Resolutions of local face modules, functoriality, and vanishing of local -vectors
Abstract.
We study the local face modules of triangulations of simplices, i.e., the modules over face rings whose Hilbert functions are local -vectors. In particular, we give resolutions of these modules by subcomplexes of Koszul complexes as well as functorial maps between modules induced by inclusions of faces. As applications, we prove a new monotonicity result for local -vectors and new results on the structure of faces in triangulations with vanishing local -vectors.
1. Introduction
In this paper, we study the modules over face rings, introduced by Athanasiadis and Stanley, whose Hilbert functions are the relative local -vectors of quasi-geometric homology triangulations of simplices, a broad class of formal subdivisions that includes all geometric triangulations and is natural from the point of view of combinatorial commutative algebra. See Section 2.1 for the precise definition and further references.
Fix an infinite field . Let be a quasi-geometric homology triangulation of a simplex, and let be a face of . Say that a face is interior if , and let be the ideal in the face ring generated by the faces that are interior relative to , i.e.,
Let , which is the Krull dimension of , and let be a special l.s.o.p., as in [Sta92, Ath12a]. See also §2.2, where we recall the definition and construction of special l.s.o.p.s.
Definition 1.1.
The local face module is the image of in .
Note that is a finite dimensional graded -vector space. The local -vector is its Hilbert function:
The local face module depends on the choice of a special l.s.o.p., but is an invariant of the triangulation with the symmetry . See §2.1 for details and references. In the past few years, there has been significant research activity on the combinatorics of local -vectors and relations to intersection homology [Ath16, KS16, Sta17, dCMM18]. Recent advances include a proof that every non-negative integer vector satisfying and is the local -vector of a quasi-geometric triangulation for [JKMS19], and a relative hard Lefschetz theorem that yields unimodality of local -vectors for regular subdivisions in a more general setting (for regular nonsimplicial polyhedral subdivisions that are not necessarily rational) [Kar19].
Here, we investigate the local face modules using methods of combinatorial commutative algebra. In particular, we describe natural combinatorial resolutions of these modules as well as natural maps of -modules, , for . Our first theorem gives explicit generators for the kernel of the natural map . Moreover, we extend this to an exact sequence of graded -modules in which each term is a direct sum of degree-shifted monomial ideals.
Label the vertices of the simplex . For a subset , let . After relabeling, we may assume that . Given , we define the ideal by
Note that for , and depends only on . For instance, and if . By the definition of a special l.s.o.p. (Definition 2.3), after reordering, we may assume
for . As a consequence, for any , multiplication by induces a degree 1 map .
Theorem 1.2.
There is an exact sequence of graded -modules
where, for , with , the differential restricted to is .
Corollary 1.3.
The kernel of the surjection is the ideal generated by
We also construct maps between local face modules, as follows. For faces in , let denote the closed star of in . We have a natural inclusion of complexes .
Theorem 1.4.
Let be faces of , with
Let be a special l.s.o.p. for , and let . Then there is a unique homomorphism of graded -algebras
whose kernel contains and satisfies for all . Moreover, there is a special l.s.o.p. for such that and, up to reordering, we have , for . With this choice of special l.s.o.p., .
Remark 1.5.
Theorem 1.4 may be viewed as a functoriality statement for local face modules. Start by fixing the special l.s.o.p. . Then is well-defined. For the special l.s.o.p. depends on some choices, but the ideal that it generates does not, nor does the map . Moreover, for , one readily checks that the maps and are independent of all choices and satisfy . Thus one obtains a functor from the poset of faces of that contain to graded vector spaces, given by .
We now give two applications of the above theorems. The first is a monotonicity property for local -vectors.
Theorem 1.6.
Let be faces of such that . Then .
The inequality in Theorem 1.6 is term by term, i.e., for all . The proof is by showing that the map given by Theorem 1.4 is surjective.
