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

Resolutions of local face modules, functoriality, and vanishing of local hh-vectors

Matt Larson, Sam Payne, and Alan Stapledon Stanford U. Department of Mathematics, 450 Jane Stanford Way, Stanford, CA 94305 [email protected] UT Department of Mathematics, 2515 Speedway, RLM 8.100, Austin, TX 78712 [email protected] Sydney Mathematics Research Institute, L4.42, Quadrangle A14, University of Sydney, NSW 2006, Australia [email protected]
Abstract.

We study the local face modules of triangulations of simplices, i.e., the modules over face rings whose Hilbert functions are local hh-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 hh-vectors and new results on the structure of faces in triangulations with vanishing local hh-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 hh-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 kk. Let σ:Γ2V\sigma\colon\Gamma\to 2^{V} be a quasi-geometric homology triangulation of a simplex, and let EE be a face of Γ\Gamma. Say that a face GΓG\in\Gamma is interior if σ(G)=V\sigma(G)=V, and let II be the ideal in the face ring k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)] generated by the faces that are interior relative to EE, i.e.,

I=(xF:FE is interior).I=(x^{F}:F\sqcup E\mbox{ is interior}).

Let d=|V||E|d=|V|-|E|, which is the Krull dimension of k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)], and let θ1,,θd\theta_{1},\ldots,\theta_{d} 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 L(Γ,E)L(\Gamma,E) is the image of II in k[lkΓ(E)]/(θ1,,θd)k[\operatorname{lk}_{\Gamma}(E)]/(\theta_{1},\ldots,\theta_{d}).

Note that L(Γ,E)L(\Gamma,E) is a finite dimensional graded kk-vector space. The local hh-vector is its Hilbert function:

(Γ,E):=(0,,d), where i:=dimL(Γ,E)i.\ell(\Gamma,E):=(\ell_{0},\ldots,\ell_{d}),\quad\quad\mbox{ where }\ \ell_{i}:=\dim L(\Gamma,E)_{i}.

The local face module L(Γ,E)L(\Gamma,E) depends on the choice of a special l.s.o.p., but (Γ,E)\ell(\Gamma,E) is an invariant of the triangulation with the symmetry i=di\ell_{i}=\ell_{d-i}. See §2.1 for details and references. In the past few years, there has been significant research activity on the combinatorics of local hh-vectors and relations to intersection homology [Ath16, KS16, Sta17, dCMM18]. Recent advances include a proof that every non-negative integer vector satisfying 0=0\ell_{0}=0 and i=di\ell_{i}=\ell_{d-i} is the local hh-vector of a quasi-geometric triangulation for E=E=\emptyset [JKMS19], and a relative hard Lefschetz theorem that yields unimodality of local hh-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 L(Γ,E)L(\Gamma,E) using methods of combinatorial commutative algebra. In particular, we describe natural combinatorial resolutions of these modules as well as natural maps of k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)]-modules, L(Γ,E)L(Γ,E)L(\Gamma,E)\to L(\Gamma,E^{\prime}), for EEE\subset E^{\prime}. Our first theorem gives explicit generators for the kernel of the natural map Ik[lkΓ(E)]/(θ1,,θd)I\to k[\operatorname{lk}_{\Gamma}(E)]/(\theta_{1},\ldots,\theta_{d}). Moreover, we extend this to an exact sequence of graded k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)]-modules in which each term is a direct sum of degree-shifted monomial ideals.

Label the vertices of the simplex V={v1,,vn}V=\{v_{1},\ldots,v_{n}\}. For a subset UVU\subset V, let Uc:=VUU^{c}:=V\smallsetminus U. After relabeling, we may assume that σ(E)c={v1,,vb}\sigma(E)^{c}=\{v_{1},\ldots,v_{b}\}. Given S{v1,,vd}S\subset\{v_{1},\ldots,v_{d}\}, we define the ideal ISk[lkΓ(E)]I_{S}\subset k[\operatorname{lk}_{\Gamma}(E)] by

IS:=(xF:σ(FE)cS).I_{S}:=(x^{F}:\,\sigma(F\sqcup E)^{c}\subset S).

Note that ISISI_{S^{\prime}}\subset I_{S} for SSS^{\prime}\subset S, and ISI_{S} depends only on S{v1,,vb}S\cap\{v_{1},\ldots,v_{b}\}. For instance, I=II_{\emptyset}=I and IS=k[lkΓ(E)]I_{S}=k[\operatorname{lk}_{\Gamma}(E)] if {v1,,vb}S\{v_{1},\ldots,v_{b}\}\subset S. By the definition of a special l.s.o.p. (Definition 2.3), after reordering, we may assume

supp(θi){wlkΓ(E):viσ(w)},\operatorname{supp}(\theta_{i})\subset\{w\in\operatorname{lk}_{\Gamma}(E):v_{i}\in\sigma(w)\},

for 1ib1\leq i\leq b. As a consequence, for any viSv_{i}\in S, multiplication by θi\theta_{i} induces a degree 1 map λi:ISIS{vi}\lambda_{i}\colon I_{S}\to I_{S\smallsetminus\{v_{i}\}}.

Theorem 1.2.

There is an exact sequence of graded k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)]-modules

0k[lkΓ(E)][d]|S|=d1IS[(d1)]|S|=1IS[1]IL(Γ,E)0,0\to k[\operatorname{lk}_{\Gamma}(E)][-d]\to\bigoplus_{|S|=d-1}I_{S}[-(d-1)]\to\dotsb\to\bigoplus_{|S|=1}I_{S}[-1]\to I\to L(\Gamma,E)\to 0,

where, for S={vi0,,vik}S=\{v_{i_{0}},\ldots,v_{i_{k}}\}, with i0<<iki_{0}<\cdots<i_{k}, the differential restricted to ISI_{S} is j=0k(1)jλij\oplus_{j=0}^{k}(-1)^{j}\lambda_{i_{j}}.

Corollary 1.3.

The kernel of the surjection IL(Γ,E)I\to L(\Gamma,E) is the ideal JJ generated by

{θixF:FE is interior }{θjxG:σ(GE)={vj}c, for 1jb}.\left\{\theta_{i}\cdot x^{F}:F\sqcup E\mbox{ is interior }\right\}\cup\left\{\theta_{j}\cdot x^{G}:\sigma(G\sqcup E)=\{v_{j}\}^{c},\mbox{ for }1\leq j\leq b\right\}.

We also construct maps between local face modules, as follows. For faces EEE\subset E^{\prime} in Γ\Gamma, let Star(EE)\operatorname{Star}(E^{\prime}\smallsetminus E) denote the closed star of EEE^{\prime}\smallsetminus E in lkΓ(E)\operatorname{lk}_{\Gamma}(E). We have a natural inclusion of complexes lkΓ(E)lkΓ(E)\operatorname{lk}_{\Gamma}(E^{\prime})\subset\operatorname{lk}_{\Gamma}(E).

Theorem 1.4.

Let EEE\subset E^{\prime} be faces of Γ\Gamma, with

d=n|E|,d=n|E|, and b=n|σ(E)|.d=n-|E|,\quad d^{\prime}=n-|E^{\prime}|,\quad\mbox{ and }\quad b^{\prime}=n-|\sigma(E^{\prime})|.

Let {θ1,,θd}\{\theta_{1},\dotsc,\theta_{d}\} be a special l.s.o.p. for k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)], and let θi:=θi|Star(EE)\theta_{i}^{\prime}:=\theta_{i}|_{\operatorname{Star}(E^{\prime}\smallsetminus E)}. Then there is a unique homomorphism of graded kk-algebras

ϕ:k[lkΓ(E)]/(θ1,,θd)k[lkΓ(E)]/(k[lkΓ(E)](θ1,,θd))\phi\colon k[\operatorname{lk}_{\Gamma}(E)]/(\theta_{1},\dotsc,\theta_{d})\to k[\operatorname{lk}_{\Gamma}(E^{\prime})]/(k[\operatorname{lk}_{\Gamma}(E^{\prime})]\cap(\theta_{1}^{\prime},\dotsc,\theta_{d}^{\prime}))

whose kernel contains {[xF]:FStar(EE)}\{[x^{F}]:F\not\in\operatorname{Star}(E^{\prime}\smallsetminus E)\} and satisfies ϕ(xF)=xF\phi(x^{F})=x^{F} for all FlkΓ(E)F\in\operatorname{lk}_{\Gamma}(E^{\prime}). Moreover, there is a special l.s.o.p. ζ1,,ζd\zeta_{1},\ldots,\zeta_{d^{\prime}} for k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E^{\prime})] such that (ζ1,,ζd)=k[lkΓ(E)](θ1,,θd)(\zeta_{1},\ldots,\zeta_{d^{\prime}})=k[\operatorname{lk}_{\Gamma}(E^{\prime})]\cap(\theta^{\prime}_{1},\ldots,\theta^{\prime}_{d}) and, up to reordering, we have θi|lkΓ(E)=ζi\theta_{i}|_{\operatorname{lk}_{\Gamma}(E^{\prime})}=\zeta_{i}, for 1ib1\leq i\leq b^{\prime}. With this choice of special l.s.o.p., ϕ(L(Γ,E))L(Γ,E)\phi(L(\Gamma,E))\subset L(\Gamma,E^{\prime}).

