcal{B}_{\mathcal{G}}^{2}\mid f(G_{m})=i\}, for i=1,,codimGm1i=1,\ldots,\operatorname{codim}G_{m}-1 have all the same cardinality and form a partition of 𝒢2\mathcal{B}_{\mathcal{G}}^{2}, and it suffices to prove that there is a grade-preserving (up to a shift by 2 in cohomology) bijection between {bmf𝒢2f(Gm)=1}\{b\,m_{f}\in\mathcal{B}_{\mathcal{G}}^{2}\mid f(G_{m})=1\} and a basis of H(t(Z),)H^{*}(t(Z),\mathbb{Z}). This extends to a grade-preserving bijection between 𝒢2\mathcal{B}_{\mathcal{G}}^{2} and a basis of H(t(Z),)ζJ\bigoplus H^{*}(t(Z),\mathbb{Z})\zeta^{J}.

Let us now recall the following result and notation from [12, Section 4]. We consider the family \mathcal{H} of subvarieties in ZZ that are the connected components of the intersections GiZG_{i}\cap Z for every i=1,,m1i=1,\ldots,m-1. Since 𝒢\mathcal{G} is well-connected, if GiZG_{i}\cap Z is not empty and not connected then its connected components belong to 𝒢\mathcal{G}. This implies that u:=||m1u:=\left|\mathcal{H}\right|\leq m-1. Now for each HH\in\mathcal{H}, we denote by s(H)s(H) the minimum index ii such that HH is a connected component of GiZG_{i}\cap Z (in particular H=Gs(H)ZH=G_{s(H)}\cap Z). We sort the set {s(H)H}\{s(H)\mid H\in\mathcal{H}\} in ascending order as {s1,,su}\{s_{1},\ldots,s_{u}\} and let ={H1,,Hu}\mathcal{H}=\{H_{1},\ldots,H_{u}\}.

Remark 4.8.

Notice that two possibility occurs for HH\in\mathcal{H}: either H=Gs(H)H=G_{s(H)} in case Gs(H)ZG_{s(H)}\subset Z, or H=Gs(H)ZH=G_{s(H)}\cap Z and the intersection is transversal.

In [12, Proposition 4.4] it is proven that \mathcal{H} is building and well-connected and from Proposition 4.6 of the same paper it follows that t(Z)t(Z) is isomorphic to the model Y(Z,)Y(Z,\mathcal{H}) obtained by blowing up \mathcal{H} in ZZ. From Theorem 2.7, point 2, we know that Z=GmZ=G_{m} is a toric variety with fan Δ=Δ({Gm})\Delta^{\prime}=\Delta(\{G_{m}\}), so Z=XΔZ=X_{\Delta^{\prime}}.

In analogy with the previous notation, for an \mathcal{H}-nested 𝒮\mathcal{S} let π𝒮\pi^{\prime}_{\mathcal{S}} be the projection π𝒮:H(XΔ,)H(XΔ(𝒮),)\pi^{\prime}_{\mathcal{S}}\colon H^{*}(X_{\Delta^{\prime}},\mathbb{Z})\to H^{*}(X_{\Delta^{\prime}(\mathcal{S})},\mathbb{Z}) and we take Θ(𝒮)\Theta^{\prime}(\mathcal{S}) as a minimal set of elements of H(XΔ,)H^{*}(X_{\Delta^{\prime}},\mathbb{Z}) such that their image via π𝒮\pi^{\prime}_{\mathcal{S}} is a basis of H(XΔ(𝒮),)H^{*}(X_{\Delta^{\prime}(\mathcal{S})},\mathbb{Z}).

Since ||<m\left|\mathcal{H}\right|<m we can apply our inductive hypothesis to Y(Z,)Y(Z,\mathcal{H}) and we get the following \mathbb{Z}-basis \mathcal{B}_{\mathcal{H}} of H(Y(Z,),)H^{*}(Y(Z,\mathcal{H}),\mathbb{Z}):

={bmgg-admissible,bΘ(suppg)}.\mathcal{B}_{\mathcal{H}}=\{b^{\prime}\,m_{g}\mid\ g\ \mathcal{H}\text{-admissible},\ b^{\prime}\in\Theta^{\prime}(\operatorname{supp}g)\}.

We are now ready to describe a bijective, grade-preserving (up to a shift by 2 in cohomology) correspondence between \mathcal{B}_{\mathcal{H}} and

L:={bmff𝒢-admissible,f(Gm)=1,bΘ(suppf)}𝒢2.L:=\{b\,m_{f}\mid f\ \mathcal{G}\text{-admissible},\ f(G_{m})=1,\ b\in\Theta(\operatorname{supp}f)\}\subset\mathcal{B}_{\mathcal{G}}^{2}.

Let f:𝒢f\colon\mathcal{G}\to\mathbb{N} be a 𝒢\mathcal{G}-admissible function with f(Gm)=1f(G_{m})=1. We associate to ff the function f¯:\overline{f}\colon\mathcal{H}\to\mathbb{N} such that f¯(H)=f(Gs(H))\overline{f}(H)=f(G_{s(H)}) for all HH\in\mathcal{H}.

Lemma 4.9.

The function f¯\overline{f} is \mathcal{H}-admissible.

Proof.

If suppf¯\operatorname{supp}\overline{f} is empty, f¯\overline{f} is admissible by definition, so from now we suppose that suppf¯\operatorname{supp}\overline{f}\neq\emptyset.

We show that suppf¯\operatorname{supp}\overline{f} is \mathcal{H}-nested by using the characterization from Proposition 4.1: we take an antichain 𝒦={K1,,Kt}\mathcal{K}=\{K_{1},\ldots,K_{t}\} (t2t\geq 2) in suppf¯\operatorname{supp}\overline{f} and prove that the intersection 𝒦\operatorname{\cap}\mathcal{K} is connected, transversal and does not belong to \mathcal{H}.

𝒦\operatorname{\cap}\mathcal{K} is connected. It is equal to Gs(K1)Gs(Kt)GmG_{s(K_{1})}\cap\dotsb\cap G_{s(K_{t})}\cap G_{m} and they all belong to suppf\operatorname{supp}f which is 𝒢\mathcal{G}-nested. In particular, their intersection is the intersection of the minimal elements among them, that is to say, the intersection of an antichain of a nested set, which is connected by Proposition 4.1.

𝒦\operatorname{\cap}\mathcal{K} is transversal. We split this part of the proof in two cases:

  1. (1)

    if there is a j{1,,t}j\in\{1,\ldots,t\} such that Kj=Gs(Kj)K_{j}=G_{s(K_{j})}, then

    K1KjKt=GmGs(K1)Gs(Kt)=Gs(K1)Gs(Kt)K_{1}\cap\dotsb\cap K_{j}\cap\dotsb\cap K_{t}=G_{m}\cap G_{s(K_{1})}\cap\dotsb\cap G_{s(K_{t})}=G_{s(K_{1})}\cap\dotsb\cap G_{s(K_{t})}

    which is transversal because {Gs(K1),,Gs(Kt)}\{G_{s(K_{1})},\ldots,G_{s(K_{t})}\} is a nested set;

  2. (2)

    if Kj=GmGs(Kj)K_{j}=G_{m}\cap G_{s(K_{j})} transversally for every j{1,,t}j\in\{1,\ldots,t\}, then the set {Gs(K1),,Gs(Kt),Gm}\{G_{s(K_{1})},\ldots,G_{s(K_{t})},G_{m}\} is a set of elements which are pairwise non-comparable (the Gs(Ki)G_{s(K_{i})}’s are not pairwise comparable because the KiK_{i}’s are not), so their intersection

    GmGs(K1)Gs(Kt)=𝒦G_{m}\cap G_{s(K_{1})}\cap\dotsb\cap G_{s(K_{t})}=\operatorname{\cap}\mathcal{K}

    is transversal.

𝒦\operatorname{\cap}\mathcal{K} does not belong to \mathcal{H}. Suppose that 𝒦=H\operatorname{\cap}\mathcal{K}=H\in\mathcal{H}. It is not possible that H=Gs(H)H=G_{s(H)}, because otherwise 𝒦\operatorname{\cap}\mathcal{K} would belong to 𝒢\mathcal{G} in contradiction with Proposition 4.1. On the other hand, if H=Gs(H)GmH=G_{s(H)}\cap G_{m} transversally, as a consequence of [12, Proposition 3.3] we have that Gs(H)G_{s(H)} and GmG_{m} are the minimal elements among the ones in 𝒢\mathcal{G} containing HH. We study the following two subcases:

  • if there is a j{1,,t}j\in\{1,\ldots,t\} such that Kj=Gs(Kj)K_{j}=G_{s(K_{j})}, then we would have HGs(Kj)GmH\subset G_{s(K_{j})}\subset G_{m} in contradiction with the minimality of GmG_{m};

  • if Kj=GmGs(Kj)K_{j}=G_{m}\cap G_{s(K_{j})} transversally for every j{1,,t}j\in\{1,\ldots,t\}, since k2k\geq 2 we would have two different 𝒢\mathcal{G}-decompositions of HH, namely

    H=GmGs(Kj)=GmGs(K1)Gs(Kt).H=G_{m}\cap G_{s(K_{j})}=G_{m}\cap G_{s(K_{1})}\cap\dotsb\cap G_{s(K_{t})}.

This concludes the proof that suppf¯\operatorname{supp}\overline{f} is nested.

Now let us consider Hsuppf¯H\in\operatorname{supp}\overline{f}. We study the following two cases:

Case H=Gs(H)H=G_{s(H)}. In this case f¯(H)=f(Gs(H))\overline{f}(H)=f(G_{s(H)}) and, since ff is 𝒢\mathcal{G}-admissible, we have

1f(Gs(H))<dimMsuppf(Gs(H))dimGs(H).1\leq f(G_{s(H)})<\dim M_{\operatorname{supp}f}(G_{s(H)})-\dim G_{s(H)}.

Now, because Gm{BsuppfGs(H)B}G_{m}\in\{B\in\operatorname{supp}f\mid G_{s(H)}\subsetneq B\} we notice that

(suppf)Gs(H)=(suppf¯)H\operatorname{\cap}(\operatorname{supp}f)_{G_{s(H)}}=\operatorname{\cap}(\operatorname{supp}\overline{f})_{H}

where suppf¯\operatorname{supp}\overline{f} is viewed as a \mathcal{H}-nested set and with the usual convention that if {Hisuppf¯HHi}\{H_{i}\in\operatorname{supp}\overline{f}\mid H\subsetneq H_{i}\} is empty the intersection is the ambient space GmG_{m}. In particular

dim(Msuppf(Gs(H)))=dim(Msuppf¯(H)),\dim\big{(}M_{\operatorname{supp}f}(G_{s(H)})\big{)}=\dim\big{(}M_{\operatorname{supp}\overline{f}}(H)\big{)},

therefore f¯(H)\overline{f}(H) ranges in the expected interval.

Case H=Gs(H)GmH=G_{s(H)}\cap G_{m} transversally. In this case f¯(H)=f(Gs(H))\overline{f}(H)=f(G_{s(H)}) where again

1f(Gs(H))<dimMsuppf(Gs(H))dimGs(H).1\leq f(G_{s(H)})<\dim M_{\operatorname{supp}f}(G_{s(H)})-\dim G_{s(H)}.