Our second application of the above theorems is to a decades old problem posed by Stanley, who introduced and studied local -vectors in the special case where and asked for a characterization of triangulations for which they vanish [Sta92, Problem 4.13]. This problem remains open, and is of enduring interest [Ath16, Problem 2.12]. The extension to the case where is not empty is particularly relevant for applications to the monodromy conjecture [Igu78, DL98, Sta17]. In [LPS22], we prove a theorem on the structure of geometric triangulations with vanishing local -vectors that is tailored to this purpose, and we use it to prove the monodromy conjectures for all singularities that are nondegenerate with respect to a simplicial Newton polyhedron. See Theorems 1.1.1, 1.4.3, and 4.1.3 in loc. cit.
Here, we apply Theorem 1.2 to prove another theorem on the structure of faces in triangulations with vanishing local -vectors. Let be a face such that is interior. Following terminology from the monodromy conjecture literature (see, e.g., [LVP11]), we say that is a pyramid with apex if is not interior. Let
The elements of correspond to the base directions of , i.e., the facets of that contain the base of , when viewed as a pyramid with apex . We say is a -pyramid if there is an apex such that . In other words, a -pyramid is a pyramid with a unique base direction, for some choice of apex.
Definition 1.7.
Let be a face. An interior partition of is a decomposition
such that and are both interior.
Theorem 1.8.
Suppose and has an interior partition such that . Then is a -pyramid.
See Remark 3.2 for a short proof in a special case that illustrates the naturality of the -pyramid condition. The method of proof breaks down when . See Example 5.3.
Remark 1.9.
The analogous theorem in [LPS22] requires that the triangulation be geometric and that the interior partition satisfies the additional condition . But then the hypothesis that is dropped entirely. So, even for geometric triangulations, there are cases of Theorem 1.8 that are not necessarily covered by [LPS22, Theorem 4.1.3]. It should be interesting to look for a common generalization of these vanishing results, and to pursue further progress on Stanley’s problem of characterizing triangulations with vanishing local -vector more generally.
Remark 1.10.
To the best of our knowledge, all of the theorems stated in the introduction are new even for regular triangulations. The reader who prefers to do so may safely restrict attention to geometric or even regular triangulations. However, while the structure results for triangulations with vanishing local -vectors in [dMGP+20] and [LPS22] rely on special properties of geometric triangulations, the proofs presented here work equally well for quasi-geometric homology triangulations, and we find it natural to work in this level of generality.
We conclude the introduction with an example illustrating the above theorems.
Example 1.11.
Let be the triforce triangulation, which figures prominently in [dMGP+20] and in the adventures of hero protagonist Link in the video game series The Legend of Zelda.
Let , , , , , . Consider first . The face ring is
and its ideal of interior faces is
A special l.s.o.p. is of the form , with
subject to the condition that the restrictions (of the corresponding affine linear functions) to the face are linearly independent. Our resolution of the local face module also involves the monomial ideals
The resolution given by Theorem 1.2 is then
In particular, we have , where
Since , , and restrict to linearly independent functions on , the elements span the 3-dimensional subspace of . Hence and .
Next, consider . Then
A special l.s.o.p. is any l.s.o.p. of the form , where . The ideal of interior faces in this case is , and the resolution given by Theorem 1.2 is
Note, in particular, that , where . Thus one sees that has dimension 1 in degree 1, i.e., .
Let us now consider Theorem 1.4 in this example. Let denote the restriction of to . Note that is supported on . Extend to a basis for , e.g., by choosing to be a linear combination of and in which the coefficient of vanishes. Then is a special l.s.o.p. for , and the map in Theorem 1.4 is given as follows. First, we set
Then, writing , with all three coefficients nonzero, we set
Note that there is no subset of whose restrictions to form an l.s.o.p. This explains and motivates our two-step process for constructing the map: first restricting to and then intersecting with to produce the special l.s.o.p. that yields the functorial map .
Let also describe how Theorems 1.6 and 1.8 manifest in this example. For Theorem 1.8, observe that the face in has an interior partition . The proof in this case shows that the classes of both and are nonzero in , for any choice of special l.s.o.p.
Finally, note that and , so there is no surjective map of graded vector space . In this case, . Thus, we see that the hypothesis cannot be dropped in Theorem 1.6.