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. θ1,,θd\theta_{1},\ldots,\theta_{d}. Then L(Γ,E)L(\Gamma,E) is well-defined. For EEE^{\prime}\supset E the special l.s.o.p. ζ1,,ζd\zeta_{1},\ldots,\zeta_{d^{\prime}} depends on some choices, but the ideal that it generates does not, nor does the map ϕ:L(Γ,E)L(Γ,E)\phi\colon L(\Gamma,E)\to L(\Gamma,E^{\prime}). Moreover, for E′′EE^{\prime\prime}\supset E^{\prime}, one readily checks that the maps ϕ:L(Γ,E)L(Γ,E′′)\phi^{\prime}\colon L(\Gamma,E^{\prime})\to L(\Gamma,E^{\prime\prime}) and ϕ′′:L(Γ,E)L(Γ,E′′)\phi^{\prime\prime}\colon L(\Gamma,E)\to L(\Gamma,E^{\prime\prime}) are independent of all choices and satisfy ϕ′′=ϕϕ\phi^{\prime\prime}=\phi^{\prime}\circ\phi. Thus one obtains a functor from the poset of faces of Γ\Gamma that contain EE to graded vector spaces, given by EL(Γ,E)E^{\prime}\mapsto L(\Gamma,E^{\prime}).

We now give two applications of the above theorems. The first is a monotonicity property for local hh-vectors.

Theorem 1.6.

Let EEE\subset E^{\prime} be faces of Γ\Gamma such that σ(E)=σ(E)\sigma(E)=\sigma(E^{\prime}). Then (Γ,E)(Γ,E)\ell(\Gamma,E)\geq\ell(\Gamma,E^{\prime}).

The inequality in Theorem 1.6 is term by term, i.e., dimL(Γ,E)idimL(Γ,E)i\dim L(\Gamma,E)_{i}\geq\dim L(\Gamma,E^{\prime})_{i} for all ii. The proof is by showing that the map ϕ:L(Γ,E)L(Γ,E)\phi\colon L(\Gamma,E)\to L(\Gamma,E^{\prime}) 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 hh-vectors in the special case where E=E=\emptyset 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 EE 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 hh-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 hh-vectors. Let FlkΓ(E)F\in\operatorname{lk}_{\Gamma}(E) be a face such that FEF\sqcup E is interior. Following terminology from the monodromy conjecture literature (see, e.g., [LVP11]), we say that FF is a pyramid with apex wFw\in F if (FE)w(F\sqcup E)\smallsetminus w is not interior. Let

𝒜F:={wF:F is a pyramid with apex w}, and Vw:=σ((FE)w)c.\mathcal{A}_{F}:=\{w\in F:F\mbox{ is a pyramid with apex }w\},\mbox{ \ and \ }V_{w}:=\sigma((F\sqcup E)\smallsetminus w)^{c}.

The elements of VwV_{w} correspond to the base directions of FF, i.e., the facets of 2V2^{V} that contain the base of FF, when viewed as a pyramid with apex ww. We say FF is a UU-pyramid if there is an apex w𝒜Fw\in\mathcal{A}_{F} such that |Vw|=1|V_{w}|=1. In other words, a UU-pyramid is a pyramid with a unique base direction, for some choice of apex.

Definition 1.7.

Let FlkΓ(E)F\in\operatorname{lk}_{\Gamma}(E) be a face. An interior partition of FF is a decomposition

F=F1F2𝒜FF=F_{1}\sqcup F_{2}\sqcup\mathcal{A}_{F}

such that F1𝒜FEF_{1}\sqcup\mathcal{A}_{F}\sqcup E and F2𝒜FEF_{2}\sqcup\mathcal{A}_{F}\sqcup E are both interior.

Theorem 1.8.

Suppose (Γ,E)=0\ell(\Gamma,E)=0 and FlkΓ(E)F\in\operatorname{lk}_{\Gamma}(E) has an interior partition F=F1F2𝒜FF=F_{1}\sqcup F_{2}\sqcup\mathcal{A}_{F} such that |F1|2|F_{1}|\leq 2. Then FF is a UU-pyramid.

See Remark 3.2 for a short proof in a special case that illustrates the naturality of the UU-pyramid condition. The method of proof breaks down when |Fi|3|F_{i}|\geq 3. 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 σ(F2E)c=w𝒜FVw\sigma(F_{2}\sqcup E)^{c}=\bigcup_{w\in\mathcal{A}_{F}}V_{w}. But then the hypothesis that |F1|2|F_{1}|\leq 2 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 hh-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 hh-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 Γ\Gamma 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.

uu vv ww cc aa bb Γ\Gamma

Let xa:=x{a}x_{a}:=x^{\{a\}}, xb:=x{b}x_{b}:=x^{\{b\}}, xc:=x{c}x_{c}:=x^{\{c\}}, xu:=x{u}x_{u}:=x^{\{u\}}, xv:=x{v}x_{v}:=x^{\{v\}}, xw:=x{w}x_{w}:=x^{\{w\}}. Consider first E=E=\emptyset. The face ring is

k[lkΓ(E)]=k[xa,xb,xc,xu,xv,xw]/(xaxu,xbxv,xcxw,xuxv,xuxw,xvxw),k[\operatorname{lk}_{\Gamma}(E)]=k[x_{a},x_{b},x_{c},x_{u},x_{v},x_{w}]/(x_{a}x_{u},x_{b}x_{v},x_{c}x_{w},x_{u}x_{v},x_{u}x_{w},x_{v}x_{w}),

and its ideal of interior faces is

I=(xaxb,xaxc,xbxc).I=(x_{a}x_{b},x_{a}x_{c},x_{b}x_{c}).

A special l.s.o.p. is of the form θ1,θ2,θ3\theta_{1},\theta_{2},\theta_{3}, with

supp(θ1)={b,c,u},supp(θ2)={a,c,v},supp(θ3)={a,b,w},\operatorname{supp}(\theta_{1})=\{b,c,u\},\quad\quad\operatorname{supp}(\theta_{2})=\{a,c,v\},\quad\quad\operatorname{supp}(\theta_{3})=\{a,b,w\},

subject to the condition that the restrictions (of the corresponding affine linear functions) to the face {a,b,c}\{a,b,c\} are linearly independent. Our resolution of the local face module L(Γ,E)L(\Gamma,E) also involves the monomial ideals

Iu=(xa,xbxc),Iv=(xb,xaxc),Iw=(xc,xaxb),Iuv=(xa,xb,xw),Iuw=(xa,xc,xv),Ivw=(xb,xc,xu).\begin{array}[]{ccc}I_{u}=(x_{a},x_{b}x_{c}),&I_{v}=(x_{b},x_{a}x_{c}),&I_{w}=(x_{c},x_{a}x_{b}),\\ I_{uv}=(x_{a},x_{b},x_{w}),&I_{uw}=(x_{a},x_{c},x_{v}),&I_{vw}=(x_{b},x_{c},x_{u}).\end{array}

The resolution given by Theorem 1.2 is then

0k[lkΓ(E)][θ1θ2θ3]IvwIuwIuv[0θ3θ2θ30θ1θ2θ10]IuIvIw[θ1θ2θ3]IL(Γ,E)00\to k[\operatorname{lk}_{\Gamma}(E)]\xrightarrow{\begin{bmatrix}\theta_{1}\\ -\theta_{2}\\ \theta_{3}\end{bmatrix}}I_{vw}\oplus I_{uw}\oplus I_{uv}\xrightarrow{\begin{bmatrix}0&-\theta_{3}&-\theta_{2}\\ -\theta_{3}&0&\theta_{1}\\ \theta_{2}&\theta_{1}&0\\ \end{bmatrix}}I_{u}\oplus I_{v}\oplus I_{w}\xrightarrow{\begin{bmatrix}\theta_{1}&\theta_{2}&\theta_{3}\end{bmatrix}}I\to L(\Gamma,E)\to 0

In particular, we have L(Γ,E)I/JL(\Gamma,E)\cong I/J, where

(θ1xa,θ2xb,θ3xc)J.(\theta_{1}\cdot x_{a},\theta_{2}\cdot x_{b},\theta_{3}\cdot x_{c})\subset J.