Now since suppf\operatorname{supp}f is 𝒢\mathcal{G}-nested and GmG_{m} does not belong to {BsuppfGs(H)B}\{B\in\operatorname{supp}f\mid G_{s(H)}\subsetneq B\} we have that

(4.4) dim(Msuppf(Gs(H))Gm)=dim(Msuppf(Gs(H)))codimGm.\dim\left(M_{\operatorname{supp}f}(G_{s(H)})\cap G_{m}\right)=\dim\left(M_{\operatorname{supp}f}(G_{s(H)})\right)-\operatorname{codim}G_{m}.

But

((suppf)Gs(H))Gm=(suppf¯)H\Big{(}\operatorname{\cap}(\operatorname{supp}f)_{G_{s(H)}}\Big{)}\cap G_{m}=\operatorname{\cap}(\operatorname{supp}\overline{f})_{H}

so we can rewrite (4.4) as

dimMsuppf¯(H)=dim(Msuppf(Gs(H)))codimGm\dim M_{\operatorname{supp}\overline{f}}(H)=\dim\left(M_{\operatorname{supp}f}(G_{s(H)})\right)-\operatorname{codim}G_{m}

and, observing that dimH=dimGs(H)codimGm\dim H=\dim G_{s(H)}-\operatorname{codim}G_{m}, we conclude that

dimMsuppf(Gs(H))dimGs(H)\displaystyle\dim M_{\operatorname{supp}f}(G_{s(H)})-\dim G_{s(H)} =dimMsuppf¯(H)+codimGmdimGs(H)\displaystyle{}=\dim M_{\operatorname{supp}\overline{f}}(H)+\operatorname{codim}G_{m}-\dim G_{s(H)}
=dimMsuppf¯(H)dimH\displaystyle{}=\dim M_{\operatorname{supp}\overline{f}}(H)-\dim H

therefore also in this case f¯(H)\overline{f}(H) ranges in the expected interval. ∎

Lemma 4.10.

If ff is 𝒢\mathcal{G}-admissible and f(Gm)=1f(G_{m})=1, then suppf{Gm}{Gs(H)H}\operatorname{supp}f\setminus\{G_{m}\}\subseteq\{G_{s(H)}\mid H\in\mathcal{H}\}.

Proof.

Let us suppose B=Gksuppf{Gm}B=G_{k}\in\operatorname{supp}f\setminus\{G_{m}\} with ks(H)k\neq s(H) for every HH\in\mathcal{H}. We notice that BGmB\nsubseteq G_{m} otherwise B=BGmB=B\cap G_{m}\in\mathcal{H} so B=Gs(B)B=G_{s(B)}. Moreover, since suppf\operatorname{supp}f is nested and contains both BB and GmG_{m} it follows that BGmB\cap G_{m}\neq\emptyset and connected by Proposition 4.1. Now we observe that from the connectedness of BGmB\cap G_{m} and the definition of \mathcal{H} we have that BGm=HjB\cap G_{m}=H_{j} for some jj and that BB and GmG_{m} are its 𝒢\mathcal{G}-factors. But Hj=GsjGmH_{j}=G_{s_{j}}\cap G_{m} is a different 𝒢\mathcal{G}-factorization of HjH_{j}, obtaining a contradiction. ∎

Therefore, given ff 𝒢\mathcal{G}-admissible with f(Gm)=1f(G_{m})=1 we can associate two monomials: mfm_{f} and mf¯m_{\overline{f}}. Now in 𝒢\mathcal{B}_{\mathcal{G}} we find elements of the form bmfb\,m_{f} with bb belonging to Θ(suppf)\Theta(\operatorname{supp}f); on the other hand in \mathcal{B}_{\mathcal{H}} we find elements of the form bmf¯b^{\prime}\,m_{\overline{f}} with bb^{\prime} belonging to Θ(suppf¯)\Theta^{\prime}(\operatorname{supp}\overline{f}). But Δ(suppf¯)=Δ(suppf)\Delta^{\prime}(\operatorname{supp}\overline{f})=\Delta(\operatorname{supp}f), therefore we can choose bb and bb^{\prime} above so that they range over the same set.

We have thus constructed a map Φ:L\Phi\colon L\to\mathcal{B}_{\mathcal{H}} such that Φ(bmf)=bmf¯\Phi(b\,m_{f})=b\,m_{\overline{f}}. If we show that Φ\Phi is a bijection, this concludes the proof of the theorem. Actually it is sufficient to prove that Φ\Phi is injective: in fact the injectivity implies

|L|||=rk(H(t(Z),))\left|L\right|\leq\left|\mathcal{B}_{\mathcal{H}}\right|=\operatorname{rk}(H^{*}(t(Z),\mathbb{Z}))

where the last equality, as we have seen, derives from the inductive hypothesis, since \mathcal{B}_{\mathcal{H}} is a basis of H(t(Z),)H^{*}(t(Z),\mathbb{Z}). This in turn implies that

|𝒢|\displaystyle\left|\mathcal{B}_{\mathcal{G}}\right| =|𝒢1|+|𝒢2|=|𝒢1|+(codimGm1)|L|=\displaystyle{}=\left|\mathcal{B}_{\mathcal{G}}^{1}\right|+\left|\mathcal{B}_{\mathcal{G}}^{2}\right|=\left|\mathcal{B}_{\mathcal{G}}^{1}\right|+(\operatorname{codim}G_{m}-1)\left|L\right|=
=rkH(Y(XΔ,𝒢m1),)+(codimGm1)|L|\displaystyle{}=\operatorname{rk}H^{*}(Y(X_{\Delta},\mathcal{G}_{m-1}),\mathbb{Z})+(\operatorname{codim}G_{m}-1)\left|L\right|
rkH(Y(XΔ,𝒢m1),)+(codimGm1)rk(H(t(Z),))=rkH(Y(XΔ,𝒢),).\displaystyle{}\leq\operatorname{rk}H^{*}(Y(X_{\Delta},\mathcal{G}_{m-1}),\mathbb{Z})+(\operatorname{codim}G_{m}-1)\operatorname{rk}(H^{*}(t(Z),\mathbb{Z}))=\operatorname{rk}H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z}).

On the other hand we already know, from the first part of this proof, that 𝒢\mathcal{B}_{\mathcal{G}} generates H(Y(XΔ,𝒢),)H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z}). It follows that

|𝒢|=rkH(Y(XΔ,𝒢),)\left|\mathcal{B}_{\mathcal{G}}\right|=\operatorname{rk}H^{*}(Y(X_{\Delta},\mathcal{G}),\mathbb{Z})

which is the claim of the theorem (and of course this also implies that Φ\Phi is actually a bijection).

To show the injectivity of Φ\Phi let f1f_{1}, f2f_{2} be two distinct 𝒢\mathcal{G}-admissible functions with f1(Gm)=f2(Gm)=1f_{1}(G_{m})=f_{2}(G_{m})=1: we prove that f1¯f2¯\overline{f_{1}}\neq\overline{f_{2}}.

Let us first suppose that suppf1suppf2\operatorname{supp}f_{1}\neq\operatorname{supp}f_{2}; up to switching f1f_{1} and f2f_{2}, we can assume that there exists Bsuppf1suppf2B\in\operatorname{supp}f_{1}\setminus\operatorname{supp}f_{2}. By Lemma 4.10 we know that B=Gs(H)B=G_{s(H)} for a certain HH\in\mathcal{H}. We deduce that Hsuppf1¯suppf2¯H\in\operatorname{supp}\overline{f_{1}}\setminus\operatorname{supp}\overline{f_{2}} and conclude that f1¯f2¯\overline{f_{1}}\neq\overline{f_{2}}.

If instead suppf1=suppf2\operatorname{supp}f_{1}=\operatorname{supp}f_{2}, f1f2f_{1}\neq f_{2} implies that there is a certain BB in their support such that f1(B)f2(B)f_{1}(B)\neq f_{2}(B). Again by Lemma 4.10 we know that B=Gs(H)B=G_{s(H)}, HH\in\mathcal{H}, therefore f1¯(H)=f1(Gs(H))f2(Gs(H))=f2¯(H)\overline{f_{1}}(H)=f_{1}(G_{s(H)})\neq f_{2}(G_{s(H)})=\overline{f_{2}}(H). This proves that Φ\Phi is injective and concludes the proof of the theorem. ∎

Example 4.11 (Example 2.19, continued).

We can compute a \mathbb{Z}-basis for the ring H(Y𝒜,)H^{*}(Y_{\mathcal{A}},\mathbb{Z}) using the admissible functions found in Example 4.5. The result is detailed in Tables 1 and 2; the tables have one line for each possible support 𝒮\mathcal{S} of admissible functions, as listed in (4.2).

Table 1. Data used to build the basis of H(Y𝒜,)H^{*}(Y_{\mathcal{A}},\mathbb{Z}).
𝒮\mathcal{S} Basis for Monomials mfm_{f}
H(XΔ(𝒮),)H^{*}(X_{\Delta(\mathcal{S})},\mathbb{Z}) with ff s.t. suppf=𝒮\operatorname{supp}f=\mathcal{S}
\emptyset Basis of H(XΔ,)H^{*}(X_{\Delta},\mathbb{Z}) {1}\{1\}
{L2}\{L_{2}\} {C7,1}\{C_{7},1\} {T4}\{T_{4}\}
{L3}\{L_{3}\} {C7,1}\{C_{7},1\} {T5}\{T_{5}\}
{P1}\{P_{1}\} {1}\{1\} {T6,T62}\{T_{6},T_{6}^{2}\}
{P2}\{P_{2}\} {1}\{1\} {T8,T82}\{T_{8},T_{8}^{2}\}
{P3}\{P_{3}\} {1}\{1\} {T7,T72}\{T_{7},T_{7}^{2}\}
{P4}\{P_{4}\} {1}\{1\} {T9,T92}\{T_{9},T_{9}^{2}\}
Table 2. Contribution to the basis of H(Y𝒜,)H^{*}(Y_{\mathcal{A}},\mathbb{Z}) and to the Betti numbers of Y𝒜Y_{\mathcal{A}}.
𝒮\mathcal{S} Contribution to Contribution to rk(Hi(Y𝒜,))\operatorname{rk}(H^{i}(Y_{\mathcal{A}},\mathbb{Z}))
the basis 𝒢\mathcal{B}_{\mathcal{G}} i=0i=0 i=2i=2 i=4i=4 i=6i=6
\emptyset Basis of H(XΔ,)H^{*}(X_{\Delta},\mathbb{Z}) 1 69 69 1
{L2}\{L_{2}\} {C7T4,T4}\{C_{7}T_{4},T_{4}\} 0 1 1 0
{L3}\{L_{3}\} {C7T5,T5}\{C_{7}T_{5},T_{5}\} 0 1 1 0
{P1}\{P_{1}\} {T6,T62}\{T_{6},T_{6}^{2}\} 0 1 1 0
{P2}\{P_{2}\} {T8,T82}\{T_{8},T_{8}^{2}\} 0 1 1 0
{P3}\{P_{3}\} {T7,T72}\{T_{7},T_{7}^{2}\} 0 1 1 0
{P4}\{P_{4}\} {T9,T92}\{T_{9},T_{9}^{2}\} 0 1 1 0
𝒢\mathcal{B}_{\mathcal{G}} 1 75 75 1
Example 4.12.

