A basis for the cohomology of compact models of toric arrangements
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 . 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 numbers2010 Mathematics Subject Classification:
14N20, 05E16, 05C301. Introduction
Let be an -dimensional torus and let be its group of characters; it is a lattice of rank and, by choosing a basis, we have isomorphisms and . Given an element , the corresponding character on will be denoted by .
Definition 1.1.
A layer in is a subvariety of of the form
where is a split direct summand and is a homomorphism. A toric arrangement is a (finite) set of layers in . 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 projective wonderful model is a smooth projective variety containing as an open set and such that the complement of 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 . 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 (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 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 of a divisorial toric arrangement, the generators of the cohomology modules over 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 . 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 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 .
One of the motivations for the construction of projective wonderful models of a toric arrangement 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 . 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 and it has a presentation which depends only on a set of initial discrete data extracted from , thus proving that in the non-divisorial case the rational homotopy type of 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) 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 with fan , then one considers the arrangement of subvarieties (in the sense of Li [34]) given by the closures of the layers of . One chooses a suitable building set of subvarieties in and blowups them in a prescribed order to obtain the projective wonderful model . The -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 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 -basis of . Every element of this basis has two factors: one is a monomial that depends essentially on a nested set of with certain labels (in analogy with the case of subspace models), the other one comes from the cohomology of a toric subvariety of associated with . 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 . We make a canonical choice of the fan , i.e. we take the fan associated with the Coxeter chambers. Therefore is the toric variety of type 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 marked points, i.e. to the minimal projective wonderful model of the hyperplane arrangement of type .
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 leaves with labels on internal vertices, equipped by an additional label: a permutation in the symmetric group , where is the number of trees. On the other side we have forests on 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 , mainly following [11] (see also [14]).
First of all, let us fix some notation that will be used throughout this paper. Given a set , we will use the symbol to denote the intersection of its elements, namely
Recall from the Introduction that is the group of characters of the torus ; likewise, we denote by the group of one-parameters subgroups of . Moreover we define the vector spaces and its dual . The usual pairing and its extension to will both be denoted by the symbol . Given , we define
(2.1) |
Given a fan in , the corresponding toric variety will be denoted by .
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 be a non-singular algebraic variety. A simple arrangement of subvarieties of is a finite set of non-singular closed connected subvarieties properly contained in such that
-
(1)
for every two , either or ;
-
(2)
if , the intersection is clean, i.e. it is non-singular and for every we have the following conditions on the tangent spaces:
Definition 2.2.
Let be a non-singular algebraic variety. An arrangement of subvarieties of is a finite set of non-singular closed connected subvarieties properly contained in such that
-
(1)
for every two , either is a disjoint union of elements of or ;
-
(2)
if , 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 is the set of the connected components of the intersections of some layers of , partially ordered by reverse inclusion.
Remark 2.4.
-
(1)
The whole torus belongs to , as it can be obtained as the intersection of no layers; we define .
-
(2)
The intersection of two layers and is the disjont union of layers of the form , i.e. they share the same , namely the saturation of .
Given a toric arrangement in a torus , we embed 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 be a fan in . An element has the equal sign property with respect to if, for every cone , either for all or for all .
Definition 2.6.
Let be a fan in and let be a layer. A -basis for is an equal sign basis with respect to if has the equal sign property for all .
We say that a toric variety is good for an arrangement if each layer of has an equal sign basis with respect to the fan . 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 let be the corresponding homogeneous subtorus and let as in (2.1), i.e.
-
(1)
For every cone , its relative interior is either entirely contained in or disjoint from .
-
(2)
The collection of cones which are contained in is a smooth fan .
-
(3)
is a smooth -variety whose fan is .
-
(4)
Let be an orbit of in and let be the corresponding cone. Then
-
(a)
if is not contained in , ;
-
(b)
If , is the -orbit in corresponding to .
-
(a)
As a consequence the set of the connected components of the intersections of the closures of the layers in is an arrangement of subvarieties according to Li’s definition.
Following [11] we now introduce the wonderful model associated with an arrangement of subvarieties in a generic non-singular algebraic variety . To do so we need to define the notion of building sets and nested sets.