Since θ1\theta_{1}, θ2\theta_{2}, and θ3\theta_{3} restrict to linearly independent functions on {a,b,c}\{a,b,c\}, the elements {θ1xa,θ2xb,θ3xc}\{\theta_{1}\cdot x_{a},\theta_{2}\cdot x_{b},\theta_{3}\cdot x_{c}\} span the 3-dimensional subspace xaxb,xaxc,xbxc\langle x_{a}x_{b},x_{a}x_{c},x_{b}x_{c}\rangle of k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)]. Hence I=JI=J and L(Γ,E)=0L(\Gamma,E)=0.

Next, consider E={c}E^{\prime}=\{c\}. Then

k[lkΓ(E)]=k[xa,xb,xu,xv]/(xaxu,xbxv,xuxv).k[\operatorname{lk}_{\Gamma}(E^{\prime})]=k[x_{a},x_{b},x_{u},x_{v}]/(x_{a}x_{u},x_{b}x_{v},x_{u}x_{v}).

A special l.s.o.p. is any l.s.o.p. of the form ζ1,ζ2\zeta_{1},\zeta_{2}, where supp(ζ1){a,b}\operatorname{supp}(\zeta_{1})\subset\{a,b\}. The ideal of interior faces in this case is I=(xa,xb)I^{\prime}=(x_{a},x_{b}), and the resolution given by Theorem 1.2 is

0k[lkΓ(E)][ζ2ζ1]k[lkΓ(E)]I[ζ1ζ2]IL(Γ,E)0.0\to k[\operatorname{lk}_{\Gamma}(E^{\prime})]\xrightarrow{\begin{bmatrix}-\zeta_{2}\\ \zeta_{1}\\ \end{bmatrix}}k[\operatorname{lk}_{\Gamma}(E^{\prime})]\oplus I^{\prime}\xrightarrow{\begin{bmatrix}\zeta_{1}&\zeta_{2}\end{bmatrix}}I^{\prime}\to L(\Gamma,E^{\prime})\to 0.

Note, in particular, that L(Γ,E)I/JL(\Gamma,E^{\prime})\cong I^{\prime}/J^{\prime}, where J=(ζ1,ζ2xa,ζ2xb)J^{\prime}=(\zeta_{1},\zeta_{2}x_{a},\zeta_{2}x_{b}). Thus one sees that L(Γ,E)L(\Gamma,E^{\prime}) has dimension 1 in degree 1, i.e., (Γ,E)=(0,1,0)\ell(\Gamma,E^{\prime})=(0,1,0).

Let us now consider Theorem 1.4 in this example. Let θi\theta^{\prime}_{i} denote the restriction of θi\theta_{i} to k[Star(EE)]k[\operatorname{Star}(E^{\prime}\smallsetminus E)]. Note that ζ1:=θ3\zeta_{1}:=\theta^{\prime}_{3} is supported on lkΓ(E)\operatorname{lk}_{\Gamma}(E^{\prime}). Extend {ζ1}\{\zeta_{1}\} to a basis for k[lkΓ(E)](θ1,θ2,θ3)k[\operatorname{lk}_{\Gamma}(E)]\cap(\theta^{\prime}_{1},\theta^{\prime}_{2},\theta^{\prime}_{3}), e.g., by choosing ζ2\zeta_{2} to be a linear combination of θ1\theta^{\prime}_{1} and θ2\theta^{\prime}_{2} in which the coefficient of xcx_{c} vanishes. Then ζ1,ζ2\zeta_{1},\zeta_{2} is a special l.s.o.p. for k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E^{\prime})], and the map ϕ\phi in Theorem 1.4 is given as follows. First, we set

ϕ(xa)=xa,ϕ(xb)=xb,ϕ(xu)=xu,ϕ(xv)=xv,ϕ(xw)=0.\phi(x_{a})=x_{a},\quad\phi(x_{b})=x_{b},\quad\phi(x_{u})=x_{u},\quad\phi(x_{v})=x_{v},\quad\phi(x_{w})=0.

Then, writing θ2=λcxc+λaxa+λvxv\theta_{2}=\lambda_{c}x_{c}+\lambda_{a}x_{a}+\lambda_{v}x_{v}, with all three coefficients nonzero, we set

ϕ(xc)=1λc(λaxa+λvxv).\phi(x_{c})=\frac{-1}{\lambda_{c}}(\lambda_{a}x_{a}+\lambda_{v}x_{v}).

Note that there is no subset of {θ1,θ2,θ3}\{\theta_{1},\theta_{2},\theta_{3}\} whose restrictions to k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E^{\prime})] form an l.s.o.p. This explains and motivates our two-step process for constructing the map: first restricting to Star(EE)\operatorname{Star}(E^{\prime}\smallsetminus E) and then intersecting with k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E^{\prime})] to produce the special l.s.o.p. that yields the functorial map ϕ:L(Γ,E)L(Γ,E)\phi\colon L(\Gamma,E)\to L(\Gamma,E^{\prime}).

Let also describe how Theorems 1.6 and 1.8 manifest in this example. For Theorem 1.8, observe that the face F={a,b}F=\{a,b\} in lkΓ(E)\operatorname{lk}_{\Gamma}(E^{\prime}) has an interior partition F={a}{b}F=\{a\}\sqcup\{b\}. The proof in this case shows that the classes of both xax_{a} and xbx_{b} are nonzero in L(Γ,E)L(\Gamma,E^{\prime}), for any choice of special l.s.o.p.

Finally, note that L(Γ,E)=0L(\Gamma,E)=0 and L(Γ,E)0L(\Gamma,E^{\prime})\neq 0, so there is no surjective map of graded vector space L(Γ,E)L(Γ,E)L(\Gamma,E)\to L(\Gamma,E^{\prime}). In this case, σ(E)σ(E)\sigma(E)\neq\sigma(E^{\prime}). Thus, we see that the hypothesis σ(E)=σ(E)\sigma(E)=\sigma(E^{\prime}) 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 kk. In particular, all rings are commutative kk-algebras and singular homology is computed with coefficients in kk.

2.1. Triangulations of simplices

In this section only, for the purposes of providing context, we allow that the field kk may be finite, and the triangulation σ:Γ2V\sigma\colon\Gamma\to 2^{V} is not necessarily quasi-geometric.

We recall the notion of a homology triangulation, following [Ath12b]. A dd-dimensional simplicial complex Γ\Gamma with trivial reduced homology is a homology ball of dimension dd if there is a subcomplex ΓΓ\partial\Gamma\subset\Gamma such that

  • Γ\partial\Gamma is a homology sphere of dimension d1d-1,

  • lkΓ(F)\operatorname{lk}_{\Gamma}(F) is a homology sphere of dimension d|F|d-|F| for FΓF\not\in\partial\Gamma.

  • lkΓ(F)\operatorname{lk}_{\Gamma}(F) is a homology ball of dimension d|F|d-|F| for all nonempty FΓF\in\partial\Gamma.

The interior faces of a homology ball Γ\Gamma are the faces not contained in Γ\partial\Gamma. A homology triangulation of the simplex 2V2^{V} is a finite simplicial complex Γ\Gamma and a map σ:Γ2V\sigma\colon\Gamma\to 2^{V} such that for every non-empty UVU\subset V,

  • the simplicial complex ΓU:=σ1(2U)\Gamma_{U}:=\sigma^{-1}(2^{U}) is a homology ball of dimension |U|1|U|-1.

  • σ1(U)\sigma^{-1}(U) is the set of interior faces of the homology ball σ1(2U)\sigma^{-1}(2^{U}).

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 kk. 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 FΓF\in\Gamma is σ(F)\sigma(F). A homology triangulation σ:Γ2V\sigma\colon\Gamma\to 2^{V} is quasi-geometric if there is no face FΓF\in\Gamma and UVU\subset V such that the dimension of ΓU\Gamma_{U} is strictly smaller than the dimension of FF and the carrier of every vertex in FF is contained in UU. A homology triangulation is geometric if it can be realized in n\mathbb{R}^{n} as the subdivision of a geometric simplex into geometric simplices. Every geometric homology triangulation is quasi-geometric.