As we have seen in Example 4.5, not all the nested sets are supports of admissible functions. In particular, for small toric arrangements in low dimensions admissible functions often are supported only on singletons. In this example we study a case where there are supports of admissible functions with cardinality 2.

Let 𝒜={𝒦1,𝒦2,𝒦3}\mathcal{A}=\{\mathcal{K}_{1},\mathcal{K}_{2},\mathcal{K}_{3}\} be the arrangement in ()4(\mathbb{C}^{*})^{4}, with coordinates (x,y,z,t)(x,y,z,t), whose layers are defined by the equations

𝒦1:z=t=1,𝒦2:y=1,𝒦3:x=1.\mathcal{K}_{1}\colon z=t=1,\qquad\mathcal{K}_{2}\colon y=1,\qquad\mathcal{K}_{3}\colon x=1.

The poset of layers 𝒞(𝒜)\mathcal{C}(\mathcal{A}) is represented in Figure 2. In this example XΔ=(1)4X_{\Delta}=(\mathbb{P}^{1})^{4} is a good toric variety for the arrangement, so we can use its associated fan Δ\Delta (recall that its 16 maximal cones are C(σ1e1,σ2e2,σ3e3,σ4e4)C(\sigma_{1}e_{1},\sigma_{2}e_{2},\sigma_{3}e_{3},\sigma_{4}e_{4}) where e1,,e4e_{1},\dotsc,e_{4} are the vectors of the canonical basis of 4\mathbb{C}^{4} and (σ1,σ2,σ3,σ4){±1}4(\sigma_{1},\sigma_{2},\sigma_{3},\sigma_{4})\in\{\pm 1\}^{4}); moreover we choose 𝒢=𝒞0(𝒜)\mathcal{G}=\mathcal{C}_{0}(\mathcal{A}) and build the model Y𝒜=Y(XΔ,𝒢)Y_{\mathcal{A}}=Y(X_{\Delta},\mathcal{G}).

𝒦1\mathcal{K}_{1}TT𝒦1\mathcal{K}_{1}𝒦2\mathcal{K}_{2}𝒦3\mathcal{K}_{3}M1M_{1}M2M_{2}LLPP
𝒦1\mathcal{K}_{1} {}\rightsquigarrow{} T1T_{1}
𝒦2\mathcal{K}_{2} {}\rightsquigarrow{} T2T_{2}
𝒦3\mathcal{K}_{3} {}\rightsquigarrow{} T3T_{3}
M1M_{1} {}\rightsquigarrow{} T4T_{4}
M2M_{2} {}\rightsquigarrow{} T5T_{5}
LL {}\rightsquigarrow{} T6T_{6}
PP {}\rightsquigarrow{} T7T_{7}
Figure 2. Poset of layers 𝒞(𝒜)\mathcal{C}(\mathcal{A}) for the arrangement of Example 4.12. Elements at the same height have the same codimension. On the right: map from elements of 𝒢=𝒞0(𝒜)\mathcal{G}=\mathcal{C}_{0}(\mathcal{A}) to the corresponding variables TiT_{i} in H(XΔ,)[T1,,T7]H^{*}(X_{\Delta},\mathbb{Z})[T_{1},\dotsc,T_{7}].

The possible non-empty supports of admissible functions are

  • {L}\{L\}: it supports one admissible function ff such that f(L)=1f(L)=1;

  • {M1}\{M_{1}\}, {M2}\{M_{2}\}: each supports two admissible functions, namely f(Mi)=1f(M_{i})=1 and f(Mi)=2f(M_{i})=2;

  • {P}\{P\}: it supports three admissible functions, where f(P)f(P) is either 11, 22 or 33;

  • {L,P}\{L,P\}: it supports one admissible function such that f(L)=1f(L)=1 and f(P)=1f(P)=1.

Table 3 details the contribution to the Betti numbers for each admissible function.

Table 3. Contribution to the Betti numbers of H(Y𝒜,)H^{*}(Y_{\mathcal{A}},\mathbb{Z}) for each admissible function, grouped by support.

𝒮\mathcal{S} Betti numbers Monomials mfm_{f} with ff Contribution to rk(Hi(Y𝒜,))\operatorname{rk}(H^{i}(Y_{\mathcal{A}},\mathbb{Z})) for H(XΔ(𝒮),)H^{*}(X_{\Delta(\mathcal{S})},\mathbb{Z}) s.t. suppf=𝒮\operatorname{supp}f=\mathcal{S} i=0i=0 i=2i=2 i=4i=4 i=6i=6 i=8i=8 \emptyset 1, 4, 6, 4, 1 {1}\{1\} 1 4 6 4 1 {L}\{L\} 1, 2, 1 {T6}\{T_{6}\} 0 1 2 1 0 {M1}\{M_{1}\} 1, 1 {T4,T42}\{T_{4},T_{4}^{2}\} 0 1 2 1 0 {M2}\{M_{2}\} 1, 1 {T5,T52}\{T_{5},T_{5}^{2}\} 0 1 2 1 0 {P}\{P\} 1 {T7,T72,T73}\{T_{7},T_{7}^{2},T_{7}^{3}\} 0 1 1 1 0 {L,P}\{L,P\} 1 {T6T7}\{T_{6}T_{7}\} 0 0 1 0 0

5. The case of root systems of type AA

In this section we will apply our main theorem to the case of the toric arrangement associated with a root system of type An1A_{n-1}.

5.1. The minimal toric model and its cohomology basis

The toric analogue of the hyperplane arrangement of type An1A_{n-1} is 𝒜An1={𝒦ij1i<jn}\mathcal{A}_{A_{n-1}}=\{\mathcal{K}_{ij}\mid{1\leq i<j\leq n}\} in T=()n/Hn()n1T=(\mathbb{C}^{*})^{n}/H_{n}\simeq(\mathbb{C}^{*})^{n-1}, where HnH_{n} is the \mathbb{C}^{*}-span of (1,,1)(1,\dotsc,1) and

𝒦ij:={[t1,,tn]Ttitj1=1}.\mathcal{K}_{ij}:=\{[t_{1},\ldots,t_{n}]\in T\mid t_{i}t_{j}^{-1}=1\}.

Its poset of intersections 𝒞(𝒜An1)\mathcal{C}(\mathcal{A}_{A_{n-1}}) is isomorphic to the poset of partitions of the set {1,,n}\{1,\ldots,n\} ordered by refiniment. More precisely, the partition {I1,,Ik}\{I_{1},\ldots,I_{k}\} with I1Ik={1,,n}I_{1}\sqcup\dotsb\sqcup I_{k}=\{1,\ldots,n\} corresponds to the layer

𝒦I1,,Ik:={[t1,,tn]Tti=tj if l such that i,jIl},\mathcal{K}_{I_{1},\ldots,I_{k}}:=\{[t_{1},\ldots,t_{n}]\in T\mid t_{i}=t_{j}\text{ if }\exists\;l\text{ such that }i,j\in I_{l}\},

where for the sake of convenience we will sometimes omit to write the blocks IjI_{j} with cardinality one.

Let An1𝒞(𝒜An1)\mathcal{F}_{A_{n-1}}\subset\mathcal{C}(\mathcal{A}_{A_{n-1}}) be the set whose elements are the 𝒦I\mathcal{K}_{I} for every I{1,,n}I\subset\{1,\ldots,n\} with |I|2\left|I\right|\geq 2. It is a building set, in fact it is the minimal one that contains the layers 𝒦ij\mathcal{K}_{ij}; it is the analogue of the “building set of irreducible elements” for the linear case (see [16, 49, 30]).

For the arrangement 𝒜An1\mathcal{A}_{A_{n-1}} there is a natural choice of a fan that produces a good toric variety, as noted in [11]: we take in V=X(T)V=X_{*}(T)\otimes_{\mathbb{Z}}\mathbb{R} the fan ΔAn1\Delta_{A_{n-1}} induced by the Weyl chambers of the root system. By construction every layer of 𝒞(𝒜An1)\mathcal{C}(\mathcal{A}_{A_{n-1}}) has an equal sign basis with respect to ΔAn1\Delta_{A_{n-1}}, so the toric variety XΔAn1X_{\Delta_{A_{n-1}}} associated with ΔAn1\Delta_{A_{n-1}} is a good toric variety for the arrangement.

In [43] Procesi studies this toric variety (and also the more general toric varieties XWX_{W} associated with the fan induced by the Weyl chambers of a Weyl group WW; see also [22]). As it is well-known, the even Betti numbers of the toric variety XΔAn1X_{\Delta_{A_{n-1}}} are the Eulerian numbers A(n,k)A(n,k) (see, for example, [48, 46, 47]). We recall briefly the main definitions and results about the numbers A(n,k)A(n,k) and the cohomology of XΔAn1X_{\Delta_{A_{n-1}}}.

Definition 5.1.

The Eulerian number A(n,k+1)A(n,k+1) is the number of permutations in SnS_{n} with kk descents444If σ\sigma is a permutation in SnS_{n}, a descent of σ\sigma is an index i{1,,n1}i\in\{1,\dotsc,n-1\} such that σ(i)>σ(i+1)\sigma(i)>\sigma(i+1). The number of descents of σ\sigma is denoted by des(σ)\operatorname{des}(\sigma). for n1n\geq 1 and 0kn10\leq k\leq n-1.

Following [8] we present the Eulerian polynomial An(q)A_{n}(q) as

An(q)={k=1nA(n,k)qk,n1,1,n=0.A_{n}(q)=\begin{cases}{\displaystyle\sum_{k=1}^{n}A(n,k)q^{k},}&n\geq 1,\\ 1,&n=0.\end{cases}

According to the above formula one can compute the first Eulerian polynomials obtaining A1(q)=qA_{1}(q)=q, A2(q)=q+q2A_{2}(q)=q+q^{2}, A3(q)=q+4q2+q3A_{3}(q)=q+4q^{2}+q^{3}. The exponential generating function of the Eulerian polynomials is (see for instance [8, Section 6.5]):

(5.1) n0An(q)tnn!=1q1qet(1q).\sum_{n\geq 0}A_{n}(q)\frac{t^{n}}{n!}=\frac{1-q}{1-qe^{t(1-q)}}.

The dimension of H2k(XΔAn1)H^{2k}(X_{\Delta_{A_{n-1}}}) is A(n,k+1)A(n,k+1) (see [48, Section 4]), so the Poincaré polynomial of XΔAn1X_{\Delta_{A_{n-1}}}, written following the convention that degq=2\deg q=2, is

P(XΔAn1,q)=k=0n1A(n,k+1)qk=1qAn(q).P(X_{\Delta_{A_{n-1}}},q)=\sum_{k=0}^{n-1}A(n,k+1)q^{k}=\frac{1}{q}A_{n}(q).