Definition 2.8.
Let be a simple arrangement of subvarieties. A subset is a building set for if for every the minimal elements (w.r.t. the inclusion) of the set intersect transversally and their intersection is . These minimal elements are called the -factors of .
Definition 2.9.
Let be a building set for a simple arrangement . A subset is called (-)nested if for any antichain111An antichain in a poset is a set of pairwise non-comparable elements. , with , there is an element in of which are the -factors.
Remark 2.10.
Since the empty set has no antichains of cardinality at least , the definition above applies vacuously for it.
Remark 2.11.
We notice that if is a subset of whose elements have empty intersection, then it cannot be contained in any -nested set.
In case the arrangement is not simple, the definitions above apply locally: first of all, we define the restriction of an arrangement of subvarieties to an open set to be the set
Definition 2.12.
Let be an arrangement of subvarieties of . A subset is a building set for if there is a cover of open sets of such that
-
(1)
for every , the restriction is simple;
-
(2)
for every , is a building set for .
Definition 2.13.
Let be a building set for an arrangement . A subset is called (-)nested if there is an open cover of such that, for every , is simple, is building for and for at least one , is -nested. (In particular for all .)
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 of connected subvarieties of is called a building set if the set of the connected components of all the possible intersections of collections of subvarieties from is an arrangement of subvarieties, called the arrangement induced by and denoted by , and is a building set for 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 of a non-singular variety and a building set for , a wonderful model can be obtained as the closure of the locally closed embedding:
where is the blowup of along . Concretely we can build 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 be a building set in a non-singular variety . Let us order the elements of in such a way that for every the set is building. Then if we set and for , we have
where denotes the dominant transform222In the blowup of a variety along a centre the dominant transform of a subvariety coincides with the proper transform if (and therefore it is isomorphic to the blowup of along ), and with if , where is the projection. We will use the same notation for both the proper and the dominant transform of , if no confusion arises. of in .
Remark 2.16.
Any total ordering of the elements of a building set which refines the ordering by inclusion, that is if , satisfies the condition of Theorem 2.15.
Let us denote by the blowup map. The boundary of admits a description in terms of -nested sets.
Theorem 2.17 (see [34, Theorem 1.2]).
The complement in of is the union of the divisors , where ranges among the elements of . Let be an open cover of such that for every is simple and is building for . Then, given and , the intersection is non-empty if and only if is -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 is considered -nested if is -nested for every . 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 of .
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 be the arrangement in , with coordinates , whose layers are defined by the equations
We can view them as where is generated by , by and by , and is the constant function equal to for . Figure 1 represents the Hasse diagram of the poset of layers .
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 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 .
The next choice is a building set ; for this example we use the subset of the elements of that are pictured in a double circle in Figure 1 and obtain the model . We will study this model in later examples of this paper.
3. Presentation of the cohomology ring
In this section we recall a presentation of the cohomology ring of the model . As we have seen, given a toric arrangement and a toric variety which is good for it, the set is an arrangement of subvarieties of according to Li.
The cohomology ring is described as a quotient of a polynomial ring with coefficients in . 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 be a smooth complete -variety. Let be the set of primitive generators of the rays of and define a polynomial indeterminate for each . Then
where
-
•
is the Stanley-Reisner ideal
-
•
is the linear equivalence ideal
Notice that for it is sufficient to take only the square-free monomials, and for it is sufficient to take only the ’s belonging to a basis of .
Furthermore the residue class of in is the cohomology class of the divisor associated with the ray for each . By abuse of notation we are going to denote this residue class in also by .
Remark 3.2.
Example 3.3 (Example 2.19, continued).
We computed the presentation of as in Theorem 3.1 for the toric variety of Example 2.19. The cohomology ring is isomorphic to a quotient of the ring where each indeterminate corresponds to the ray as listed in the table in Appendix A. We won’t report the full presentation here and give only the Betti numbers:
The presentation of the cohomology ring of has been computed with an additional hypothesis on the building set.
Definition 3.4.
A building set for is well-connected if for any subset , if the intersection has two or more connected components, then each of these components belongs to .