The local hh-vector, which we have defined in the introduction as the Hilbert function of the local face module, can be expressed in terms of hh-vectors of subcomplexes of links of faces in the homology balls ΓU\Gamma_{U}:

(1) (Γ,E)=Uσ(E)(1)|V||U|h(lkΓU(E)).\ell(\Gamma,E)=\sum_{U\supset\sigma(E)}(-1)^{|V|-|U|}h(\operatorname{lk}_{\Gamma_{U}}(E)).

Note that (1) makes sense even when kk is finite or σ:Γ2V\sigma\colon\Gamma\to 2^{V} is not quasi-geometric, and should be taken as the definition of the local hh-vector in this broader context.

Theorem 2.1 ([Sta92, Ath12b, KS16]).

Let σ:Γ2V\sigma\colon\Gamma\to 2^{V} be a homology triangulation, let EE be a face of Γ\Gamma and let d=|V||E|d=|V|-|E|. Then the local hh-vector (0,,d)(\ell_{0},\ldots,\ell_{d}) satisfies:
    \bullet (symmetry) i=di;\ell_{i}=\ell_{d-i};     \bullet (non-negativity) if Γ\Gamma is quasi-geometric, then i0;\ell_{i}\geq 0;     \bullet (unimodality) if Γ\Gamma is regular, then 01d/2\ell_{0}\leq\ell_{1}\leq\cdots\leq\ell_{\lfloor d/2\rfloor}.

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 kk 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 kk is fixed and infinite, and all triangulations are quasi-geometric homology triangulations.

Given a finite simplicial complex Γ\Gamma with vertex set V={v1,,vn}V=\{v_{1},\ldots,v_{n}\}, let k[Γ]k[\Gamma] denote the face ring. In other words, for each subset FVF\subset V, let xFx^{F} be the corresponding squarefree monomial in the polynomial ring k[x1,,xn]k[x_{1},\ldots,x_{n}], i.e., xF:=viFxi.x^{F}:=\prod_{v_{i}\in F}x_{i}. Then the face ring is

k[Γ]:=k[x1,,xn]/(xF:F is not a face in Γ).k[\Gamma]:=k[x_{1},\ldots,x_{n}]/(x^{F}:F\mbox{ is not a face in }\Gamma).

Given a subcomplex Γ\Gamma^{\prime} of Γ\Gamma, we have a natural restriction map k[Γ]k[Γ]k[\Gamma]\rightarrow k[\Gamma^{\prime}], taking xFx^{F} to xFx^{F} if FΓF\in\Gamma^{\prime} and to 0 otherwise. Given θk[Γ]\theta\in k[\Gamma], let θ|Γ\theta|_{\Gamma^{\prime}} denote the image of θ\theta in k[Γ]k[\Gamma^{\prime}]. In particular, each FF in Γ\Gamma may be viewed as a subcomplex, and we write θ|F\theta|_{F} for the restriction of θ\theta to this subcomplex.

Note that k[Γ]k[\Gamma] is graded by degree. By definition, a linear system of parameters (l.s.o.p.) for a finitely generated graded kk-algebra RR of Krull dimension dd is a sequence of elements θ1,,θd\theta_{1},\ldots,\theta_{d} in R1R_{1} such that R/(θ1,,θd)R/(\theta_{1},\ldots,\theta_{d}) is a finite-dimensional kk-vector space. If Γ\Gamma is a Cohen-Macaulay complex (i.e., if k[Γ]k[\Gamma] is a Cohen-Macaulay ring) and θ1,,θd\theta_{1},\ldots,\theta_{d} is an l.s.o.p. for k[Γ]k[\Gamma], then (θ1,,θd)(\theta_{1},\ldots,\theta_{d}) is a regular sequence and the hh-polynomial of Γ\Gamma is the Hilbert series of k[Γ]/(θ1,,θd)k[\Gamma]/(\theta_{1},\ldots,\theta_{d}). Links of faces in triangulations of simplices are Cohen-Macaulay [Rei76].

Suppose Γ\Gamma has dimension d1d-1, so k[Γ]k[\Gamma] has Krull dimension dd. Then a sequence of elements θ1,,θd\theta_{1},\ldots,\theta_{d} in k[Γ]1k[\Gamma]_{1} is an l.s.o.p. for k[Γ]k[\Gamma] if and only if the following condition is satisfied [Sta96, Lemma 2.4(a)]:

  • ()(*)

    For every face FΓF\in\Gamma (or equivalently, for every facet FΓF\in\Gamma), the restrictions θ1|F,,θd|F\theta_{1}|_{F},\ldots,\theta_{d}|_{F} span a vector space of dimension |F||F|.

This characterization provides flexibility in constructing l.s.o.p.s in which the linear functions have specified support, where the support of θ=aixi\theta=\sum a_{i}x_{i} is supp(θ):={vi:ai0}\operatorname{supp}(\theta):=\{v_{i}:a_{i}\neq 0\}.

Lemma 2.2.

Let S1,,SdS_{1},\ldots,S_{d} be subsets of the vertices of Γ\Gamma. Then there is an l.s.o.p. θ1,,θd\theta_{1},\ldots,\theta_{d} for k[Γ]k[\Gamma] such that supp(θi)=Si\operatorname{supp}(\theta_{i})=S_{i} for 1id1\leq i\leq d if and only if, for every face FΓF\in\Gamma,

(2) |{Si:SiF}||F|.|\{S_{i}:S_{i}\cap F\neq\emptyset\}|\geq|F|.
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 S1,,SdS_{1},\ldots,S_{d} are chosen such that (2) holds for every FΓF\in\Gamma. Let N=|S1|++|Sd|N=|S_{1}|+\cdots+|S_{d}|, and consider the space kNk^{N} parametrizing tuples (θ1,,θd)(\theta_{1},\ldots,\theta_{d}) with supp(θi)Si\operatorname{supp}(\theta_{i})\subset S_{i}. Fix F={v1,,vk}ΓF=\{v_{1},\dotsc,v_{k}\}\in\Gamma. Let XFkNX_{F}\subset k^{N} parametrize the tuples whose restrictions to FF span a vector space of dimension |F||F|. Note that XFX_{F} is Zariski open. By Hall’s Marriage Theorem, there is a permutation σ𝔖d\sigma\in\mathfrak{S}_{d} such that viSσ(i)v_{i}\in S_{\sigma(i)}. If we set θσ(i)=xi\theta_{\sigma(i)}=x_{i} for 1ik1\leq i\leq k, and θσ(i)=0\theta_{\sigma(i)}=0 for i>ki>k, then θXF\theta\in X_{F}, and hence XFX_{F} is nonempty. Also, the subset of kNk^{N} where all coordinates are nonzero is Zariski open and nonempty. Since kk is infinite, the intersection of these nonempty Zariski open subsets of kNk^{N} is nonempty, and hence there is an l.s.o.p. θ1,,θd\theta_{1},\ldots,\theta_{d} with supp(θi)=Si\operatorname{supp}(\theta_{i})=S_{i}. ∎

Let σ:Γ2V\sigma\colon\Gamma\to 2^{V} be a quasi-geometric homology triangulation, and let EΓE\in\Gamma be a face.

Definition 2.3 ([Sta92, Ath12a]).

A linear system of parameters θ1,,θd\theta_{1},\dotsc,\theta_{d} for k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)] is special if, for each vertex vVv\in V with vσ(E)v\not\in\sigma(E), there is an element θv\theta_{v} of the l.s.o.p. such that supp(θv)\operatorname{supp}(\theta_{v}) consists of vertices in lkΓ(E)\operatorname{lk}_{\Gamma}(E) whose carrier contains vv, and such that θvθv\theta_{v}\not=\theta_{v^{\prime}} for vvv\not=v^{\prime}.

In other words, after reordering so that σ(E)c={v1,,vb}\sigma(E)^{c}=\{v_{1},\ldots,v_{b}\}, an l.s.o.p. for k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)] is special if we can order it θ1,,θd\theta_{1},\ldots,\theta_{d} such that

supp(θi){wlkΓ(E):viσ(w)},\operatorname{supp}(\theta_{i})\subset\{w\in\operatorname{lk}_{\Gamma}(E):v_{i}\in\sigma(w)\},

for 1ib1\leq i\leq b. 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 E=E=\emptyset. For completeness, we provide a short proof.

Proposition 2.4.

Suppose kk is infinite. Let σ:Γ2V\sigma\colon\Gamma\to 2^{V} be a quasi-geometric homology triangulation of a simplex, and let EE be a face of Γ\Gamma. Then there is a special l.s.o.p. for k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)].

Proof.