From Theorem 4.7 we know that a basis for the cohomology of 𝒴T(An1)Y(XΔAn1,An1)\mathcal{Y}^{T}(A_{n-1})\coloneqq Y(X_{\Delta_{A_{n-1}}},\mathcal{F}_{A_{n-1}}) is given by the elements of An1\mathcal{B}_{\mathcal{F}_{A_{n-1}}}. Recall that these elements are products of the form bmfb\,m_{f}, where ff is an admissible function and bΘ(suppf)b\in\Theta(\operatorname{supp}f); we are going to study these two factors in this case.

Monomial mfm_{f}. In analogy with [30] and [49], we can associate in a natural way an admissible function f:An1f\colon\mathcal{F}_{A_{n-1}}\to\mathbb{N} with a so-called admissible forest (on nn leaves).

Definition 5.2.

An admissible tree on mm leaves is a labeled directed rooted tree such that

  • it has mm leaves, each labeled with a distinct non-zero natural number;

  • each non-leaf vertex vv has kv3k_{v}\geq 3 outgoing edges, and it is labeled with the symbol qiq^{i} where i{1,,kv2}i\in\{1,\dotsc,k_{v}-2\}.

By convention, the graph with one vertex and no edges is an admissible tree on one leaf (actually the only one). The degree of an admissible tree is the sum of the exponents of the labels of the non-leaf vertices.

Definition 5.3.

An admissible forest on nn leaves is the disjoint union of admissible trees such that the sets of labels of their leaves form a partition of {1,,n}\{1,\dotsc,n\}. The degree of an admissible forest is the sum of the degrees of its connected components.

As illustrated by Example 5.5, the association between admissible forests and functions is the following: given an admissible forest FF, for each internal vertex vv let I(v)I(v) be the set of labels of the leaves that descend from vv and let i(v)i(v) be such that qi(v)q^{i(v)} is the label of vv; then the admissible function ff associated with FF has suppf={𝒦I(v)v internal vertex of F}\operatorname{supp}f=\{\mathcal{K}_{I(v)}\mid v\textrm{ internal vertex of }F\} and, for each vv, f:𝒦I(v)i(v)f\colon\mathcal{K}_{I(v)}\mapsto i(v).

Remark 5.4.

The degree of an admissible forest associated with the function ff is equal to the degree of the monomial mfm_{f}.

Example 5.5.

The admissible forest of Figure 3 is associated with the function ff with

suppf={𝒦{1,7,9,12},𝒦{8,10,13},𝒦{1,5,7,8,9,10,12,13},𝒦{2,6,11,14}}\operatorname{supp}f=\{\mathcal{K}_{\{1,7,9,12\}},\mathcal{K}_{\{8,10,13\}},\mathcal{K}_{\{1,5,7,8,9,10,12,13\}},\mathcal{K}_{\{2,6,11,14\}}\}

and such that f(𝒦{1,7,9,12})=2f(\mathcal{K}_{\{1,7,9,12\}})=2, f(𝒦{8,10,13})=1f(\mathcal{K}_{\{8,10,13\}})=1, f(𝒦{1,5,7,8,9,10,12,13})=1f(\mathcal{K}_{\{1,5,7,8,9,10,12,13\}})=1, f(𝒦{2,6,11,14})=1f(\mathcal{K}_{\{2,6,11,14\}})=1. If we denote by TIT_{I} the variable corresponding to 𝒦IA13\mathcal{K}_{I}\in\mathcal{F}_{A_{13}}, the monomial mfm_{f} in this case is

T{1,7,9,12}2T{8,10,13}T{1,5,7,8,9,10,12,13}T{2,6,11,14}.T_{\{1,7,9,12\}}^{2}T_{\{8,10,13\}}T_{\{1,5,7,8,9,10,12,13\}}T_{\{2,6,11,14\}}.
11779912125588101013132266111114143344q2q^{2}q1q^{1}q1q^{1}q1q^{1}
Figure 3. An example of an admissible forest on 14 leaves with degree 5.

Element bΘ(suppf)b\in\Theta(\operatorname{supp}f). To study the elements bΘ(suppf)b\in\Theta(\operatorname{supp}f), first of all we need to analyze H(XΔAn1(suppf))H^{*}(X_{\Delta_{A_{n-1}}(\operatorname{supp}f)}): it is easy to show that the subfan ΔAn1(suppf)\Delta_{A_{n-1}}(\operatorname{supp}f) is isomorphic to ΔAk1\Delta_{A_{k-1}}, where kk is the number of connected components of the forest associated with ff (the isomorphism is obtained by identifying the coordinates associated with the leaves of the same tree). Therefore the elements of Θ(suppf)\Theta(\operatorname{supp}f) are in bijection with the permutations of SkS_{k}, and any statistics on SkS_{k} that is equidistributed with the statistic des\operatorname{des} makes this bijection grade-preserving. We choose to use the so-called lec statistic, first introduced in [29]. To describe it, we need a couple of definitions. In the following, a permutation in SnS_{n} will be denoted by the ordered nn-tuple [σ(1),,σ(n)][\sigma(1),\dotsc,\sigma(n)].

5.2. Some remarks on the statistic lec

Given an ordered list of distinct numbers (not necessarily a permutation), say σ=[σ1,,σn]\sigma=[\sigma_{1},\ldots,\sigma_{n}], we denote by inv(σ)\operatorname{inv}(\sigma) the set of inversions of σ\sigma:

inv(σ):={(i,j)i<j,σi>σj}.\operatorname{inv}(\sigma):=\{(i,j)\mid i<j,\ \sigma_{i}>\sigma_{j}\}.
Definition 5.6.

A hook is an ordered list of distinct non-zero natural numbers τ=[t1,,th]\tau=[t_{1},\dotsc,t_{h}], with h2h\geq 2, such that t1>t2t_{1}>t_{2} and t2<t3<<tht_{2}<t_{3}<\dotsb<t_{h} (this second condition applies only for h3h\geq 3).

Remark 5.7.

Given ss numbers 1j1<<jsn1\leq j_{1}<\dotsb<j_{s}\leq n and i{1,,s1}i\in\{1,\ldots,s-1\} there is a unique way to sort {j1,,js}\{j_{1},\ldots,j_{s}\} so that they form a hook with exactly ii inversions, namely [ji+1,j1,,ji,ji+2,,js][j_{i+1},j_{1},\ldots,j_{i},j_{i+2},\ldots,j_{s}].

It is easy to observe that every list of distinct numbers has a unique hook factorization (this notion comes from [31]), i.e. it is possible to write σ\sigma as a concatenation σ=pτ1τk\sigma=p\tau_{1}\dotsm\tau_{k} where each τi\tau_{i} is a hook and pp is a list of increasing numbers. Notice that it is possible to have k=0k=0, if σ\sigma is an increasing sequence; also it may happen that p=p=\emptyset (σ=[3,1,2]\sigma=[3,1,2] is an example with k=1k=1). The statistic lec\operatorname{lec} is defined as

lec(σ)=i=1k|inv(τi)|\operatorname{lec}(\sigma)=\sum_{i=1}^{k}\left|\operatorname{inv}(\tau_{i})\right|

where pτ1τkp\tau_{1}\dotsb\tau_{k} is the hook factorization of σ\sigma.

Example 5.8.

Let σ=[10,13,14,8,3,6,5,4,7,11,12,9,1,2]\sigma=[10,13,14,8,3,6,5,4,7,11,12,9,1,2]. Its hook factorization is

[10,13,14][8,3,6][5,4,7,11,12][9,1,2][10,13,14]\,[8,3,6]\,[5,4,7,11,12]\,[9,1,2]

and lec(σ)=2+1+2=5\operatorname{lec}(\sigma)=2+1+2=5.

Our choice of the statistic lec\operatorname{lec} has been again inspired by the theory of wonderful models. As it is well known (see for instance [3]), XΔAn1X_{\Delta_{A_{n-1}}} can be also seen as a projective wonderful model for the boolean hyperplane arrangement in n1\mathbb{P}^{n-1}. More precisely, it is the maximal model: the (projective) hyperplanes are

Hi={[z1,z2,,zn]n1zi=0}H_{i}=\{[z_{1},z_{2},\dotsc,z_{n}]\in\mathbb{P}^{n-1}\mid z_{i}=0\}

for i=1,,ni=1,\dotsc,n and the building set is provided by the full poset of their intersections. The nested sets in this case are simply the chains of elements in this poset.

Therefore from [30] we know how to describe a monomial basis of H(XΔAn1,)H^{*}(X_{\Delta_{A_{n-1}}},\mathbb{Z}). In fact we will describe a basis of the cohomology of the corresponding non-projective model, but the two cohomologies are isomorphic (this is a general property, see [16, 30]).

A monomial in this basis is a product of Chern classes associated with an admissible function (in analogy with our previous definitions; see also [49, 30]). In particular, the support of the (function associated with the) monomial is a chain of subsets of {1,,n}\{1,\ldots,n\}.

As an example let n=10n=10 and consider the monomial

ζ{1,2}ζ{1,2,4,5,6}2ζ{1,2,4,5,6,7,8}\zeta_{\{1,2\}}\zeta_{\{1,2,4,5,6\}}^{2}\zeta_{\{1,2,4,5,6,7,8\}}

which is an element of the basis of the cohomology of XΔA9X_{\Delta_{A_{9}}}; the variable ζI\zeta_{I}, I{1,,10}I\subset\{1,\ldots,10\}, is the Chern class of the irreducible divisor obtained as proper transform of the subspace HI:=iIHiH_{I}:=\cap_{i\in I}H_{i}. Notice that, for instance, the exponent of ζ{1,2,4,5,6}\zeta_{\{1,2,4,5,6\}} is strictly less than 33, i.e. the codimension of H{1,2,4,5,6}H_{\{1,2,4,5,6\}} in H{1,2}H_{\{1,2\}}.

We show an algorithm producing a bijection between this monomial basis of H(XΔAn1,)H^{*}(X_{\Delta_{A_{n-1}}},\mathbb{Z}) and SnS_{n}, which is grade-preserving provided that we consider in SnS_{n} the grade induced by the statistic lec\operatorname{lec}. The idea is to write a permutation σSn\sigma\in S_{n} in terms of its hook decomposition, associating a hook with every power of Chern class appearing in the monomial.

  • We first look at the elements in {1,,n}\{1,\ldots,n\} that do not appear in the support of the monomial. We write them in increasing order obtaining the non-hook part pp of σ\sigma. In our example we have {1,,10}{1,2,4,5,6,7,8}={3,9,10}\{1,\dotsc,10\}\setminus\{1,2,4,5,6,7,8\}=\{3,9,10\} so p=[3,9,10]p=[3,9,10].

  • We then create the first hook of σ\sigma by using Remark 5.7 with the numbers in the smallest set of the support of the monomial, and the number of the inversions given by the corresponding exponent. In our example the smallest set is {1,2}\{1,2\} with exponent 11 so τ1=[2,1]\tau_{1}=[2,1].

  • The second hook of σ\sigma is formed using the numbers of the second set of the chain that do not appear in the smallest one. In the example those numbers are {4,5,6}={1,2,4,5,6}{1,2}\{4,5,6\}=\{1,2,4,5,6\}\setminus\{1,2\}, so we form the hook with two inversions since 22 is the exponent of ζ{1,2,4,5,6}\zeta_{\{1,2,4,5,6\}} in the monomial: τ2=[6,4,5]\tau_{2}=[6,4,5].