Example 3.5 (Example 2.19, continued).
The building set 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 . Let be an indeterminate and let viewed as as in Theorem 3.1. (For brevity we will use again the symbols instead of the corresponding equivalence classes in the quotient.) For every we denote by the lattice such that .
Given a pair with , we can choose a basis for such that it is equal sign with respect to and that , with , is a basis for (if is the whole variety , then we choose any equal sign basis of and let ). We define the polynomials as
If , we set since it is an empty product. As shown in [12, Proposition 6.3], different choices of the ’s are possible; for example, in [40, Remark 4.4], the authors suggest the polynomials
Now we define the set
where is the power set of . Notice that for all . For each pair we define a relation in the following way: let be the unique connected component of that contains (as usual, if then ), and for let ; with this information we define the polynomial as
Finally let . For each we define the polynomial as
Theorem 3.6 ([12, Theorem 7.1]).
Let , and be as in the beginning of this section and let be a well-connected building set for . Let also viewed as a polynomial ring as in Theorem 3.1. The cohomology ring of the wonderful model is isomorphic to the quotient of by the ideal generated by
-
(1)
the products , with and such that does not belong to ;
-
(2)
the polynomials for every pair ;
-
(3)
the polynomials for every .
The isomorphism is given by sending for to the cohomology class associated with the divisor in the boundary which is the transform of (). Putting all together, we have
Remark 3.7.
It was already noted in [11, Theorem 9.1] that the cohomology of the projective wonderful model is a free -module and for odd.
Example 3.8 (Example 2.19, continued).
We computed the presentation of as in Theorem 3.6 for the model of Example 2.19. The cohomology ring is isomorphic to a quotient of the ring , where the ’s are the same as Example 3.3 and each corresponds to an element of in the following way:
Once again we won’t report the full presentation here and give only the Betti numbers:
In the next Section we are going to give a description of a monomial basis of .
4. Main theorem
Let be a toric arrangement, let be a good toric variety for it with associated fan , let be the poset of intersections of the closures of the layers of in which is an arrangement of subvarieties, and let be a well-connected building set for . In this Section we show a basis of the cohomology of 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 -nested sets in this case, which will be useful in the proof of our main theorem.
Proposition 4.1.
Let be a well-connected building set. A subset is -nested if and only if for any antichain with at least two elements the intersection is non-empty, connected, transversal, and does not belong to .
Proof of Proposition 4.1.
Let be a -nested set an let be an antichain with . The intersection is not empty by Definition 2.13 and Remark 2.11. If is not connected then it would be the disjoint union of at least two elements in by well-connectedness. Let us consider as in Definition 2.13, so that is -nested and for every , therefore is an antichain of . This implies that
would be the -decomposition of , where is one of the connected components of . We reached a contradiction because , and we deduce that is connected and furthermore that it is not an element of . The transversality of the intersection in also implies that is transversal.
We now suppose that for every antichain with at least two elements the intersection is non-empty, connected, transversal and does not belong to .
Let be an open cover as in Definition 2.12. Let us first notice that is equal to the intersection of the antichain given by the minimal elements of , therefore it is non-empty and connected. Let such that is not empty. We will prove that is -nested.
Consider an antichain in with (in particular is non-empty for every ): we will prove that are the -factors of their intersection.
Let us put . We observe that is an antichain, therefore by hypothesis we know that is non-empty, connected, transversal and does not belong to . Since is building, we know that is the transversal intersection of the minimal elements of among the ones containing . We let those minimal elements be . For simplicity, in the rest of the proof we omit the reference to .
By minimality each of the ’s contains some of the ’s: in the next paragraph we partition in subsets according to this.
Up to reordering the indices of the ’s, we can assume that is contained in some of the ’s, and define . If we are done, otherwise there is another contained in the elements of , and up to reordering we assume that this is . We define . If is not empty we repeat the process with and obtain . We stop after steps, where , when we have .
By construction we have:
(4.1) |
so
and this implies that , that is to say, .
Recall that is transversal and so is ; since each is an antichain, is transversal too, and it follows that is transversal. From this information and from (4.1) we deduce that for all , therefore .