Acknowledgments. We thank the referees for their helpful comments. The work of ML is supported by an NDSEG fellowship and the work of SP is supported in part by NSF DMS–2001502 and DMS–2053261.
2. Preliminaries
We begin by recalling definitions and background results that will be used throughout, following [Sta96, Chapter III] and [Ath16]. We work over a field . In particular, all rings are commutative -algebras and singular homology is computed with coefficients in .
2.1. Triangulations of simplices
In this section only, for the purposes of providing context, we allow that the field may be finite, and the triangulation is not necessarily quasi-geometric.
We recall the notion of a homology triangulation, following [Ath12b]. A -dimensional simplicial complex with trivial reduced homology is a homology ball of dimension if there is a subcomplex such that
-
•
is a homology sphere of dimension ,
-
•
is a homology sphere of dimension for .
-
•
is a homology ball of dimension for all nonempty .
The interior faces of a homology ball are the faces not contained in . A homology triangulation of the simplex is a finite simplicial complex and a map such that for every non-empty ,
-
•
the simplicial complex is a homology ball of dimension .
-
•
is the set of interior faces of the homology ball .
Note that the Betti numbers of a simplicial complex, and hence the property of being a homology ball, depend only on the characteristic of the field . Homology triangulations are a special case of the (strong) formal subdivisions of Eulerian posets considered in [Sta92, §7] and [KS16, §3].
The carrier of a face is . A homology triangulation is quasi-geometric if there is no face and such that the dimension of is strictly smaller than the dimension of and the carrier of every vertex in is contained in . A homology triangulation is geometric if it can be realized in as the subdivision of a geometric simplex into geometric simplices. Every geometric homology triangulation is quasi-geometric.
The local -vector, which we have defined in the introduction as the Hilbert function of the local face module, can be expressed in terms of -vectors of subcomplexes of links of faces in the homology balls :
(1) |
Note that (1) makes sense even when is finite or is not quasi-geometric, and should be taken as the definition of the local -vector in this broader context.
Theorem 2.1 ([Sta92, Ath12b, KS16]).
Let be a homology triangulation, let be a face of and let . Then the local -vector satisfies:
(symmetry)
(non-negativity)
if is quasi-geometric, then
(unimodality)
if is regular, then .
Note that the proof of non-negativity for quasi-geometric triangulations, due to Stanley and Athanasiadis, is via the identification with the Hilbert function of the local face module. It suffices to consider the case where is infinite, since (1) is invariant under field extensions.
2.2. Face rings and special l.s.o.p.s
Here, and for the remainder of the paper, the field is fixed and infinite, and all triangulations are quasi-geometric homology triangulations.
Given a finite simplicial complex with vertex set , let denote the face ring. In other words, for each subset , let be the corresponding squarefree monomial in the polynomial ring , i.e., Then the face ring is
Given a subcomplex of , we have a natural restriction map , taking to if and to 0 otherwise. Given , let denote the image of in . In particular, each in may be viewed as a subcomplex, and we write for the restriction of to this subcomplex.
Note that is graded by degree. By definition, a linear system of parameters (l.s.o.p.) for a finitely generated graded -algebra of Krull dimension is a sequence of elements in such that is a finite-dimensional -vector space. If is a Cohen-Macaulay complex (i.e., if is a Cohen-Macaulay ring) and is an l.s.o.p. for , then is a regular sequence and the -polynomial of is the Hilbert series of . Links of faces in triangulations of simplices are Cohen-Macaulay [Rei76].
Suppose has dimension , so has Krull dimension . Then a sequence of elements in is an l.s.o.p. for if and only if the following condition is satisfied [Sta96, Lemma 2.4(a)]:
-
For every face (or equivalently, for every facet ), the restrictions span a vector space of dimension .
This characterization provides flexibility in constructing l.s.o.p.s in which the linear functions have specified support, where the support of is .
Lemma 2.2.
Let be subsets of the vertices of . Then there is an l.s.o.p. for such that for if and only if, for every face ,
(2) |
Proof.