  • We go on building the ii-th hook τi\tau_{i} by looking at the numbers in the ii-th set of the support of the monomial that do not appear in the (i1)(i-1)-th set. In our case there is only one set remaining: we pick {7,8}\{7,8\} from {1,2,4,5,6,7,8}\{1,2,4,5,6,7,8\} and form the hook [8,7][8,7].

In the end we obtain σ=[3,9,10][2,1][6,4,5][8,7]\sigma=[3,9,10]\,[2,1]\,[6,4,5]\,[8,7], which has lec(σ)=4\operatorname{lec}(\sigma)=4.

Notice that this bijection, if one already knows that the Betti numbers of XΔAn1X_{\Delta_{A_{n-1}}} coincide with the Eulerian numbers, gives a geometric interpretation of the fact that the lec\operatorname{lec} statistic is Eulerian.

5.3. The toric model and the subspace model: an explicit bijection between their cohomology bases

In Section 5.1 we have established a bijection between the basis An1\mathcal{B}_{\mathcal{F}_{A_{n-1}}} and the set of pairs (F,σ)(F,\sigma) where:

  • FF is an admissible forest on nn leaves,

  • σ\sigma is a permutation in SmS_{m}, where mm is the number of connected components of the forest FF.

If we define the degree of a pair as deg(F,σ)=deg(F)+lec(σ)\deg(F,\sigma)=\deg(F)+\operatorname{lec}(\sigma), this bijection is grade-preserving.

Now, it can be proved that the model 𝒴T(An1)\mathcal{Y}^{T}(A_{n-1}) is isomorphic to the projective model 𝒴H(An)\mathcal{Y}^{H}(A_{n}) for the hyperplane arrangement of type AnA_{n}, obtained by blowing up the building set of irreducible elements. Even if we don’t use this fact in the present paper (we mention it only as inspiring additional information), we sketch here a proof.

The first step consists in noticing that both the toric and the hyperplane arrangements can be seen as the same subspace arrangement in a projective space of dimension n1n-1. On one side, the hyperplane arrangement of type AnA_{n} can be seen as the arrangement in

V={(0,x1,,xn)xi}n+1V=\{(0,x_{1},\dotsc,x_{n})\mid x_{i}\in\mathbb{C}\}\subseteq\mathbb{C}^{n+1}

given by the hyperplanes

x1=0,,xn=0,xixj=0 for 1i<jn.x_{1}=0,\dotsc,\ x_{n}=0,\ x_{i}-x_{j}=0\textrm{ for }1\leq i<j\leq n.

The corresponding projective arrangement in (V)\mathbb{P}(V) is given by the hyperplanes

y1=0,,yn=0,yiyj=0 for 1i<jny_{1}=0,\dotsc,\ y_{n}=0,\ y_{i}-y_{j}=0\textrm{ for }1\leq i<j\leq n

where, omitting the first zero, we denote by [y1,,yn][y_{1},\dotsc,y_{n}] the projective coordinates of a point in (V)\mathbb{P}(V). On the other side we observe that we can identify ()n1(\mathbb{C}^{*})^{n-1} with

(n)i=0n1{ti=0}\mathbb{P}(\mathbb{C}^{n})\setminus\bigcup_{i=0}^{n-1}\{t_{i}=0\}

via the map (t1,,tn1)[1,t1,,tn1](t_{1},\dotsc,t_{n-1})\mapsto[1,t_{1},\dotsc,t_{n-1}], where we denote by [t0,,tn1][t_{0},\dotsc,t_{n-1}] the projective coordinates in (n)\mathbb{P}(\mathbb{C}^{n}). In this setting, the divisorial layers of the toric arrangement of type An1A_{n-1} T=()n/Hn()n1T=(\mathbb{C}^{*})^{n}/H_{n}\simeq(\mathbb{C}^{*})^{n-1} are given by

ti=tj for 0i<jn1.t_{i}=t_{j}\textrm{ for }0\leq i<j\leq n-1.

Overall, we are considering in (n)\mathbb{P}(\mathbb{C}^{n}) the hyperplanes

t0=0,,tn1=0,titj=0 for 0i<jn1.t_{0}=0,\dotsc,\ t_{n-1}=0,\ t_{i}-t_{j}=0\textrm{ for }0\leq i<j\leq n-1.

The second step of the proof consists now in noticing that the two models 𝒴T(An1)\mathcal{Y}^{T}(A_{n-1}) and 𝒴H(An)\mathcal{Y}^{H}(A_{n}) are obtained by blowing up the same subspaces; however, the two constructions differ in the order in which the blow ups are carried out, but thanks to [34, Theorem 1.3] the two resulting varieties are isomorphic.

This suggests us to search for a grade-preserving bijection between the bases of the cohomologies of 𝒴T(An1)\mathcal{Y}^{T}(A_{n-1}) and 𝒴H(An)\mathcal{Y}^{H}(A_{n}). Recall that a basis for the cohomology of 𝒴H(An)\mathcal{Y}^{H}(A_{n}) is in grade-preserving bijection with the set of admissible forests on n+1n+1 leaves [49, 30]. So we describe an algorithm that produces an explicit bijection Ψ\Psi, associating a pair (F,σ)(F,\sigma) (given by an admissible forest FF on nn leaves with mm trees and a permutation σSm\sigma\in S_{m}) with an admissible forest \mathcal{F} on n+1n+1 leaves.

As a preliminary step we fix an ordering of the trees in FF. For example we can say that T<TT<T^{\prime} if the minimum index labelling the leaves in TT is smaller than the minimum index labelling the leaves in TT^{\prime}. We denote the trees accordingly as T1<T2<<TmT_{1}<T_{2}<\dotsb<T_{m}.

Let σ=pτ1τk\sigma=p\tau_{1}\dotsm\tau_{k} be the hook factorization of σ\sigma.

Base step: k=0k=0. In this case \mathcal{F} is obtained from FF by simply adding a connected component with a single vertex-leaf labeled with n+1n+1.

Inductive step: k>0k>0. Let τk=[M,a1,,a]\tau_{k}=[M,a_{1},\dotsc,a_{\ell}] be the last hook of σ\sigma and let i=|inv(τk)|i=\left|\operatorname{inv}(\tau_{k})\right|, i.e. the number of inversions. We produce a tree TT connecting with a new internal labeled vertex the roots of the trees TM,Ta1,,TaT_{M},T_{a_{1}},\ldots,T_{a_{\ell}} and an extra leaf labeled with n+1n+1. We label this new internal vertex, which is the root of TT, with qiq^{i}. Figure 4 shows the situation.

\ldotsTa1T_{a_{1}}\ldotsTMT_{M}\ldotsTaT_{a_{\ell}}n+1n+1\ldotsqiq^{i}
Figure 4. The new tree TT obtained using the last hook τk\tau_{k} of σ\sigma.

Now we consider the forest FF^{\prime} obtained from FF by removing the trees TM,Ta1,,TaT_{M},T_{a_{1}},\ldots,T_{a_{\ell}} and the list σ=pτ1τk1\sigma^{\prime}=p\tau_{1}\dotsb\tau_{k-1} and we apply the same construction to the pair (F,σ)(F^{\prime},\sigma^{\prime}), with the difference that now, instead of connecting the trees to an extra leaf labeled n+1n+1, we connect them to the root of the tree TT obtained in the previous step. The algorithm is repeated inductively until there are no hooks remaining.

To prove that this algorithm defines a bijection, we present the reverse algorithm that computes Ψ1\Psi^{-1}, associating an admissible forest \mathcal{F} on n+1n+1 leaves with a pair (F,σ)(F,\sigma).

Description of the forest FF. As the first step of the reverse algorithm, we remove all the internal vertices of \mathcal{F} that have the leaf labeled n+1n+1 among their descendants. When we remove a vertex, we remove also its label and all its outgoing edges (but not their descendants). Then we remove the leaf labeled with n+1n+1. In this way, we have obtained a forest FF on nn leaves, and we sort its connected components according to the usual ordering T1,,TmT_{1},\ldots,T_{m} (see Figure 5).

11445522337766q1q^{1}q1q^{1}\longrightarrow114455223366q1q^{1}
Figure 5. An example of the application of the reverse algorithm.

Description of the permutation σ\sigma. We now need to describe σSm\sigma\in S_{m}. If \mathcal{F} has no internal vertices with the leaf n+1n+1 as a descendant, we just take σ=e\sigma=e, the identity in SmS_{m}. Otherwise, let vv be the vertex in \mathcal{F} that covers n+1n+1; let qiq^{i} be its label and let {Ta1,,Tas,n+1}\{T_{a_{1}},\ldots,T_{a_{s}},n+1\} be the set of connected components of the forest obtained by removing vv and its outgoing edges from the subtree of \mathcal{F} with root vv. The situation is described in Figure 6. Since \mathcal{F} is an admissible forest we have 1is11\leq i\leq s-1, so we can apply Remark 5.7 to the set {a1,,as}\{a_{1},\ldots,a_{s}\} and obtain the hook [ai+1,a1,,ai,ai+2,,as][a_{i+1},a_{1},\ldots,a_{i},a_{i+2},\ldots,a_{s}], which will be the last hook of the permutation σ\sigma.

\ldotsTa2T_{a_{2}}\ldotsTa1T_{a_{1}}\ldotsTasT_{a_{s}}n+1n+1\ldotsvvqiq^{i}
Figure 6. The vertex vv covers the leaf labeled with n+1n+1.

Let now {1,,m}{a1,,as}={b1,,bms}\{1,\ldots,m\}\setminus\{a_{1},\ldots,a_{s}\}=\{b_{1},\ldots,b_{m-s}\}, with b1<<bmsb_{1}<\dotsb<b_{m-s}. If in \mathcal{F} there are no vertices that cover vv we define

σ=[b1,,bms,ai+1,a1,,ai,ai+2,,as];\sigma=[b_{1},\ldots,b_{m-s},a_{i+1},a_{1},\ldots,a_{i},a_{i+2},\ldots,a_{s}];

if instead there is a vertex, say ww, that covers vv in \mathcal{F} we have a picture like Figure 7, with c1<<chc_{1}<\dotsb<c_{h} and 1rh11\leq r\leq h-1. We repeat the same step as we did for vv, obtaining a new hook [cr+1,c1,,cr,cr+2,,ch][c_{r+1},c_{1},\ldots,c_{r},c_{r+2},\ldots,c_{h}] so that the last part of σ\sigma is now

[cr+1,c1,,cr,cr+2,,ch,ai+1,a1,,ai1,ai+2,,as].[c_{r+1},c_{1},\ldots,c_{r},c_{r+2},\ldots,c_{h},a_{i+1},a_{1},\ldots,a_{i-1},a_{i+2},\ldots,a_{s}].
\ldotsTc1T_{c_{1}}\ldotsTchT_{c_{h}}\ldots(Fig. 6)vv\ldotswwqrq^{r}
Figure 7. The vertex ww covers the vertex vv of Figure 6.

We repeat the previous steps as long as there are internal vertices in \mathcal{F} covering the last vertex that we removed.