This implies that for any , otherwise if that would contradict the fact that the intersection of an antichain does not belong to .
Since we deduce that and, up to a relabeling, we can assume that . We conclude that and this shows that are the -factors of . ∎
Remark 4.2.
Proposition 4.1 implies that when is well-connected the property of being nested can be expressed in global terms (without charts).
Let be a well-connected building set and let be a -nested set. Given , we define and for every we denote by the (connected) intersection . We will omit the nested set and write just for brevity when it is clear from the context which is the involved nested set.
Definition 4.4.
A function is (-)admissible if it has both the following properties:
-
(1)
is -nested;
-
(2)
for every we have .
Notice that the zero function, i.e. the function such that for every , is admissible since its support is the empty set.
Example 4.5 (Example 2.19, continued).
For each nested set 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 (see Definition 4.4). It turns out that only 7 of the 48 nested sets give rise to admissible functions, namely
(4.2) |
In particular we find 11 admissible functions:
Support | Values |
---|---|
for all | |
Let us fix some notation. Given a nested set we know that the intersection is non-empty and it is a layer of type for some and . Let be the subtorus associated with as in Theorem 2.7, namely . By Theorem 2.7, given a fan , is a smooth fan which we denote by . Let be the corresponding toric variety and let be the projection , which is the restriction map induced by the inclusion, i.e. the one of (3.1). Let be a minimal set of elements of such that their image via is a basis of .
Given a (non necessarily admissible) function we define the monomial in viewed as
where and is the (class of the) variable associated with , and denote by the following set of elements of :
(4.3) |
Remark 4.6.
For , the only admissible function is the zero function. In this case the associated monomial is , and is a basis of . In fact by the usual convention, which is the closure of that, as a layer, has and . So and is the identity function. In particular contains the set .
Theorem 4.7.
Let , , and be as in the beginning of this section. The set defined as in (4.3) is a -basis of .
Proof.
This proof is divided in two parts: we first show that the elements in generate as a -module; then we see that they are independent by counting them.
generates as a -module. First of all notice that the relations in Theorem 3.6 imply that, given an admissible function and such that in then in . In fact is a polynomial in terms of variables where does not belong to —this follows from Theorem 2.7. Therefore we can prove that generates by showing that the set
generates as a -module. To show this, let us consider a function which is not admissible: we will prove that can be obtained as a -linear combination of monomials in .
If is non-nested then the intersection of the elements in is empty by Theorem 2.2 of [11], so in . From now on we can assume that is nested, and following the notation introduced before Definition 4.4 we will write .
Let us consider as a set partially ordered by inclusion. On this poset we define a “pseudo-rank” function as follows: and for let be the maximal length of a chain between and in the Hasse diagram of the poset.
We say that is a bad component for if . To every bad component we assign the pair
and we put on the lexicographic order.333i.e. if or if and . Since we assumed that is nested but that is not admissible, the set of bad components for is not empty, so we can define the evaluation of as the maximal pair associated with the bad components of . We will proceed by induction on the evaluation.
Base step. Let us consider a non-admissible function whose evaluation is , for some . This means that the maximal bad components of are minimal elements in w.r.t. inclusion. Let be such a bad component. We can partition as where and . The monomial associated with is of the following type:
From Theorem 3.6 we know that in the ideal there is the element
where .
Now we notice that is a polynomial in of the following form:
By writing as
we see that the leading term of , namely , divides , so when we reduce modulo we obtain a polynomial of the following form:
for some and some suitable . Notice that (eventually without ) and that the ’s coincide with in , whereas
Therefore every monomial appearing in this formula either is in or its evaluation is with . We can apply the same argument to the latter monomials until we get a linear combination of monomials in .
Inductive step. Let us suppose that our claim is true for non-admissible functions (with nested support) whose evaluation is with and . Let us consider a non-admissible function with evaluation . This means that there is at least one bad component whose associated pair is .
As before we consider the element of :
where and . The polynomial is of type
where is a polynomial in with degree in strictly less than .