The argument is similar to that given by Stanley in [Sta92, Corollary 4.4]. The necessity of (2) follows immediately from (*). We now prove its sufficiency. Suppose are chosen such that (2) holds for every . Let , and consider the space parametrizing tuples with . Fix . Let parametrize the tuples whose restrictions to span a vector space of dimension . Note that is Zariski open. By Hall’s Marriage Theorem, there is a permutation such that . If we set for , and for , then , and hence is nonempty. Also, the subset of where all coordinates are nonzero is Zariski open and nonempty. Since is infinite, the intersection of these nonempty Zariski open subsets of is nonempty, and hence there is an l.s.o.p. with . ∎
Let be a quasi-geometric homology triangulation, and let be a face.
Definition 2.3 ([Sta92, Ath12a]).
A linear system of parameters for is special if, for each vertex with , there is an element of the l.s.o.p. such that consists of vertices in whose carrier contains , and such that for .
In other words, after reordering so that , an l.s.o.p. for is special if we can order it such that
for . The existence of special l.s.o.p.s is well-known to experts and the proof is similar to Stanley’s argument in the case . For completeness, we provide a short proof.
Proposition 2.4.
Suppose is infinite. Let be a quasi-geometric homology triangulation of a simplex, and let be a face of . Then there is a special l.s.o.p. for .
Proof.
Let . After renumbering, we may assume that . Fix . Note that . We define subsets of the vertices in , as follows. For , let be the set of vertices such that . For , let be the set of all vertices of . Because is quasi-geometric, for each face of , the union of the sets , as ranges over vertices of , has size at least . It follows that . Since for , we conclude that . Hence, by Lemma 2.2, there is an l.s.o.p. for with . ∎
3. A resolution of the local face module
In this section, we prove Theorem 1.2, giving an explicit resolution of the local face module by a subcomplex of the Koszul resolution of . We continue to use the notation established above. In particular, is a quasi-geometric homology triangulation of the simplex with vertex set . We consider a face with and . After reordering, we assume . For , we consider the ideal given by
Let be a special l.s.o.p. for . We may assume that
for . For any , multiplication by gives a map , and we consider the complex of graded -modules
(3) |
in which the differential restricted to , for , with , is .
Example 3.1.
If is an interior face of then every l.s.o.p. is special, for all , and (3) is the Koszul resolution of .
Proof of Theorem 1.2.
We must show (3) is exact. We begin by considering two complexes of -modules studied by Stanley and Athanasiadis. Recall that, for , we write .
Say and , with . For , let be the restriction map. The first complex we consider is
(4) |
in which the differential restricted to is . Next, we consider its quotient by :
(5) |
For any , with , let be defined as
Then and it follows that the restriction of to is nonzero if and only if . Furthermore, is a special l.s.o.p. for . Stanley and Athanasiadis proved that both (4) and (5) are exact, and the kernel of the first arrow in (5) is . (We will recall the proofs below.) Using the additivity of Hilbert functions in exact sequences, they deduced that the Hilbert function of satisfies (1) [Sta92, Ath12b].
With the goal of proving that (3) is exact, we take Koszul resolutions of each term in (5) to build a double complex of -modules. Since is Cohen-Macauley, the special l.s.o.p. is a regular sequence. Hence the corresponding Koszul complex
is exact. Here, for a graded module and a finite set , we write . Replacing each term in (5) with its corresponding Koszul resolution, gives a complex of complexes
(6) |
which may be expanded as the commuting double complex shown in Figure 1.
The columns of this complex are exact by construction. We claim that the rows are also exact, and prove this using ideas from [Sta92, Theorem 4.6]. First, we show that all rows except for the top row are exact. Choose a subset of , and consider the piece of the complex indexed by :
(7) |
When , we obtain (4). Observe that the complex (7) is multigraded by , where is the number of vertices of . Explicitly, . Therefore it suffices to show exactness on graded pieces. Fix . By the definition of the face ring, every term of (7) will have in the graded piece corresponding to unless the set of vertices with forms a face , in which case the -graded part can be identified with the augmented cochain complex of a simplex, indexed by all that contain , and hence is exact.