5.4. A combinatorial proof, with a geometrical interpretation, that lec is Eulerian

The bijection Ψ\Psi described above allows us to give a new proof that lec\operatorname{lec} is an Eulerian statistic. This proof is purely combinatorial, and therefore in particular it differs from the one sketched in Section 5.2, which uses the fact that the Betti numbers of XΔAn1X_{\Delta_{A_{n-1}}} are Eulerian numbers. Nevertheless our proof has a geometric inspiration that comes from counting elements of monomial bases of cohomologies of models. We first need to introduce some generating functions.

Let λ(q,t)\lambda(q,t) be the generating function of the admissible trees, i.e. the series whose coefficient of qitk/k!{q^{i}t^{k}}/{k!} counts the number of admissible trees of degree ii on kk leaves (see [49, 30, 37]). There are explicit combinatorial ways to compute the series λ\lambda, as the following theorem shows.

Theorem 5.9 ([30, Theorem 4.1]).

Let λ\lambda defined as above. Then we have the following recurrence relation:

tλ=1+tλq1(eqλqeλ+q1).\frac{\partial}{\partial t}\lambda=1+\frac{\frac{\partial}{\partial t}\lambda}{q-1}(e^{q\lambda}-qe^{\lambda}+q-1).

In other words

(5.2) tλ=1qeqλqeλ.\frac{\partial}{\partial t}\lambda=\frac{1-q}{e^{q\lambda}-qe^{\lambda}}.

The first few terms of λ\lambda are

λ(q,t)=t+qt33!+(q+q2)t44!+\lambda(q,t)=t+q\frac{t^{3}}{3!}+(q+q^{2})\frac{t^{4}}{4!}+\dotsb

By standard combinatorial arguments we deduce that the generating function of the admissible forests is eλ1e^{\lambda}-1, and in particular the number of the admissible forests with kk connected components on nn leaves and degree dd is counted by the coefficient of qdtn/n!q^{d}{t^{n}}/{n!} in the series λk(q,t)/k!{\lambda^{k}(q,t)}/{k!}.

We define now the exponential generating function for the lec\operatorname{lec} statistic

(q,t)n1(σSnqlec(σ))tnn!\mathcal{L}(q,t)\coloneqq\sum_{n\geq 1}\left(\sum_{\sigma\in S_{n}}q^{\operatorname{lec}(\sigma)}\right)\frac{t^{n}}{n!}

and the usual exponential generating function for any Eulerian statistic

(q,t)n1(σSnqdes(σ))tnn!=n1An(q)qtnn!.\mathcal{E}(q,t)\coloneqq\sum_{n\geq 1}\left(\sum_{\sigma\in S_{n}}q^{\operatorname{des}(\sigma)}\right)\frac{t^{n}}{n!}=\sum_{n\geq 1}\frac{A_{n}(q)}{q}\frac{t^{n}}{n!}.

Our goal is to prove that (q,t)=(q,t)\mathcal{L}(q,t)=\mathcal{E}(q,t). This is equivalent to prove that

(q,λ(q,t))=(q,λ(q,t))\mathcal{L}(q,\lambda(q,t))=\mathcal{E}(q,\lambda(q,t))

since λ\lambda, viewed as a series in [q][[t]]\mathbb{Z}[q][[t]], is invertible with respect to the composition (its constant term is zero and its degree 1 term is invertible in [q]\mathbb{Z}[q]).

Notice that the coefficient of qdtn/n!q^{d}t^{n}/n! in the series (q,λ(q,t))\mathcal{L}(q,\lambda(q,t)) counts the pairs (F,σ)(F,\sigma) where FF is an admissible forest on nn leaves and deg(F,σ)=d\deg(F,\sigma)=d; from the bijection Ψ\Psi we deduce that the series (q,λ(q,t))\mathcal{L}(q,\lambda(q,t)) is equal to the series eλ1e^{\lambda}-1 shifted by one, i.e.

(5.3) (q,λ(q,t))=t(eλ(q,t)1)1.\mathcal{L}(q,\lambda(q,t))=\frac{\partial}{\partial t}(e^{\lambda(q,t)}-1)-1.

Now, a simple computation of formal power series gives that

(5.4) t(eλ(q,t)1)1=(q,λ(q,t)).\frac{\partial}{\partial t}(e^{\lambda(q,t)}-1)-1=\mathcal{E}(q,\lambda(q,t)).

In fact from (5.1) we can write

q(q,λ(q,t))=1q1qeλ(q,t)(1q)1q\cdot\mathcal{E}(q,\lambda(q,t))=\frac{1-q}{1-qe^{\lambda(q,t)\cdot(1-q)}}-1

from which we have

(q,λ(q,t))=1+eλ(q,t)(1q)1qeλ(q,t)(1q)=eλ(q,t)eλ(q,t)qeλ(q,t)qqeλ(q,t);\mathcal{E}(q,\lambda(q,t))=\frac{-1+e^{\lambda(q,t)\cdot(1-q)}}{1-qe^{\lambda(q,t)\cdot(1-q)}}=\frac{e^{\lambda(q,t)}-e^{\lambda(q,t)\cdot q}}{e^{\lambda(q,t)\cdot q}-qe^{\lambda(q,t)}};

on the other hand

t(eλ(q,t)1)1\displaystyle\frac{\partial}{\partial t}(e^{\lambda(q,t)}-1)-1 =eλ(q,t)t(λ(q,t))1=\displaystyle{}\stackrel{{\scriptstyle\phantom{\eqref{eq_lambda_primo}}}}{{=}}e^{\lambda(q,t)}\frac{\partial}{\partial t}(\lambda(q,t))-1=
=(5.2)eλ(q,t)(1q)eλ(q,t)qqeλ(q,t)1=eλ(q,t)eλ(q,t)qeλ(q,t)qqeλ(q,t).\displaystyle{}\stackrel{{\scriptstyle\eqref{eq_lambda_primo}}}{{=}}\frac{e^{\lambda(q,t)}(1-q)}{e^{\lambda(q,t)\cdot q}-qe^{\lambda(q,t)}}-1=\frac{e^{\lambda(q,t)}-e^{\lambda(q,t)\cdot q}}{e^{\lambda(q,t)\cdot q}-qe^{\lambda(q,t)}}.

By combining (5.3) and (5.4) we conclude.

This proof has used only combinatorial arguments, but, as we remarked above, it has a geometric inspiration. We notice that some of the power series involved are actually the generating functions for the Poincaré polynomials of the compact models 𝒴T(An1)\mathcal{Y}^{T}(A_{n-1}) and 𝒴H(An1)\mathcal{Y}^{H}(A_{n-1}), defined as

ΦT(q,t)\displaystyle\Phi^{T}(q,t) :=n1P(𝒴T(An1),q)tnn!,\displaystyle{}:=\sum_{n\geq 1}P(\mathcal{Y}^{T}(A_{n-1}),q)\frac{t^{n}}{n!},
ΦH(q,t)\displaystyle\Phi^{H}(q,t) :=n1P(𝒴H(An1),q)tnn!.\displaystyle{}:=\sum_{n\geq 1}P(\mathcal{Y}^{H}(A_{n-1}),q)\frac{t^{n}}{n!}.

In fact we already know that ΦH(q,t)=eλ(q,t)1\Phi^{H}(q,t)=e^{\lambda(q,t)}-1, and the description of the basis for the toric model gives that ΦT(q,t)=(q,λ(q,t))\Phi^{T}(q,t)=\mathcal{E}(q,\lambda(q,t)). From this point of view we can read Equation (5.4) as

tΦH(q,t)1=ΦT(q,t),\frac{\partial}{\partial t}\Phi^{H}(q,t)-1=\Phi^{T}(q,t),

which reveals itself to be a consequence of the isomorphism between 𝒴T(An1)\mathcal{Y}^{T}(A_{n-1}) and 𝒴H(An)\mathcal{Y}^{H}(A_{n}).

Appendix A Description of the fan in Example 2.19

The following table lists the rays of the fan Δ\Delta associated with a good toric variety for the arrangement of Example 2.19.

r1:(0,2,1)r_{1}\colon\left(0,-2,-1\right) r2:(0,1,1)r_{2}\colon\left(0,-1,-1\right) r3:(2,1,1)r_{3}\colon\left(-2,1,-1\right) r4:(1,1,1)r_{4}\colon\left(-1,1,-1\right)
r5:(0,1,1)r_{5}\colon\left(0,-1,1\right) r6:(0,1,0)r_{6}\colon\left(0,1,0\right) r7:(2,1,1)r_{7}\colon\left(-2,1,1\right) r8:(1,1,1)r_{8}\colon\left(-1,1,1\right)
r9:(1,1,1)r_{9}\colon\left(1,-1,1\right) r10:(1,3,1)r_{10}\colon\left(-1,3,1\right) r11:(2,3,1)r_{11}\colon\left(-2,3,1\right) r12:(1,2,1)r_{12}\colon\left(1,-2,-1\right)
r13:(1,1,1)r_{13}\colon\left(1,-1,-1\right) r14:(6,3,1)r_{14}\colon\left(6,-3,-1\right) r15:(2,1,1)r_{15}\colon\left(2,-1,-1\right) r16:(2,2,1)r_{16}\colon\left(2,-2,-1\right)
r17:(2,1,2)r_{17}\colon\left(2,-1,-2\right) r18:(0,0,1)r_{18}\colon\left(0,0,-1\right) r19:(2,1,1)r_{19}\colon\left(2,-1,1\right) r20:(0,0,1)r_{20}\colon\left(0,0,1\right)
r21:(0,2,1)r_{21}\colon\left(0,2,1\right) r22:(1,0,1)r_{22}\colon\left(1,0,1\right) r23:(1,0,1)r_{23}\colon\left(1,0,-1\right) r24:(6,3,1)r_{24}\colon\left(-6,3,1\right)
r25:(2,2,1)r_{25}\colon\left(-2,-2,-1\right) r26:(2,1,1)r_{26}\colon\left(-2,-1,-1\right) r27:(4,3,1)r_{27}\colon\left(4,-3,-1\right) r28:(0,1,0)r_{28}\colon\left(0,-1,0\right)
r29:(2,0,1)r_{29}\colon\left(2,0,-1\right) r30:(5,3,1)r_{30}\colon\left(-5,3,1\right) r31:(2,1,1)r_{31}\colon\left(-2,-1,1\right) r32:(2,0,1)r_{32}\colon\left(2,0,1\right)
r33:(1,1,1)r_{33}\colon\left(-1,-1,1\right) r34:(1,2,1)r_{34}\colon\left(-1,-2,-1\right) r35:(1,1,0)r_{35}\colon\left(-1,1,0\right) r36:(2,1,0)r_{36}\colon\left(-2,1,0\right)
r37:(2,2,1)r_{37}\colon\left(2,2,1\right) r38:(3,3,1)r_{38}\colon\left(3,-3,-1\right) r39:(1,1,0)r_{39}\colon\left(1,-1,0\right) r40:(4,2,1)r_{40}\colon\left(-4,2,1\right)
r41:(1,1,1)r_{41}\colon\left(-1,-1,-1\right) r42:(2,1,0)r_{42}\colon\left(2,-1,0\right) r43:(1,2,1)r_{43}\colon\left(1,2,1\right) r44:(5,3,1)r_{44}\colon\left(5,-3,-1\right)
r45:(3,3,1)r_{45}\colon\left(-3,3,1\right) r46:(2,0,1)r_{46}\colon\left(-2,0,1\right) r47:(1,0,1)r_{47}\colon\left(-1,0,1\right) r48:(3,2,1)r_{48}\colon\left(3,-2,-1\right)
r49:(2,0,1)r_{49}\colon\left(-2,0,-1\right) r50:(1,0,1)r_{50}\colon\left(-1,0,-1\right) r51:(1,2,1)r_{51}\colon\left(-1,2,1\right) r52:(2,2,1)r_{52}\colon\left(-2,2,1\right)
r53:(1,0,0)r_{53}\colon\left(1,0,0\right) r54:(0,3,1)r_{54}\colon\left(0,-3,-1\right) r55:(2,3,1)r_{55}\colon\left(2,-3,-1\right) r56:(4,3,1)r_{56}\colon\left(-4,3,1\right)
r57:(0,1,1)r_{57}\colon\left(0,1,-1\right) r58:(2,1,1)r_{58}\colon\left(2,1,-1\right) r59:(1,3,1)r_{59}\colon\left(1,-3,-1\right) r60:(0,1,1)r_{60}\colon\left(0,1,1\right)
r61:(2,1,2)r_{61}\colon\left(-2,1,2\right) r62:(2,1,1)r_{62}\colon\left(2,1,1\right) r63:(0,3,1)r_{63}\colon\left(0,3,1\right) r64:(2,3,1)r_{64}\colon\left(-2,-3,-1\right)
r65:(1,3,1)r_{65}\colon\left(-1,-3,-1\right) r66:(1,3,1)r_{66}\colon\left(1,3,1\right) r67:(4,2,1)r_{67}\colon\left(4,-2,-1\right) r68:(1,0,0)r_{68}\colon\left(-1,0,0\right)
r69:(2,3,1)r_{69}\colon\left(2,3,1\right) r70:(1,1,1)r_{70}\colon\left(1,1,-1\right) r71:(1,1,1)r_{71}\colon\left(1,1,1\right) r72:(3,2,1)r_{72}\colon\left(-3,2,1\right)