As in the base step, we notice that the leading term of divides and this allows us to write modulo as a -linear combination of monomials. If these are not modulo either they are in or their evaluation is strictly less than . If some of them have evaluation with we can again use the relations in and in a finite number of steps we get a -linear combination of monomials that are in or have evaluation with . To these latter monomials we can apply the inductive hypothesis.
is a set of independent elements. To show that the elements in are linearly independent over it suffices to show that is equal to the rank of which is a free -module (see Theorem 3.6). We proceed by induction on , the cardinality of .
Base step. If then and is the blowup of along . As it is well known (see for example [32, Chapter 4, Section 6]) we have the following isomorphism of graded -modules:
where is a symbol that shifts the degrees of . We now split into two disjoint subsets:
We observe that by Theorem 2.7, point 3, is isomorphic to . Therefore there is a grade-preserving bijection between and a basis of which, together with the known bijection between and a basis of , 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 . Consider and assume that the labelling is a refinement of the ordering in by inclusion. We put for every and ; we denote then the proper transform of in the variety . Then is obtained as the blow up of along .
We can use again the result from [32] and obtain the following graded isomorphism of -modules:
Following the idea from the base step, we write as the union of the disjoint sets:
There is a bijective correspondence, provided by the restriction, between the set and the set . By the inductive hypothesis, is in bijection with the basis of and this correspondence is grade-preserving.
Notice that because is maximal in , then is maximal in for every admissible function such that . So the possible values for are .
Now we observe that, given an element (so that ), we also have in the monomials for all the admissible functions that coincide with on and such that . As a consequence the sets , for have all the same cardinality and form a partition of , and it suffices to prove that there is a grade-preserving (up to a shift by 2 in cohomology) bijection between and a basis of . This extends to a grade-preserving bijection between and a basis of .
Let us now recall the following result and notation from [12, Section 4]. We consider the family of subvarieties in that are the connected components of the intersections for every . Since is well-connected, if is not empty and not connected then its connected components belong to . This implies that . Now for each , we denote by the minimum index such that is a connected component of (in particular ). We sort the set in ascending order as and let .
Remark 4.8.
Notice that two possibility occurs for : either in case , or and the intersection is transversal.
In [12, Proposition 4.4] it is proven that is building and well-connected and from Proposition 4.6 of the same paper it follows that is isomorphic to the model obtained by blowing up in . From Theorem 2.7, point 2, we know that is a toric variety with fan , so .
In analogy with the previous notation, for an -nested let be the projection and we take as a minimal set of elements of such that their image via is a basis of .
Since we can apply our inductive hypothesis to and we get the following -basis of :
We are now ready to describe a bijective, grade-preserving (up to a shift by 2 in cohomology) correspondence between and
Let be a -admissible function with . We associate to the function such that for all .
Lemma 4.9.
The function is -admissible.
Proof.
If is empty, is admissible by definition, so from now we suppose that .
We show that is -nested by using the characterization from Proposition 4.1: we take an antichain () in and prove that the intersection is connected, transversal and does not belong to .
is connected. It is equal to and they all belong to which is -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.
is transversal. We split this part of the proof in two cases:
-
(1)
if there is a such that , then
which is transversal because is a nested set;
-
(2)
if transversally for every , then the set is a set of elements which are pairwise non-comparable (the ’s are not pairwise comparable because the ’s are not), so their intersection
is transversal.
does not belong to . Suppose that . It is not possible that , because otherwise would belong to in contradiction with Proposition 4.1. On the other hand, if transversally, as a consequence of [12, Proposition 3.3] we have that and are the minimal elements among the ones in containing . We study the following two subcases:
-
•
if there is a such that , then we would have in contradiction with the minimality of ;
-
•
if transversally for every , since we would have two different -decompositions of , namely
This concludes the proof that is nested.
Now let us consider . We study the following two cases:
Case . In this case and, since is -admissible, we have
Now, because we notice that
where is viewed as a -nested set and with the usual convention that if is empty the intersection is the ambient space . In particular
therefore ranges in the expected interval.
Case transversally. In this case where again
Now since is -nested and does not belong to we have that
(4.4) |
But
so we can rewrite (4.4) as
and, observing that , we conclude that
therefore also in this case ranges in the expected interval. ∎
Lemma 4.10.