We now recall the proof that the top row of the double complex, (5), is exact.
The proof involves showing that the quotients of (4) by is exact by induction on . The case of is the exactness of the second row.
Now assume that (4) remains exact after quotienting by . Let denote the th term of (4) tensored with . By the induction hypothesis, we have an exact sequence
Set . Recall that if , and that is a special l.s.o.p. for . Also, for , if and only if . Hence, we have an exact sequence
(8) |
where
For example, when , and . Up to signs and a degree shift, we can then identify with the complex (4) for quotiented by . Then is exact by the induction hypothesis applied to . By breaking (8) up into two short exact sequences we see that as desired.
Now that we know the exactness of (6), let
Then, by construction, we have an exact sequence of complexes
As above, we repeatedly apply the long exact sequence on cohomology to see that is exact. We may then identify with the exact sequence
Since surjects onto and , we conclude that , as required. ∎
Remark 3.2.
Let be a quasi-geometric homology triangulation of a simplex, and let be a face of . Let such that is interior, and suppose that is an interior partition of , i.e., with . Suppose that is not a -pyramid. By Corollary 1.3, is generated by elements of the form for interior or for some with . Because is not a -pyramid, no monomial appearing in any of these generators divides , so is nonzero in . This proves Theorem 1.8 in the special case when .
4. Functorial properties of local face modules
In this section, we prove Theorem 1.4, giving natural maps between local face modules. Consider a quasi-geometric homology triangulation , and let be faces of .
Lemma 4.1.
Let be a graded -algebra with . Let be an l.s.o.p. for , where each has degree . Then there is a unique graded -algebra isomorphism
Moreover, any -basis for is an l.s.o.p. for and generates .
Proof.
Consider the exact sequence of -linear maps
where the right hand map takes to , for any and . This restricts to an exact sequence of -linear maps
where the surjectivity of the right-hand map follows from the fact that is an l.s.o.p. Hence, for , we can write , for some and . For any -algebra map we must have that , so there is a unique such map. On the other hand, the -algebra homomorphism defined by is well-defined, since if , for some and , then . Note that the unique -algebra homomorphism from to is the inverse of .
Since is an isomorphism and factors through , we conclude that the -ideal is generated in degree and hence any -basis for is an l.s.o.p. for . ∎
Proof of Theorem 1.4.
Note that is the join of with . The face ring is therefore a polynomial ring over . Its Krull dimension is equal to , and hence the restrictions form an l.s.o.p., where . By Lemma 4.1, there is a unique graded -algebra homomorphism , which lifts to the unique homomorphism in the statement of the theorem. It remains to construct a special l.s.o.p. for with the specified properties.
After reordering, we may assume that
Note, in particular, that is supported on vertices in the link of , for . By Lemma 4.1, any -basis for is an l.s.o.p. for . Set , for , and note that is linearly independent. Extending this independent set to a basis produces a special l.s.o.p. for . It remains to verify that . Let be a face with interior. If is not in , then . Otherwise, can be written uniquely as the join of possibly empty faces and . Then is interior, and . Hence , as required. ∎
Proof of Theorem 1.6.
Let be faces of a quasi-geometric homology triangulation of a simplex, and assume that . It is enough to show that the induced map given by Theorem 1.4 is surjective. Note that is generated by the monomials such that and is interior. If is such a face, then it is also in the link of and, since , the face is also interior. Then , and the theorem follows. ∎
5. Restrictions of local face modules
In this section, we use the resolution found in Theorem 1.2 to show that the vanishing of a local face module implies the vanishing of a restricted local face module for certain interior partitions . We then develop algebraic arguments, inspired by ideas from [dMGP+20], to show that being a -pyramid is necessary for the vanishing of the restricted local face module when and thus prove Theorem 1.8.
We use the notation introduced in the introduction. Let be a subcomplex of . For any -module , the restriction of to is , where is a -module via the restriction map. By the resolution of in Theorem 1.2 and the right exactness of tensoring with , we have an exact sequence
(9) |
Recall from Corollary 1.3 that , where is the ideal generated by and . Hence, , where are the -ideals
(10) |
(11) |
For example, if is a face of , then is a polynomial ring with variables indexed by the vertices of , and is identified with a quotient of ideals in this polynomial ring.