Let V={v1,,vn}V=\{v_{1},\ldots,v_{n}\}. After renumbering, we may assume that σ(E)c={v1,,vb}\sigma(E)^{c}=\{v_{1},\dotsc,v_{b}\}. Fix d=n|E|d=n-|E|. Note that bdb\leq d. We define subsets S1,S2,,SdS_{1},S_{2},\dotsc,S_{d} of the vertices in lkΓ(E)\operatorname{lk}_{\Gamma}(E), as follows. For ibi\leq b, let SiS_{i} be the set of vertices ww such that viσ(w)v_{i}\in\sigma(w). For i>bi>b, let SiS_{i} be the set of all vertices of lkΓ(E)\operatorname{lk}_{\Gamma}(E). Because σ\sigma is quasi-geometric, for each face FF of lkΓ(E)\operatorname{lk}_{\Gamma}(E), the union of the sets σ(w)V\sigma(w)\subset V, as ww ranges over vertices of EFE\sqcup F, has size at least |E|+|F||E|+|F|. It follows that |{ib:SiF}||F|(db)|\{i\leq b:S_{i}\cap F\neq\emptyset\}|\geq|F|-(d-b). Since SjFS_{j}\cap F\neq\emptyset for j>bj>b, we conclude that |{i:SiF}||F||\{i:S_{i}\cap F\neq\emptyset\}|\geq|F|. Hence, by Lemma 2.2, there is an l.s.o.p. θ1,,θd\theta_{1},\ldots,\theta_{d} for k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)] with supp(θi)=Si\operatorname{supp}(\theta_{i})=S_{i}. ∎

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 L(Γ,E)L(\Gamma,E) by a subcomplex of the Koszul resolution of k[lkΓ(E)]/(θ1,,θd)k[\operatorname{lk}_{\Gamma}(E)]/(\theta_{1},\dotsc,\theta_{d}). We continue to use the notation established above. In particular, σ:Γ2V\sigma\colon\Gamma\to 2^{V} is a quasi-geometric homology triangulation of the simplex with vertex set V={v1,,vn}V=\{v_{1},\ldots,v_{n}\}. We consider a face EΓE\in\Gamma with d=n|E|d=n-|E| and b=n|σ(E)|b=n-|\sigma(E)|. After reordering, we assume σ(E)c={v1,,vb}\sigma(E)^{c}=\{v_{1},\dotsc,v_{b}\}. For S{v1,,vd}S\subset\{v_{1},\ldots,v_{d}\}, we consider the ideal ISk[lkΓ(E)]I_{S}\subset k[\operatorname{lk}_{\Gamma}(E)] given by

IS:=(xF:σ(FE)cS).I_{S}:=(x^{F}:\,\sigma(F\sqcup E)^{c}\subset S).

Let θ1,θd\theta_{1},\ldots\theta_{d} be a special l.s.o.p. for k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)]. We may assume that

supp(θi){wlkΓ(E):viσ(w)},\operatorname{supp}(\theta_{i})\subset\{w\in\operatorname{lk}_{\Gamma}(E):v_{i}\in\sigma(w)\},

for 1ib1\leq i\leq b. For any viSv_{i}\in S, multiplication by θi\theta_{i} gives a map λi:ISIS{vi}\lambda_{i}\colon I_{S}\to I_{S\smallsetminus\{v_{i}\}}, and we consider the complex of graded k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)]-modules

(3) 0k[lkΓ(E)][d]|S|=d1IS[(d1)]|S|=1IS[1]IL(Γ,E)0,0\to k[\operatorname{lk}_{\Gamma}(E)][-d]\to\bigoplus_{|S|=d-1}I_{S}[-(d-1)]\to\dotsb\to\bigoplus_{|S|=1}I_{S}[-1]\to I\to L(\Gamma,E)\to 0,

in which the differential restricted to ISI_{S}, for S={vi0,,vik}S=\{v_{i_{0}},\ldots,v_{i_{k}}\}, with i0<<iki_{0}<\cdots<i_{k}, is j=0k(1)jλij\oplus_{j=0}^{k}(-1)^{j}\lambda_{i_{j}}.

Example 3.1.

If EE is an interior face of Γ\Gamma then every l.s.o.p. is special, IS=k[lkΓ(E)]I_{S}=k[\operatorname{lk}_{\Gamma}(E)] for all SS, and (3) is the Koszul resolution of L(Γ,E)=k[lkΓ(E)]/(θ1,,θd)L(\Gamma,E)=k[\operatorname{lk}_{\Gamma}(E)]/(\theta_{1},\ldots,\theta_{d}).

Proof of Theorem 1.2.

We must show (3) is exact. We begin by considering two complexes of k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)]-modules studied by Stanley and Athanasiadis. Recall that, for UVU\subset V, we write ΓU:=σ1(2U)\Gamma_{U}:=\sigma^{-1}(2^{U}).

Say Uσ(E)U\supset\sigma(E) and Uσ(E)={vi0,,vik}U\smallsetminus\sigma(E)=\{v_{i_{0}},\ldots,v_{i_{k}}\}, with i0<<iki_{0}<\cdots<i_{k}. For 0jk0\leq j\leq k, let ρj:k[lkΓU(E)]k[lkΓU{vij}(E)]\rho_{j}\colon k[\operatorname{lk}_{\Gamma_{U}}(E)]\to k[\operatorname{lk}_{\Gamma_{U\smallsetminus\{v_{i_{j}}\}}}(E)] be the restriction map. The first complex we consider is

(4) k[lkΓ(E)]{k[\operatorname{lk}_{\Gamma}(E)]}Uσ(E)|U|=n1k[lkΓU(E)]{\bigoplus\limits_{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-1\end{subarray}}k[\operatorname{lk}_{\Gamma_{U}}(E)]}Uσ(E)|U|=n2k[lkΓU(E)]{\bigoplus\limits_{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-2\end{subarray}}k[\operatorname{lk}_{\Gamma_{U}}(E)]}{\cdots}k[lkΓσ(E)(E)]{k[\operatorname{lk}_{\Gamma_{\sigma(E)}}(E)]}0,{0,}

in which the differential restricted to k[lkΓU(E)]k[\operatorname{lk}_{\Gamma_{U}}(E)] is j(1)jρj\bigoplus_{j}(-1)^{j}\rho_{j}. Next, we consider its quotient by (θ1,,θd)(\theta_{1},\ldots,\theta_{d}):

(5) k[lkΓ(E)](θ1,,θd){\frac{k[\operatorname{lk}_{\Gamma}(E)]}{(\theta_{1},\dotsc,\theta_{d})}}Uσ(E)|U|=n1k[lkΓU(E)](θ1,,θd){\bigoplus\limits_{\mathclap{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-1\end{subarray}}}\frac{k[\operatorname{lk}_{\Gamma_{U}}(E)]}{(\theta_{1},\dotsc,\theta_{d})}}𝒰σ()|𝒰|=𝓃2k[lkΓU(E)](θ1,,θd){\bigoplus\limits_{\mathcal{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-2\end{subarray}}}\frac{k[\operatorname{lk}_{\Gamma_{U}}(E)]}{(\theta_{1},\dotsc,\theta_{d})}}{\cdots}k[lkΓσ(E)(E)](θ1,,θd){\frac{k[\operatorname{lk}_{\Gamma_{\sigma(E)}}(E)]}{(\theta_{1},\dotsc,\theta_{d})}}0.{0.}

For any UVU\subset V, with Uσ(E)U\supset\sigma(E), let SUS_{U} be defined as

SU:=(U{v1,,vb}){vb+1,,vd}.S_{U}:=(U\cap\{v_{1},\ldots,v_{b}\})\cup\{v_{b+1},\ldots,v_{d}\}.