If is -admissible and , then .
Proof.
Let us suppose with for every . We notice that otherwise so . Moreover, since is nested and contains both and it follows that and connected by Proposition 4.1. Now we observe that from the connectedness of and the definition of we have that for some and that and are its -factors. But is a different -factorization of , obtaining a contradiction. ∎
Therefore, given -admissible with we can associate two monomials: and . Now in we find elements of the form with belonging to ; on the other hand in we find elements of the form with belonging to . But , therefore we can choose and above so that they range over the same set.
We have thus constructed a map such that . If we show that is a bijection, this concludes the proof of the theorem. Actually it is sufficient to prove that is injective: in fact the injectivity implies
where the last equality, as we have seen, derives from the inductive hypothesis, since is a basis of . This in turn implies that
On the other hand we already know, from the first part of this proof, that generates . It follows that
which is the claim of the theorem (and of course this also implies that is actually a bijection).
To show the injectivity of let , be two distinct -admissible functions with : we prove that .
Let us first suppose that ; up to switching and , we can assume that there exists . By Lemma 4.10 we know that for a certain . We deduce that and conclude that .
If instead , implies that there is a certain in their support such that . Again by Lemma 4.10 we know that , , therefore . This proves that is injective and concludes the proof of the theorem. ∎
Example 4.11 (Example 2.19, continued).
We can compute a -basis for the ring 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 of admissible functions, as listed in (4.2).
Basis for | Monomials | |
---|---|---|
with s.t. | ||
Basis of | ||
Contribution to | Contribution to | ||||
the basis | |||||
Basis of | 1 | 69 | 69 | 1 | |
0 | 1 | 1 | 0 | ||
0 | 1 | 1 | 0 | ||
0 | 1 | 1 | 0 | ||
0 | 1 | 1 | 0 | ||
0 | 1 | 1 | 0 | ||
0 | 1 | 1 | 0 | ||
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 be the arrangement in , with coordinates , whose layers are defined by the equations
The poset of layers is represented in Figure 2. In this example is a good toric variety for the arrangement, so we can use its associated fan (recall that its 16 maximal cones are where are the vectors of the canonical basis of and ); moreover we choose and build the model .
The possible non-empty supports of admissible functions are
-
•
: it supports one admissible function such that ;
-
•
, : each supports two admissible functions, namely and ;
-
•
: it supports three admissible functions, where is either , or ;
-
•
: it supports one admissible function such that and .
Table 3 details the contribution to the Betti numbers for each admissible function.
Betti numbers Monomials with Contribution to for s.t. 1, 4, 6, 4, 1 1 4 6 4 1 1, 2, 1 0 1 2 1 0 1, 1 0 1 2 1 0 1, 1 0 1 2 1 0 1 0 1 1 1 0 1 0 0 1 0 0
5. The case of root systems of type
In this section we will apply our main theorem to the case of the toric arrangement associated with a root system of type .
5.1. The minimal toric model and its cohomology basis
The toric analogue of the hyperplane arrangement of type is in , where is the -span of and
Its poset of intersections is isomorphic to the poset of partitions of the set ordered by refiniment. More precisely, the partition with corresponds to the layer
where for the sake of convenience we will sometimes omit to write the blocks with cardinality one.
Let be the set whose elements are the for every with . It is a building set, in fact it is the minimal one that contains the layers ; it is the analogue of the “building set of irreducible elements” for the linear case (see [16, 49, 30]).
For the arrangement there is a natural choice of a fan that produces a good toric variety, as noted in [11]: we take in the fan induced by the Weyl chambers of the root system. By construction every layer of has an equal sign basis with respect to , so the toric variety associated with is a good toric variety for the arrangement.
In [43] Procesi studies this toric variety (and also the more general toric varieties associated with the fan induced by the Weyl chambers of a Weyl group ; see also [22]). As it is well-known, the even Betti numbers of the toric variety are the Eulerian numbers (see, for example, [48, 46, 47]). We recall briefly the main definitions and results about the numbers and the cohomology of .
Definition 5.1.
The Eulerian number is the number of permutations in with descents444If is a permutation in , a descent of is an index such that . The number of descents of is denoted by . for and .