The following table lists the maximal cones of the fan Δ\Delta associated with a good toric variety for the arrangement of Example 2.19. Each cone is given by its generating rays.

C(r6,r53,r69)C(r_{6},r_{53},r_{69}) C(r37,r53,r69)C(r_{37},r_{53},r_{69}) C(r37,r53,r62)C(r_{37},r_{53},r_{62}) C(r6,r66,r69)C(r_{6},r_{66},r_{69})
C(r37,r66,r69)C(r_{37},r_{66},r_{69}) C(r37,r43,r66)C(r_{37},r_{43},r_{66}) C(r37,r43,r62)C(r_{37},r_{43},r_{62}) C(r43,r62,r71)C(r_{43},r_{62},r_{71})
C(r6,r63,r66)C(r_{6},r_{63},r_{66}) C(r21,r63,r66)C(r_{21},r_{63},r_{66}) C(r21,r43,r66)C(r_{21},r_{43},r_{66}) C(r21,r43,r60)C(r_{21},r_{43},r_{60})
C(r43,r60,r71)C(r_{43},r_{60},r_{71}) C(r32,r53,r62)C(r_{32},r_{53},r_{62}) C(r22,r32,r62)C(r_{22},r_{32},r_{62}) C(r22,r62,r71)C(r_{22},r_{62},r_{71})
C(r22,r60,r71)C(r_{22},r_{60},r_{71}) C(r20,r22,r60)C(r_{20},r_{22},r_{60}) C(r6,r53,r58)C(r_{6},r_{53},r_{58}) C(r6,r58,r70)C(r_{6},r_{58},r_{70})
C(r6,r57,r70)C(r_{6},r_{57},r_{70}) C(r29,r53,r58)C(r_{29},r_{53},r_{58}) C(r23,r29,r58)C(r_{23},r_{29},r_{58}) C(r23,r58,r70)C(r_{23},r_{58},r_{70})
C(r18,r23,r70)C(r_{18},r_{23},r_{70}) C(r18,r57,r70)C(r_{18},r_{57},r_{70}) C(r19,r28,r39)C(r_{19},r_{28},r_{39}) C(r19,r42,r53)C(r_{19},r_{42},r_{53})
C(r19,r39,r42)C(r_{19},r_{39},r_{42}) C(r9,r19,r28)C(r_{9},r_{19},r_{28}) C(r5,r9,r28)C(r_{5},r_{9},r_{28}) C(r19,r32,r53)C(r_{19},r_{32},r_{53})
C(r19,r22,r32)C(r_{19},r_{22},r_{32}) C(r19,r20,r22)C(r_{19},r_{20},r_{22}) C(r9,r19,r20)C(r_{9},r_{19},r_{20}) C(r5,r9,r20)C(r_{5},r_{9},r_{20})
C(r28,r54,r59)C(r_{28},r_{54},r_{59}) C(r1,r54,r59)C(r_{1},r_{54},r_{59}) C(r1,r12,r59)C(r_{1},r_{12},r_{59}) C(r1,r2,r12)C(r_{1},r_{2},r_{12})
C(r28,r39,r55)C(r_{28},r_{39},r_{55}) C(r28,r55,r59)C(r_{28},r_{55},r_{59}) C(r12,r55,r59)C(r_{12},r_{55},r_{59}) C(r15,r53,r67)C(r_{15},r_{53},r_{67})
C(r14,r42,r53)C(r_{14},r_{42},r_{53}) C(r14,r53,r67)C(r_{14},r_{53},r_{67}) C(r27,r39,r42)C(r_{27},r_{39},r_{42}) C(r27,r42,r44)C(r_{27},r_{42},r_{44})
C(r27,r44,r48)C(r_{27},r_{44},r_{48}) C(r15,r48,r67)C(r_{15},r_{48},r_{67}) C(r14,r42,r44)C(r_{14},r_{42},r_{44}) C(r14,r44,r48)C(r_{14},r_{44},r_{48})
C(r14,r48,r67)C(r_{14},r_{48},r_{67}) C(r27,r38,r39)C(r_{27},r_{38},r_{39}) C(r27,r38,r48)C(r_{27},r_{38},r_{48}) C(r16,r38,r48)C(r_{16},r_{38},r_{48})
C(r15,r16,r48)C(r_{15},r_{16},r_{48}) C(r13,r15,r16)C(r_{13},r_{15},r_{16}) C(r2,r12,r13)C(r_{2},r_{12},r_{13}) C(r38,r39,r55)C(r_{38},r_{39},r_{55})
C(r16,r38,r55)C(r_{16},r_{38},r_{55}) C(r13,r16,r55)C(r_{13},r_{16},r_{55}) C(r12,r13,r55)C(r_{12},r_{13},r_{55}) C(r15,r29,r53)C(r_{15},r_{29},r_{53})
C(r15,r23,r29)C(r_{15},r_{23},r_{29}) C(r13,r15,r17)C(r_{13},r_{15},r_{17}) C(r15,r17,r23)C(r_{15},r_{17},r_{23}) C(r13,r17,r18)C(r_{13},r_{17},r_{18})
C(r17,r18,r23)C(r_{17},r_{18},r_{23}) C(r2,r13,r18)C(r_{2},r_{13},r_{18}) C(r6,r10,r63)C(r_{6},r_{10},r_{63}) C(r10,r21,r63)C(r_{10},r_{21},r_{63})
C(r10,r21,r51)C(r_{10},r_{21},r_{51}) C(r21,r51,r60)C(r_{21},r_{51},r_{60}) C(r6,r11,r35)C(r_{6},r_{11},r_{35}) C(r6,r10,r11)C(r_{6},r_{10},r_{11})
C(r10,r11,r51)C(r_{10},r_{11},r_{51}) C(r7,r40,r68)C(r_{7},r_{40},r_{68}) C(r24,r36,r68)C(r_{24},r_{36},r_{68}) C(r24,r40,r68)C(r_{24},r_{40},r_{68})
C(r35,r36,r56)C(r_{35},r_{36},r_{56}) C(r30,r36,r56)C(r_{30},r_{36},r_{56}) C(r30,r56,r72)C(r_{30},r_{56},r_{72}) C(r7,r40,r72)C(r_{7},r_{40},r_{72})
C(r24,r30,r36)C(r_{24},r_{30},r_{36}) C(r24,r30,r72)C(r_{24},r_{30},r_{72}) C(r24,r40,r72)C(r_{24},r_{40},r_{72}) C(r35,r45,r56)C(r_{35},r_{45},r_{56})
C(r45,r56,r72)C(r_{45},r_{56},r_{72}) C(r45,r52,r72)C(r_{45},r_{52},r_{72}) C(r7,r52,r72)C(r_{7},r_{52},r_{72}) C(r7,r8,r52)C(r_{7},r_{8},r_{52})
C(r8,r51,r60)C(r_{8},r_{51},r_{60}) C(r11,r35,r45)C(r_{11},r_{35},r_{45}) C(r11,r45,r52)C(r_{11},r_{45},r_{52}) C(r8,r11,r52)C(r_{8},r_{11},r_{52})
C(r8,r11,r51)C(r_{8},r_{11},r_{51}) C(r7,r46,r68)C(r_{7},r_{46},r_{68}) C(r7,r46,r47)C(r_{7},r_{46},r_{47}) C(r7,r8,r61)C(r_{7},r_{8},r_{61})
C(r7,r47,r61)C(r_{7},r_{47},r_{61}) C(r8,r20,r61)C(r_{8},r_{20},r_{61}) C(r20,r47,r61)C(r_{20},r_{47},r_{61}) C(r8,r20,r60)C(r_{8},r_{20},r_{60})
C(r3,r6,r35)C(r_{3},r_{6},r_{35}) C(r3,r36,r68)C(r_{3},r_{36},r_{68}) C(r3,r35,r36)C(r_{3},r_{35},r_{36}) C(r3,r4,r6)C(r_{3},r_{4},r_{6})
C(r4,r6,r57)C(r_{4},r_{6},r_{57}) C(r3,r49,r68)C(r_{3},r_{49},r_{68}) C(r3,r49,r50)C(r_{3},r_{49},r_{50}) C(r3,r18,r50)C(r_{3},r_{18},r_{50})
C(r3,r4,r18)C(r_{3},r_{4},r_{18}) C(r4,r18,r57)C(r_{4},r_{18},r_{57}) C(r28,r31,r68)C(r_{28},r_{31},r_{68}) C(r28,r31,r33)C(r_{28},r_{31},r_{33})
C(r5,r28,r33)C(r_{5},r_{28},r_{33}) C(r31,r46,r68)C(r_{31},r_{46},r_{68}) C(r31,r46,r47)C(r_{31},r_{46},r_{47}) C(r31,r33,r47)C(r_{31},r_{33},r_{47})
C(r20,r33,r47)C(r_{20},r_{33},r_{47}) C(r5,r20,r33)C(r_{5},r_{20},r_{33}) C(r28,r64,r68)C(r_{28},r_{64},r_{68}) C(r25,r64,r68)C(r_{25},r_{64},r_{68})
C(r25,r26,r68)C(r_{25},r_{26},r_{68}) C(r28,r64,r65)C(r_{28},r_{64},r_{65}) C(r25,r64,r65)C(r_{25},r_{64},r_{65}) C(r25,r34,r65)C(r_{25},r_{34},r_{65})
C(r25,r26,r34)C(r_{25},r_{26},r_{34}) C(r26,r34,r41)C(r_{26},r_{34},r_{41}) C(r28,r54,r65)C(r_{28},r_{54},r_{65}) C(r1,r54,r65)C(r_{1},r_{54},r_{65})
C(r1,r34,r65)C(r_{1},r_{34},r_{65}) C(r1,r2,r34)C(r_{1},r_{2},r_{34}) C(r2,r34,r41)C(r_{2},r_{34},r_{41}) C(r26,r49,r68)C(r_{26},r_{49},r_{68})
C(r26,r49,r50)C(r_{26},r_{49},r_{50}) C(r26,r41,r50)C(r_{26},r_{41},r_{50}) C(r2,r41,r50)C(r_{2},r_{41},r_{50}) C(r2,r18,r50)C(r_{2},r_{18},r_{50})

Acknowledgments

The authors would like to thank Michele D’Adderio for the useful conversations and for pointing out the statistic lec\operatorname{lec} introduced in [29]. The authors also acknowledge the support of INdAM-GNSAGA.