Then dimk[lkΓU(E)]=|SU|\dim k[\operatorname{lk}_{\Gamma_{U}}(E)]=|S_{U}| and it follows that the restriction of θi\theta_{i} to lkΓU(E)\operatorname{lk}_{\Gamma_{U}}(E) is nonzero if and only if viSUv_{i}\in S_{U}. Furthermore, {θi|lkΓU(E):viSU}\{\theta_{i}|_{\operatorname{lk}_{\Gamma_{U}}(E)}:v_{i}\in S_{U}\} is a special l.s.o.p. for k[lkΓU(E)]k[\operatorname{lk}_{\Gamma_{U}}(E)]. Stanley and Athanasiadis proved that both (4) and (5) are exact, and the kernel of the first arrow in (5) is L(Γ,E)L(\Gamma,E). (We will recall the proofs below.) Using the additivity of Hilbert functions in exact sequences, they deduced that the Hilbert function of L(Γ,E)L(\Gamma,E) 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 k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)]-modules. Since k[lkΓU(E)]k[\operatorname{lk}_{\Gamma_{U}}(E)] is Cohen-Macauley, the special l.s.o.p. {θi|lkΓU(E):viSU}\{\theta_{i}|_{\operatorname{lk}_{\Gamma_{U}}(E)}:v_{i}\in S_{U}\} is a regular sequence. Hence the corresponding Koszul complex KUK^{\bullet}_{U}

0{0}k[lkΓU(E)]SU{k[\operatorname{lk}_{\Gamma_{U}}(E)]_{S_{U}}}SSU|S|=|SU|1k[lkΓU(E)]S{\bigoplus\limits_{\mathclap{\begin{subarray}{c}S\subset S_{U}\\ |S|=|S_{U}|-1\end{subarray}}}k[\operatorname{lk}_{\Gamma_{U}}(E)]_{S}}{\cdots}SSU|S|=1k[lkΓU(E)]S{\bigoplus\limits_{\mathclap{\begin{subarray}{c}S\subset S_{U}\\ |S|=1\end{subarray}}}k[\operatorname{lk}_{\Gamma_{U}}(E)]_{S}}k[lkΓU(E)]{k[\operatorname{lk}_{\Gamma_{U}}(E)]}k[lkΓU(E)](θ1,,θd){\frac{k[\operatorname{lk}_{\Gamma_{U}}(E)]}{(\theta_{1},\dotsc,\theta_{d})}}0,{0,}

is exact. Here, for a graded module MM and a finite set SS, we write MS:=M[|S|]M_{S}:=M[-|S|]. Replacing each term in (5) with its corresponding Koszul resolution, gives a complex of complexes

(6) KV{K_{V}^{\bullet}}Uσ(E)|U|=n1KU{\bigoplus\limits_{\mathclap{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-1\end{subarray}}}K_{U}^{\bullet}}Uσ(E)|U|=n2KU{\bigoplus\limits_{\mathclap{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-2\end{subarray}}}K_{U}^{\bullet}}{\cdots}Kσ(E){K_{\sigma(E)}^{\bullet}}0,{0,}

which may be expanded as the commuting double complex shown in Figure 1.

0{0}0{0}0{0}{\cdots}0{0}k[lkΓ(E)](θ1,,θd){\frac{k[\operatorname{lk}_{\Gamma}(E)]}{(\theta_{1},\dotsc,\theta_{d})}}Uσ(E)|U|=n1k[lkΓU(E)](θ1,,θd){\bigoplus\limits_{\mathclap{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-1\end{subarray}}}\frac{k[\operatorname{lk}_{\Gamma_{U}}(E)]}{(\theta_{1},\dotsc,\theta_{d})}}𝒰σ()|𝒰|=𝓃2k[lkΓU(E)](θ1,,θd){\bigoplus\limits_{\mathcal{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-2\end{subarray}}}\frac{k[\operatorname{lk}_{\Gamma_{U}}(E)]}{(\theta_{1},\dotsc,\theta_{d})}}{\cdots}k[lkΓσ(E)(E)](θ1,,θd){\frac{k[\operatorname{lk}_{\Gamma_{\sigma(E)}}(E)]}{(\theta_{1},\dotsc,\theta_{d})}}0{0}k[lkΓ(E)]{k[\operatorname{lk}_{\Gamma}(E)]}Uσ(E)|U|=n1k[lkΓU(E)]{\bigoplus\limits_{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-1\end{subarray}}k[\operatorname{lk}_{\Gamma_{U}}(E)]}Uσ(E)|U|=n2k[lkΓU(E)]{\bigoplus\limits_{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-2\end{subarray}}k[\operatorname{lk}_{\Gamma_{U}}(E)]}{\cdots}k[lkΓσ(E)(E)]{k[\operatorname{lk}_{\Gamma_{\sigma(E)}}(E)]}0{0}|S|=1k[lkΓ(E)]S{\bigoplus\limits_{\mathclap{\begin{subarray}{c}|S|=1\end{subarray}}}k[\operatorname{lk}_{\Gamma}(E)]_{S}}Uσ(E)|U|=n1SSU|S|=1k[lkΓU(E)]S{\bigoplus\limits_{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-1\end{subarray}}\quad\bigoplus\limits_{\mathclap{\begin{subarray}{c}S\subset S_{U}\\ |S|=1\end{subarray}}}k[\operatorname{lk}_{\Gamma_{U}}(E)]_{S}}Uσ(E)|U|=n2SSU|S|=1k[lkΓU(E)]S{\bigoplus\limits_{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-2\end{subarray}}\quad\bigoplus\limits_{\mathclap{\begin{subarray}{c}S\subset S_{U}\\ |S|=1\end{subarray}}}k[\operatorname{lk}_{\Gamma_{U}}(E)]_{S}}{\cdots}SSσ(E)|S|=1k[lkΓσ(E)(E)]S{\bigoplus\limits_{\mathclap{\begin{subarray}{c}S\subset S_{\sigma(E)}\\ |S|=1\end{subarray}}}k[\operatorname{lk}_{\Gamma_{\sigma(E)}}(E)]_{S}}0{0}|S|=2k[lkΓ(E)]S{\bigoplus\limits_{\mathclap{\begin{subarray}{c}|S|=2\end{subarray}}}k[\operatorname{lk}_{\Gamma}(E)]_{S}}Uσ(E)|U|=n1SSU|S|=2k[lkΓU(E)]S{\bigoplus\limits_{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-1\end{subarray}}\quad\bigoplus\limits_{\mathclap{\begin{subarray}{c}S\subset S_{U}\\ |S|=2\end{subarray}}}k[\operatorname{lk}_{\Gamma_{U}}(E)]_{S}}Uσ(E)|U|=n2SSU|S|=2k[lkΓU(E)]S{\bigoplus\limits_{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-2\end{subarray}}\quad\bigoplus\limits_{\mathclap{\begin{subarray}{c}S\subset S_{U}\\ |S|=2\end{subarray}}}k[\operatorname{lk}_{\Gamma_{U}}(E)]_{S}}{\cdots}SSσ(E)|S|=2k[lkΓσ(E)(E)]S{\bigoplus\limits_{\mathclap{\begin{subarray}{c}S\subset S_{\sigma(E)}\\ |S|=2\end{subarray}}}k[\operatorname{lk}_{\Gamma_{\sigma(E)}}(E)]_{S}}0.{0.}{\vdots}{\vdots}{\vdots}{\cdots}{\vdots}|S|=d1k[lkΓ(E)]S{\bigoplus\limits_{\mathclap{\begin{subarray}{c}|S|=d-1\end{subarray}}}k[\operatorname{lk}_{\Gamma}(E)]_{S}}Uσ(E)|U|=n1k[lkΓ(E)]SU{\bigoplus\limits_{\mathclap{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-1\end{subarray}}}k[\operatorname{lk}_{\Gamma}(E)]_{S_{U}}}0{0}k[lkΓ(E)]{v1,,vd}{k[\operatorname{lk}_{\Gamma}(E)]_{\{v_{1},\ldots,v_{d}\}}}0{0}0{0}
Figure 1. The double complex obtained by taking the Koszul resolution of (5).

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 SS of {v1,,vd}\{v_{1},\ldots,v_{d}\}, and consider the piece of the complex indexed by SS:

(7) k[lkΓ(E)]S{k[\operatorname{lk}_{\Gamma}(E)]_{S}}SSU|U|=n1k[lkΓU(E)]S{\bigoplus\limits_{\begin{subarray}{c}S\subset S_{U}\\ |U|=n-1\end{subarray}}k[\operatorname{lk}_{\Gamma_{U}}(E)]_{S}}SSU|U|=n2k[lkΓU(E)]S{\bigoplus\limits_{\begin{subarray}{c}S\subset S_{U}\\ |U|=n-2\end{subarray}}k[\operatorname{lk}_{\Gamma_{U}}(E)]_{S}}{\cdots}0.{0.}

When S=S=\emptyset, we obtain (4). Observe that the complex (7) is multigraded by m\mathbb{N}^{m}, where mm is the number of vertices of lkΓ(E)\operatorname{lk}_{\Gamma}(E). Explicitly, degx1α1xmαm=(α1,,αm)\deg x_{1}^{\alpha_{1}}\dotsb x_{m}^{\alpha_{m}}=(\alpha_{1},\dotsc,\alpha_{m}). Therefore it suffices to show exactness on graded pieces. Fix α=(α1,,αm)\alpha=(\alpha_{1},\dotsc,\alpha_{m}). By the definition of the face ring, every term of (7) will have 0 in the graded piece corresponding to α\alpha unless the set of vertices with αi0\alpha_{i}\not=0 forms a face FF, in which case the α\alpha-graded part can be identified with the augmented cochain complex of a simplex, indexed by all UU that contain σ(E)σ(F)S\sigma(E)\cup\sigma(F)\cup S, and hence is exact.