Lemma 5.1.
Let be a quasi-geometric homology triangulation of a simplex, and let be a face of . Let be a face with interior. Assume that is not a -pyramid. Then there is a surjective graded -module homomorphism
where the second term is a -module via the restriction map .
Proof.
If is a subcomplex of contained in the closed star of , then is a non-zero divisor in . In particular, is a non-zero divisor in (this is also clear since is a polynomial ring). Note that every face of with carrier codimension at most contains . Thus and , where and are the ideals in
Then we have surjective graded -module homomorphisms
where the first map is the isomorphism taking and the second map is restriction. Finally the right hand term can be identified with . ∎
We will derive Theorem 1.8 from the following more technical statement.
Theorem 5.2.
Let be a quasi-geometric homology triangulation, and let be a face. Let be a face with interior. Suppose and admits an interior partition . Assume that has no faces with interior and . If , then is non-zero in degree .
Example 5.3.
Proof of Theorem 1.8.
We may assume that is an interior partition of with minimal among all possible interior partitions of . In particular, if , then there is no vertex such that is interior, as then would be an interior partition. Hence there are no faces of with interior and with cardinality smaller than . By Theorem 5.2, is non-zero in degree . Then, by Lemma 5.1, is nonzero in degree . ∎
We now proceed with the proof of Theorem 5.2. We begin with a series of three lemmas. Inspired by the results of [dMGP+20] in the case , we consider the internal edge graph of a subcomplex . This is the graph contained in the -skeleton of consisting of edges with interior.
Lemma 5.4.
Assume has codimension at least . Let be a subcomplex of , and assume has no vertices with interior. If is zero in degree , then each connected component of the internal edge graph of satisfies one of the following.
-
(1)
The component is a tree, and it has at most one vertex with having carrier codimension more than .
-
(2)
The component has a unique cycle, and the carrier codimension of is equal to for every vertex in the component.
Proof.
From (9), we have the following exact sequence for the degree part of .
Because , the first map in the above complex is surjective. As has no vertices with interior, we see that
(12) |
Thus the number of edges with interior is less than or equal to the number of vertices with the carrier codimension of equal to . If and , then
In particular, both vector spaces in (12) naturally decompose into a direct sum of vector spaces indexed by the connected components of the internal edge graph. Therefore, in each connected component of the internal edge graph, the number of edges with interior is less than or equal to the number of vertices with of carrier codimension . As the only connected graphs where are either trees or contain a unique cycle, the result follows. ∎
Lemma 5.5.
Assume has codimension at least . Let be a face. Assume has no vertices with interior. If is zero in degree , then no component of the internal edge graph of contains a cycle of length .
Proof.
Suppose a component of the internal edge graph contains a -cycle of vertices . By Lemma 5.4, this is the unique cycle in this component and every vertex has of carrier codimension . Because is a face and there are no -cycles in this component of the internal edge graph, we may assume that and . Restricting to and using that , we have that
The relation expands into a relation between the generators of the right-hand side. But the left-hand side is -dimensional, a contradiction. ∎
Lemma 5.6.
Assume has codimension . Let be a subcomplex. Then
Proof.
By considering the degree part of (9), as the codimension of is , we get the following exact sequence.
and the result follows. ∎
Proof of Theorem 5.2.
We must show that is non-zero in degree . Recall that is isomorphic to , where and are described in (10) and (11) respectively. First we handle the cases when . If , then is interior and , but is a proper ideal as it is generated by elements of positive degree, so . If , then we assume that is not an interior face. Then is generated by elements of degree at least , so .
Suppose . We assume that there are no vertices with interior and is not interior. If has codimension , then both and must have a vertex with interior. Then by Lemma 5.6, we see that . Hence we may assume that has codimension at least .
Let and assume that has no non-zero elements in degree . Consider the connected component of the internal edge graph containing . By Lemma 5.4, we may assume that . Note that . There is a vertex such that , so is interior. Therefore either or has carrier codimension .