References

  • [1] Karim Adiprasito, June Huh and Eric Katz “Hodge theory for combinatorial geometries” In Annals of Math. 188.2, 2018, pp. 381–452 DOI: 10.4007/annals.2018.188.2.1
  • [2] Omid Amini and Matthieu Piquerez “Homology of tropical fans”, 2021 arXiv:2105.01504 [math.AG]
  • [3] Victor Batyrev and Mark Blume “The functor of toric varieties associated with Weyl chambers and Losev-Manin moduli spaces” In Tohoku Math. J. 63.4, 2011, pp. 581–604
  • [4] Filippo Callegaro and Emanuele Delucchi “The integer cohomology algebra of toric arrangements” In Adv. Math. 313, 2017, pp. 746–802
  • [5] Filippo Callegaro and Emanuele Delucchi “Erratum to “The integer cohomology algebra of toric arrangements””, 2019 arXiv:1910.13836 [math.AT]
  • [6] Filippo Callegaro, Giovanni Gaiffi and Pierre Lochak “Divisorial inertia and central elements in braid groups” In J. Algebra 457, 2016, pp. 26–44 DOI: 10.1016/j.jalgebra.2016.03.006
  • [7] Filippo Callegaro et al. “Orlik-Solomon type presentations for the cohomology algebra of toric arrangements” In Trans. Amer. Math. Soc. 373.3, 2020, pp. 1909–1940 DOI: 10.1090/tran/7952
  • [8] Louis Comtet “Advanced Combinatorics” D. Reidel Publishing Company, 1974
  • [9] Michele D’Adderio and Luca Moci “Arithmetic matroids, the Tutte polynomial and toric arrangements” In Adv. Math. 232.1, 2013, pp. 335–367 DOI: 10.1016/j.aim.2012.09.001
  • [10] Vladimir I. Danilov “The geometry of toric varieties” In Russian Math. Surveys 33.2, 1978, pp. 97–154
  • [11] Corrado De Concini and Giovanni Gaiffi “Projective wonderful models for toric arrangements” In Adv. Math. 327, 2018, pp. 390–409 DOI: 10.1016/j.aim.2017.06.019
  • [12] Corrado De Concini and Giovanni Gaiffi “Cohomology Rings of Compactifications of Toric Arrangements” In Algebraic and Geometric Topology 19, 2019, pp. 503–532 DOI: 10.2140/agt.2019.19.503
  • [13] Corrado De Concini and Giovanni Gaiffi “A differential algebra and the homotopy type of the complement of a toric arrangement” In Atti Accad. Naz. Lincei Cl. Sci. Fis. Mat. Natur. 32.1, 2021, pp. 1–21 DOI: 10.4171/RLM/924
  • [14] Corrado De Concini, Giovanni Gaiffi and Oscar Papini “On projective wonderful models for toric arrangements and their cohomology” In European Jour. of Math. 6.3, 2020, pp. 790–816 DOI: 10.1007/s40879-020-00414-z
  • [15] Corrado De Concini and Claudio Procesi “Hyperplane Arrangements and Holonomy Equations” In Selecta Mathematica, New Series 1.3, 1995, pp. 495–535
  • [16] Corrado De Concini and Claudio Procesi “Wonderful Models of Subspace Arrangements” In Selecta Mathematica, New Series 1.3, 1995, pp. 459–494
  • [17] Corrado De Concini and Claudio Procesi “On the Geometry of Toric Arrangements” In Transformation Groups 10.3, 2005, pp. 387–422 DOI: 10.1007/s00031-005-0403-3
  • [18] Corrado De Concini and Claudio Procesi “Topics in Hyperplane Arrangements, Polytopes and Box-Splines”, Universitext Springer, 2010
  • [19] Corrado De Concini, Claudio Procesi and Michèle Vergne “Vector partition functions and index of transversally elliptic operators” In Transformation Groups 15.4, 2010, pp. 775–811 DOI: 10.1007/s00031-010-9101-x
  • [20] Graham Denham “Toric and tropical compactifications of hyperplane complements” In Annales de la Faculté des Sciences de Toulouse XXIII.2, 2014, pp. 297–333 DOI: 10.5802/afst.1408
  • [21] Graham Denham and Alexander I. Suciu “Local systems on arrangements of smooth, complex algebraic hypersurfaces” In Forum of Mathematics, Sigma 6, 2018, pp. e6 DOI: 10.1017/fms.2018.5
  • [22] Igor Dolgachev and Valery Lunts “A Character Formula for the Representation of a Weyl Group in the Cohomology of the Associated Toric Variety” In J. Algebra 168, 1994, pp. 741–772
  • [23] Vladimir G. Drinfeld In Algebra i Analiz 2.4, 1990, pp. 149–181
  • [24] Vladimir G. Drinfeld “On quasitriangular quasi-Hopf algebras and on a group that is closely connected with Gal(𝐐¯/𝐐)\mathrm{Gal}(\overline{\mathbf{Q}}/\mathbf{Q}) In St. Petersburg Math. J. 2.4, 1991, pp. 829–860
  • [25] Clément Dupont “Purity, Formality, and Arrangement Complements” In Int. Math. Res. Not. 2016.13, 2016, pp. 4132–4144 DOI: 10.1093/imrn/rnv260
  • [26] Pavel Etingof, André Henriques, Joel Kamnitzer and Eric M. Rains “The cohomology ring of the real locus of the moduli space of stable curves of genus 0 with marked points” In Annals of Math. 171.2, 2010, pp. 731–777
  • [27] Eva Maria Feichtner and Bernd Sturmfels “Matroid polytopes, nested sets and Bergman fans” In Portugaliae Mathematica, Nova Série 62.4, 2005, pp. 437–468
  • [28] Eva Maria Feichtner and Sergey Yuzvinsky “Chow rings of toric varieties defined by atomic lattices” In Invent. math. 155.3, 2004, pp. 515–536
  • [29] Dominique Foata and Guo-Niu Han “Fix-mahonian calculus III; a quadruple distribution” In Monatsh. Math. 154, 2008, pp. 177–197 DOI: 10.1007/s00605-007-0512-2
  • [30] Giovanni Gaiffi “Blowups and cohomology bases for De Concini-Procesi models of subspace arrangements” In Selecta Mathematica, New Series 3.3, 1997, pp. 315–333 DOI: 10.1007/s000290050013
  • [31] Ira M. Gessel “A Coloring Problem” In Amer. Math. Monthly 98.6, 1991, pp. 530–533
  • [32] Phillip Griffiths and Joseph Harris “Principles of Algebraic Geometry” Wiley, 1978
  • [33] Anthony Henderson “Representations of wreath products on cohomology of De Concini-Procesi compactifications” In Int. Math. Res. Not. 2004.20, 2004, pp. 983–1021 DOI: 10.1155/S1073792804132510
  • [34] Li Li “Wonderful compactification of an arrangement of subvarieties” In Michigan Math. Jour. 58.2, 2009, pp. 535–563
  • [35] Eduard Looijenga “Cohomology of 3\mathcal{M}_{3} and 31\mathcal{M}^{1}_{3} In Mapping Class Groups and Moduli Spaces of Riemann Surfaces 150, Contemp. Math. Amer. Math. Soc., Providence, RI, 1993, pp. 205–228 DOI: 10.1090/conm/150/01292
  • [36] Robert MacPherson and Claudio Procesi “Making conical compactification wonderful” In Selecta Mathematica, New Series 4.1, 1998, pp. 125–139
  • [37] Yuri I. Manin “Generating Functions in Algebraic Geometry and Sums Over Trees” In The Moduli Space of Curves Birkhäuser Boston, 1995, pp. 401–417
  • [38] Luca Moci “Combinatorics and topology of toric arrangements defined by root systems” In Atti Accad. Naz. Lincei Cl. Sci. Fis. Mat. Natur. 19.4, 2008, pp. 293–308 DOI: 10.4171/RLM/526
  • [39] Luca Moci “A Tutte polynomial for toric arrangements” In Trans. Amer. Math. Soc. 364.2, 2012, pp. 1067–1088 DOI: 10.1090/S0002-9947-2011-05491-7
  • [40] Luca Moci and Roberto Pagaria “On the cohomology of arrangements of subtori” In J. London Math. Soc., 2022 DOI: 10.1112/jlms.12616
  • [41] Roberto Pagaria “Two examples of toric arrangements” In J. Combin. Theory, Series A 167, 2019, pp. 389–402 DOI: 10.1016/j.jcta.2019.05.006
  • [42] Oscar Papini “Computational Aspects of Line and Toric Arrangements”, 2018 URL: https://etd.adm.unipi.it/t/etd-09262018-113931/
  • [43] Claudio Procesi “The toric variety associated to Weyl chambers” In Mots: mélanges offerts à M.-P. Schützenberger, Lang. Raison. Calc. Hermès Science, 1990, pp. 153–161
  • [44] Eric M. Rains “The homology of real subspace arrangements” In Jour. of Topology 3.4, 2010, pp. 786–818
  • [45] The Sage Developers “SageMath, the Sage Mathematics Software System (Version 9.0)”, 2020 URL: https://www.sagemath.org
  • [46] Richard P. Stanley “Log-Concave and Unimodal Sequences in Algebra, Combinatorics, and Geometry” In Annals of the New York Academy of Sciences 576, 1989, pp. 500–535 DOI: 10.1111/j.1749-6632.1989.tb16434.x
  • [47] Richard P. Stanley “Enumerative Combinatorics, Vol. 1” Cambridge University Press, 2011
  • [48] John R. Stembridge “Eulerian numbers, tableaux, and the Betti numbers of a toric variety” In Discrete Math. 99.1–3, 1992, pp. 307–320
  • [49] Sergey Yuzvinsky “Cohomology bases for the De Concini-Procesi models of hyperplane arrangements and sums over trees” In Invent. Math. 127.2, 1997, pp. 319–335 DOI: 10.1007/s002220050122