We now recall the proof that the top row of the double complex, (5), is exact.

k[lkΓ(E)](θ1,,θd){\frac{k[\operatorname{lk}_{\Gamma}(E)]}{(\theta_{1},\dotsc,\theta_{d})}}Uσ(E)|U|=n1k[lkΓU(E)](θ1,,θd){\bigoplus\limits_{\mathclap{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-1\end{subarray}}}\frac{k[\operatorname{lk}_{\Gamma_{U}}(E)]}{(\theta_{1},\dotsc,\theta_{d})}}𝒰σ()|𝒰|=𝓃2k[lkΓU(E)](θ1,,θd){\bigoplus\limits_{\mathcal{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-2\end{subarray}}}\frac{k[\operatorname{lk}_{\Gamma_{U}}(E)]}{(\theta_{1},\dotsc,\theta_{d})}}{\cdots}k[lkΓσ(E)(E)](θ1,,θd){\frac{k[\operatorname{lk}_{\Gamma_{\sigma(E)}}(E)]}{(\theta_{1},\dotsc,\theta_{d})}}0{0}

The proof involves showing that the quotients of (4) by (θd,,θd(r1))(\theta_{d},\dotsc,\theta_{d-(r-1)}) is exact by induction on rr. The case of r=0r=0 is the exactness of the second row.

Now assume that (4) remains exact after quotienting by (θd,,θd(r1))(\theta_{d},\dotsc,\theta_{d-(r-1)}). Let CiC^{i} denote the iith term of (4) tensored with k[lkΓ(E)]/(θd,,θd(r1))k[\operatorname{lk}_{\Gamma}(E)]/(\theta_{d},\dotsc,\theta_{d-(r-1)}). By the induction hypothesis, we have an exact sequence

C:C0C1Cb0.C^{\bullet}\colon\enskip C^{0}\to C^{1}\to\dotsb\to C^{b}\to 0.

Set m=drm=d-r. Recall that θi=0k[lkΓU(E)]\theta_{i}=0\in k[\operatorname{lk}_{\Gamma_{U}}(E)] if viSUv_{i}\notin S_{U}, and that {θi|lkΓU(E):viSU}\{\theta_{i}|_{\operatorname{lk}_{\Gamma_{U}}(E)}:v_{i}\in S_{U}\} is a special l.s.o.p. for k[lkΓU(E)]k[\operatorname{lk}_{\Gamma_{U}}(E)]. Also, for σ(E)U\sigma(E)\subset U, vmSUv_{m}\notin S_{U} if and only if vmUv_{m}\notin U. Hence, we have an exact sequence

(8) 0BCθmCC/(θm)0,0\to B^{\bullet}\to C^{\bullet}\xrightarrow{\theta_{m}}C^{\bullet}\to C^{\bullet}/(\theta_{m})\to 0,

where

Bi=Uσ(E),|U|=nivmUk[lkΓU(E)]/(θd,,θm+1).B^{i}=\bigoplus_{\begin{subarray}{c}U\supset\sigma(E),\enskip|U|=n-i\\ v_{m}\not\in U\end{subarray}}k[\operatorname{lk}_{\Gamma_{U}}(E)]/(\theta_{d},\dotsc,\theta_{m+1}).

For example, when m>bm>b, vmσ(E)v_{m}\in\sigma(E) and B=0B^{\bullet}=0. Up to signs and a degree shift, we can then identify BB^{\bullet} with the complex (4) for Γ|{vm}c\Gamma|_{\{v_{m}\}^{c}} quotiented by (θd,,θm+1)(\theta_{d},\dotsc,\theta_{m+1}). Then BB^{\bullet} is exact by the induction hypothesis applied to Γ|{vm}c\Gamma|_{\{v_{m}\}^{c}}. By breaking (8) up into two short exact sequences we see that Hi(C/(θm))Hi+2(B)=0H^{i}(C^{\bullet}/(\theta_{m}))\cong H^{i+2}(B^{\bullet})=0 as desired.

Now that we know the exactness of (6), let

A=ker(KVUσ(E)|U|=n1KU).\begin{split}A^{\bullet}&=\ker\Bigg{(}K_{V}^{\bullet}\to\bigoplus\limits_{\mathclap{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-1\end{subarray}}}K_{U}^{\bullet}\Bigg{)}.\end{split}

Then, by construction, we have an exact sequence of complexes

0{0}A{A^{\bullet}}KV{K_{V}^{\bullet}}Uσ(E)|U|=n1KU{\bigoplus\limits_{\mathclap{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-1\end{subarray}}}K_{U}^{\bullet}}Uσ(E)|U|=n2KU{\bigoplus\limits_{\mathclap{\begin{subarray}{c}U\supset\sigma(E)\\ |U|=n-2\end{subarray}}}K_{U}^{\bullet}}{\cdots}Kσ(E){K_{\sigma(E)}^{\bullet}}0.{0.}

As above, we repeatedly apply the long exact sequence on cohomology to see that AA^{\bullet} is exact. We may then identify AA^{\bullet} with the exact sequence

0k[lkΓ(E)][n]|S|=d1IS[(n1)]|S|=1IS[1]IA00.0\to k[\operatorname{lk}_{\Gamma}(E)][-n]\to\oplus_{|S|=d-1}I_{S}[-(n-1)]\to\dotsb\to\oplus_{|S|=1}I_{S}[-1]\to I\to A^{0}\to 0.

Since II surjects onto A0A^{0} and A0k[lkΓ(E)]/(θ1,,θd)A^{0}\subset k[\operatorname{lk}_{\Gamma}(E)]/(\theta_{1},\dotsc,\theta_{d}), we conclude that A0=L(Γ,E)A^{0}=L(\Gamma,E), as required. ∎

Remark 3.2.

Let σ:Γ2V\sigma\colon\Gamma\to 2^{V} be a quasi-geometric homology triangulation of a simplex, and let EE be a face of Γ\Gamma. Let FlkΓ(E)F\in\operatorname{lk}_{\Gamma}(E) such that FEF\sqcup E is interior, and suppose that F=𝒜FF=\mathcal{A}_{F} is an interior partition of FF, i.e., with F1=F2=F_{1}=F_{2}=\emptyset. Suppose that FF is not a UU-pyramid. By Corollary 1.3, JJ is generated by elements of the form θixF\theta_{i}\cdot x^{F} for FEF\sqcup E interior or θjxG\theta_{j}\cdot x^{G} for some GG with σ(GE)={vj}c\sigma(G\sqcup E)=\{v_{j}\}^{c}. Because FF is not a UU-pyramid, no monomial appearing in any of these generators divides xFx^{F}, so xFx^{F} is nonzero in L(Γ,E)L(\Gamma,E). This proves Theorem 1.8 in the special case when F1=F2=F_{1}=F_{2}=\emptyset.

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 σ:Γ2V\sigma\colon\Gamma\to 2^{V}, and let EEE\subset E^{\prime} be faces of Γ\Gamma.

Lemma 4.1.

Let RR be a graded kk-algebra with R0=kR_{0}=k. Let {θ1,,θn}\{\theta_{1},\dotsc,\theta_{n}\} be an l.s.o.p. for R[x1,,xm]R[x_{1},\dotsc,x_{m}], where each xjx_{j} has degree 11. Then there is a unique graded RR-algebra isomorphism

ϕ:R[x1,,xm]/(θ1,,θn)R/R(θ1,,θn).\phi\colon R[x_{1},\dotsc,x_{m}]/(\theta_{1},\dotsc,\theta_{n})\rightarrow R/R\cap(\theta_{1},\dotsc,\theta_{n}).

Moreover, any kk-basis for R1(θ1,,θn)R_{1}\cap(\theta_{1},\dotsc,\theta_{n}) is an l.s.o.p. for RR and generates R(θ1,,θn)R\cap(\theta_{1},\dotsc,\theta_{n}).

Proof.