Following [8] we present the Eulerian polynomial as
According to the above formula one can compute the first Eulerian polynomials obtaining , , . The exponential generating function of the Eulerian polynomials is (see for instance [8, Section 6.5]):
(5.1) |
The dimension of is (see [48, Section 4]), so the Poincaré polynomial of , written following the convention that , is
From Theorem 4.7 we know that a basis for the cohomology of is given by the elements of . Recall that these elements are products of the form , where is an admissible function and ; we are going to study these two factors in this case.
Monomial . In analogy with [30] and [49], we can associate in a natural way an admissible function with a so-called admissible forest (on leaves).
Definition 5.2.
An admissible tree on leaves is a labeled directed rooted tree such that
-
•
it has leaves, each labeled with a distinct non-zero natural number;
-
•
each non-leaf vertex has outgoing edges, and it is labeled with the symbol where .
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 leaves is the disjoint union of admissible trees such that the sets of labels of their leaves form a partition of . 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 , for each internal vertex let be the set of labels of the leaves that descend from and let be such that is the label of ; then the admissible function associated with has and, for each , .
Remark 5.4.
The degree of an admissible forest associated with the function is equal to the degree of the monomial .
Example 5.5.
The admissible forest of Figure 3 is associated with the function with
and such that , , , . If we denote by the variable corresponding to , the monomial in this case is
Element . To study the elements , first of all we need to analyze : it is easy to show that the subfan is isomorphic to , where is the number of connected components of the forest associated with (the isomorphism is obtained by identifying the coordinates associated with the leaves of the same tree). Therefore the elements of are in bijection with the permutations of , and any statistics on that is equidistributed with the statistic 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 will be denoted by the ordered -tuple .
5.2. Some remarks on the statistic lec
Given an ordered list of distinct numbers (not necessarily a permutation), say , we denote by the set of inversions of :
Definition 5.6.
A hook is an ordered list of distinct non-zero natural numbers , with , such that and (this second condition applies only for ).
Remark 5.7.
Given numbers and there is a unique way to sort so that they form a hook with exactly inversions, namely .
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 as a concatenation where each is a hook and is a list of increasing numbers. Notice that it is possible to have , if is an increasing sequence; also it may happen that ( is an example with ). The statistic is defined as
where is the hook factorization of .
Example 5.8.
Let . Its hook factorization is
and .
Our choice of the statistic has been again inspired by the theory of wonderful models. As it is well known (see for instance [3]), can be also seen as a projective wonderful model for the boolean hyperplane arrangement in . More precisely, it is the maximal model: the (projective) hyperplanes are
for 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 . 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 .
As an example let and consider the monomial
which is an element of the basis of the cohomology of ; the variable , , is the Chern class of the irreducible divisor obtained as proper transform of the subspace . Notice that, for instance, the exponent of is strictly less than , i.e. the codimension of in .
We show an algorithm producing a bijection between this monomial basis of and , which is grade-preserving provided that we consider in the grade induced by the statistic . The idea is to write a permutation 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 that do not appear in the support of the monomial. We write them in increasing order obtaining the non-hook part of . In our example we have so .
-
•
We then create the first hook of 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 with exponent so .
-
•
The second hook of 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 , so we form the hook with two inversions since is the exponent of in the monomial: .
-
•
We go on building the -th hook by looking at the numbers in the -th set of the support of the monomial that do not appear in the -th set. In our case there is only one set remaining: we pick from and form the hook .
In the end we obtain , which has .
Notice that this bijection, if one already knows that the Betti numbers of coincide with the Eulerian numbers, gives a geometric interpretation of the fact that the 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 and the set of pairs where:
-
•
is an admissible forest on leaves,
-
•
is a permutation in , where is the number of connected components of the forest .
If we define the degree of a pair as , this bijection is grade-preserving.