If , then there is a vertex such that . First assume and are distinct. Since at least one of and has carrier codimension , it follows that is interior. Then forms a -cycle, contradicting Lemma 5.5.
If , then the internal edge graph contains a cycle and hence every vertex in it (including ) has of carrier codimension . As is interior and has carrier codimension , there is a vertex such that is interior. But then either or is interior, contradicting the uniqueness of the cycle in Lemma 5.4.
If does not have carrier codimension , then we may assume that . Choose a vertex with . Then is interior, so has carrier codimension . If , then is interior. If , then is interior. In either case, there is a cycle and a vertex with of carrier codimension more than in the internal edge graph, contradicting Lemma 5.4. ∎
Remark 5.7.
One can use the same overall strategy more generally to show that other combinatorial types of faces cannot appear in triangulations with vanishing local -vectors. For instance, suppose and is a geometric triangulation with a facet such that
Then the interior -faces of are , , , and . But has no interior vertices or edges, and it has only three edges with carrier codimension one, namely and . Thus is non-zero in degree three. Note that is not a pyramid and does not admit an interior partition.
References
- [Ath12a] C. Athanasiadis, Cubical subdivisions and local -vectors, Ann. Comb. 16 (2012), no. 3, 421–448.
- [Ath12b] by same author, Flag subdivisions and -vectors, Pacific J. Math. 259 (2012), no. 2, 257–278.
- [Ath16] by same author, A survey of subdivisions and local -vectors, The mathematical legacy of Richard P. Stanley, Amer. Math. Soc., Providence, RI, 2016, pp. 39–51.
- [dCMM18] M. de Cataldo, L. Migliorini, and M. Mustaţă, Combinatorics and topology of proper toric maps, J. Reine Angew. Math. 744 (2018), 133–163.
- [DL98] J. Denef and F. Loeser, Motivic Igusa zeta functions, J. Algebraic Geom. 7 (1998), no. 3, 505–537.
- [dMGP+20] A. de Moura, E. Gunther, S. Payne, J. Schuchardt, and A. Stapledon, Triangulations of simplices with vanishing local -polynomial, Algebr. Comb. 3 (2020), no. 6, 1417–1430.
- [Igu78] J.-i. Igusa, Forms of higher degree, Tata Institute of Fundamental Research Lectures on Mathematics and Physics, vol. 59, Tata Institute of Fundamental Research, Bombay; by the Narosa Publishing House, New Delhi, 1978.
- [JKMS19] M. Juhnke-Kubitzke, S. Murai, and R. Sieg, Local -vectors of quasi-geometric and barycentric subdivisions, Discrete Comput. Geom. 61 (2019), no. 2, 364–379.
- [Kar19] K. Karu, Relative hard Lefschetz theorem for fans, Adv. Math. 347 (2019), 859–903.
- [KS16] E. Katz and A. Stapledon, Local -polynomials, invariants of subdivisions, and mixed Ehrhart theory, Adv. Math. 286 (2016), 181–239.
- [LPS22] M. Larson, S. Payne, and A. Stapledon, The local motivic monodromy conjecture for simplicial nondegenerate singularities, arXiv:2209.03553, 2022.
- [LVP11] A. Lemahieu and L. Van Proeyen, Monodromy conjecture for nondegenerate surface singularities, Trans. Amer. Math. Soc. 363 (2011), no. 9, 4801–4829.
- [Rei76] G. Reisner, Cohen-Macaulay quotients of polynomial rings, Adv. Math. 21 (1976), no. 1, 30–49.
- [Sta92] R. Stanley, Subdivisions and local -vectors, J. Amer. Math. Soc. 5 (1992), no. 4, 805–851.
- [Sta96] by same author, Combinatorics and commutative algebra, second ed., Progress in Mathematics, vol. 41, Birkhäuser Boston Inc., Boston, MA, 1996.
- [Sta17] A. Stapledon, Formulas for monodromy, Res. Math. Sci. 4 (2017), Paper No. 8, 42 pp.