Now, it can be proved that the model is isomorphic to the projective model for the hyperplane arrangement of type , 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 . On one side, the hyperplane arrangement of type can be seen as the arrangement in
given by the hyperplanes
The corresponding projective arrangement in is given by the hyperplanes
where, omitting the first zero, we denote by the projective coordinates of a point in . On the other side we observe that we can identify with
via the map , where we denote by the projective coordinates in . In this setting, the divisorial layers of the toric arrangement of type are given by
Overall, we are considering in the hyperplanes
The second step of the proof consists now in noticing that the two models and 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 and . Recall that a basis for the cohomology of is in grade-preserving bijection with the set of admissible forests on leaves [49, 30]. So we describe an algorithm that produces an explicit bijection , associating a pair (given by an admissible forest on leaves with trees and a permutation ) with an admissible forest on leaves.
As a preliminary step we fix an ordering of the trees in . For example we can say that if the minimum index labelling the leaves in is smaller than the minimum index labelling the leaves in . We denote the trees accordingly as .
Let be the hook factorization of .
Base step: . In this case is obtained from by simply adding a connected component with a single vertex-leaf labeled with .
Inductive step: . Let be the last hook of and let , i.e. the number of inversions. We produce a tree connecting with a new internal labeled vertex the roots of the trees and an extra leaf labeled with . We label this new internal vertex, which is the root of , with . Figure 4 shows the situation.
Now we consider the forest obtained from by removing the trees and the list and we apply the same construction to the pair , with the difference that now, instead of connecting the trees to an extra leaf labeled , we connect them to the root of the tree 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 , associating an admissible forest on leaves with a pair .
Description of the forest . As the first step of the reverse algorithm, we remove all the internal vertices of that have the leaf labeled 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 . In this way, we have obtained a forest on leaves, and we sort its connected components according to the usual ordering (see Figure 5).
Description of the permutation . We now need to describe . If has no internal vertices with the leaf as a descendant, we just take , the identity in . Otherwise, let be the vertex in that covers ; let be its label and let be the set of connected components of the forest obtained by removing and its outgoing edges from the subtree of with root . The situation is described in Figure 6. Since is an admissible forest we have , so we can apply Remark 5.7 to the set and obtain the hook , which will be the last hook of the permutation .
Let now , with . If in there are no vertices that cover we define
if instead there is a vertex, say , that covers in we have a picture like Figure 7, with and . We repeat the same step as we did for , obtaining a new hook so that the last part of is now
We repeat the previous steps as long as there are internal vertices in covering the last vertex that we removed.
5.4. A combinatorial proof, with a geometrical interpretation, that lec is Eulerian
The bijection described above allows us to give a new proof that 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 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 be the generating function of the admissible trees, i.e. the series whose coefficient of counts the number of admissible trees of degree on leaves (see [49, 30, 37]). There are explicit combinatorial ways to compute the series , as the following theorem shows.
Theorem 5.9 ([30, Theorem 4.1]).
Let defined as above. Then we have the following recurrence relation:
In other words
(5.2) |
The first few terms of are
By standard combinatorial arguments we deduce that the generating function of the admissible forests is , and in particular the number of the admissible forests with connected components on leaves and degree is counted by the coefficient of in the series .
We define now the exponential generating function for the statistic
and the usual exponential generating function for any Eulerian statistic
Our goal is to prove that . This is equivalent to prove that
since , viewed as a series in , is invertible with respect to the composition (its constant term is zero and its degree 1 term is invertible in ).
Notice that the coefficient of in the series counts the pairs where is an admissible forest on leaves and ; from the bijection we deduce that the series is equal to the series shifted by one, i.e.
(5.3) |
Now, a simple computation of formal power series gives that
(5.4) |
In fact from (5.1) we can write
from which we have
on the other hand
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 and , defined as
In fact we already know that , and the description of the basis for the toric model gives that . From this point of view we can read Equation (5.4) as
which reveals itself to be a consequence of the isomorphism between and .
Appendix A Description of the fan in Example 2.19
The following table lists the rays of the fan associated with a good toric variety for the arrangement of Example 2.19.
The following table lists the maximal cones of the fan associated with a good toric variety for the arrangement of Example 2.19. Each cone is given by its generating rays.
Acknowledgments
The authors would like to thank Michele D’Adderio for the useful conversations and for pointing out the statistic 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 ” 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 and ” 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