Consider the exact sequence of kk-linear maps

0R1R[x1,,xm]1(x1,,xm)10,0\rightarrow R_{1}\rightarrow R[x_{1},\dotsc,x_{m}]_{1}\rightarrow(x_{1},\dotsc,x_{m})_{1}\rightarrow 0,

where the right hand map takes r+iαixir+\sum_{i}\alpha_{i}x_{i} to iαixi\sum_{i}\alpha_{i}x_{i}, for any rR1r\in R_{1} and αik\alpha_{i}\in k. This restricts to an exact sequence of kk-linear maps

0R1(θ1,,θn)1(θ1,,θn)1(x1,,xm)10,0\rightarrow R_{1}\cap(\theta_{1},\dotsc,\theta_{n})_{1}\rightarrow(\theta_{1},\dotsc,\theta_{n})_{1}\rightarrow(x_{1},\dotsc,x_{m})_{1}\rightarrow 0,

where the surjectivity of the right-hand map follows from the fact that θ1,,θn\theta_{1},\dotsc,\theta_{n} is an l.s.o.p. Hence, for 1im1\leq i\leq m, we can write xi=ri+six_{i}=r_{i}+s_{i}, for some riR1r_{i}\in R_{1} and si(θ1,,θn)1s_{i}\in(\theta_{1},\dotsc,\theta_{n})_{1}. For any RR-algebra map ϕ:R[x1,,xm]/(θ1,,θn)R/R(θ1,,θn),\phi\colon R[x_{1},\dotsc,x_{m}]/(\theta_{1},\dotsc,\theta_{n})\rightarrow R/R\cap(\theta_{1},\dotsc,\theta_{n}), we must have that ϕ(xi)=ri\phi(x_{i})=r_{i}, so there is a unique such map. On the other hand, the RR-algebra homomorphism defined by ϕ(xi)=ri\phi(x_{i})=r_{i} is well-defined, since if xi=ri+six_{i}=r_{i}^{\prime}+s_{i}^{\prime}, for some riR1r_{i}^{\prime}\in R_{1} and si(θ1,,θn)1s_{i}^{\prime}\in(\theta_{1},\dotsc,\theta_{n})_{1}, then ririR1(θ1,,θn)1r_{i}-r_{i}^{\prime}\in R_{1}\cap(\theta_{1},\dotsc,\theta_{n})_{1}. Note that the unique RR-algebra homomorphism from R/R(θ1,,θn)R/R\cap(\theta_{1},\dotsc,\theta_{n}) to R[x1,,xm]/(θ1,,θn)R[x_{1},\dotsc,x_{m}]/(\theta_{1},\dotsc,\theta_{n}) is the inverse of ϕ\phi.

Since ϕ\phi is an isomorphism and factors through R/(R1(θ1,,θn)1)R/(R_{1}\cap(\theta_{1},\dotsc,\theta_{n})_{1}), we conclude that the RR-ideal R(θ1,,θn)R\cap(\theta_{1},\dotsc,\theta_{n}) is generated in degree 11 and hence any kk-basis for R1(θ1,,θn)R_{1}\cap(\theta_{1},\dotsc,\theta_{n}) is an l.s.o.p. for RR. ∎

Proof of Theorem 1.4.

Note that Star(EE)\operatorname{Star}(E^{\prime}\smallsetminus E) is the join of EEE^{\prime}\smallsetminus E with lkΓ(E)\operatorname{lk}_{\Gamma}(E^{\prime}). The face ring k[Star(EE)]k[\operatorname{Star}(E^{\prime}\smallsetminus E)] is therefore a polynomial ring over k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E^{\prime})]. Its Krull dimension is equal to d=dimk[lkΓ(E)]d=\dim k[\operatorname{lk}_{\Gamma}(E)], and hence the restrictions θ1,,θd\theta^{\prime}_{1},\ldots,\theta^{\prime}_{d} form an l.s.o.p., where θi:=θi|Star(EE)\theta^{\prime}_{i}:=\theta_{i}|_{\operatorname{Star}(E^{\prime}\smallsetminus E)}. By Lemma 4.1, there is a unique graded k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E^{\prime})]-algebra homomorphism k[Star(EE)]/(θ1,,θd)k[lkΓ(E)]/(k[lkΓ(E)](θ1,,θd))k[\operatorname{Star}(E^{\prime}\smallsetminus E)]/(\theta^{\prime}_{1},\ldots,\theta^{\prime}_{d})\to k[\operatorname{lk}_{\Gamma}(E^{\prime})]/(k[\operatorname{lk}_{\Gamma}(E^{\prime})]\cap(\theta^{\prime}_{1},\ldots,\theta^{\prime}_{d})), which lifts to the unique homomorphism ϕ\phi in the statement of the theorem. It remains to construct a special l.s.o.p. for k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E^{\prime})] with the specified properties.

After reordering, we may assume that

σ(E)c={v1,,vb},supp(θi){w:viσ(w)}, for 1ib, and σ(E)c={v1,,vb}.\sigma(E)^{c}=\{v_{1},\ldots,v_{b}\},\quad\operatorname{supp}(\theta_{i})\subset\{w:v_{i}\in\sigma(w)\},\ \mbox{ for }1\leq i\leq b,\quad\mbox{ and }\sigma(E^{\prime})^{c}=\{v_{1},\ldots,v_{b^{\prime}}\}.

Note, in particular, that θi\theta^{\prime}_{i} is supported on vertices in the link of EE^{\prime}, for 1ib1\leq i\leq b^{\prime}. By Lemma 4.1, any kk-basis for k[lkΓ(E)](θ1,,θd)k[\operatorname{lk}_{\Gamma}(E^{\prime})]\cap(\theta^{\prime}_{1},\ldots,\theta^{\prime}_{d}) is an l.s.o.p. for k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E^{\prime})]. Set ζi=θi|lkΓ(E)\zeta_{i}=\theta_{i}|_{\operatorname{lk}_{\Gamma}(E^{\prime})}, for 1ib1\leq i\leq b^{\prime}, and note that {ζ1,,ζb}\{\zeta_{1},\ldots,\zeta_{b^{\prime}}\} is linearly independent. Extending this independent set to a basis produces a special l.s.o.p. for k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E^{\prime})]. It remains to verify that ϕ(L(Γ,E))L(Γ,E)\phi(L(\Gamma,E))\subset L(\Gamma,E^{\prime}). Let FlkΓ(E)F\in\operatorname{lk}_{\Gamma}(E) be a face with FEF\sqcup E interior. If FF is not in Star(EE)\operatorname{Star}(E^{\prime}\smallsetminus E), then ϕ(xF)=0\phi(x^{F})=0. Otherwise, FF can be written uniquely as the join of possibly empty faces F1EEF_{1}\subset E^{\prime}\smallsetminus E and F2lkΓ(E)F_{2}\in\operatorname{lk}_{\Gamma}(E^{\prime}). Then F2EF_{2}\sqcup E^{\prime} is interior, and ϕ(xF)=ϕ(xF1)xF2(xF2)\phi(x^{F})=\phi(x^{F_{1}})x^{F_{2}}\in(x^{F_{2}}). Hence ϕ(xF)L(Γ,E)\phi(x^{F})\in L(\Gamma,E^{\prime}), as required. ∎

Proof of Theorem 1.6.

Let EEE\subset E^{\prime} be faces of a quasi-geometric homology triangulation Γ\Gamma of a simplex, and assume that σ(E)=σ(E)\sigma(E)=\sigma(E^{\prime}). It is enough to show that the induced map ϕ:L(Γ,E)L(Γ,E)\phi\colon L(\Gamma,E)\to L(\Gamma,E^{\prime}) given by Theorem 1.4 is surjective. Note that L(Γ,E)L(\Gamma,E^{\prime}) is generated by the monomials xFx^{F} such that FlkΓ(E)F\in\operatorname{lk}_{\Gamma}(E^{\prime}) and FEF\sqcup E^{\prime} is interior. If FF is such a face, then it is also in the link of EE and, since σ(E)=σ(E)\sigma(E)=\sigma(E^{\prime}), the face (FE)<(FE)(F\sqcup E)<(F\sqcup E^{\prime}) is also interior. Then ϕ(xF)=xF\phi(x^{F})=x^{F}, 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 L(Γ,E)L(\Gamma,E) implies the vanishing of a restricted local face module L(Γ,𝒜FE)|F1F2,L(\Gamma,\mathcal{A}_{F}\sqcup E)|_{F_{1}\sqcup F_{2}}, for certain interior partitions F1F2𝒜FF_{1}\sqcup F_{2}\sqcup\mathcal{A}_{F}. We then develop algebraic arguments, inspired by ideas from [dMGP+20], to show that FF being a UU-pyramid is necessary for the vanishing of the restricted local face module when |F1|2|F_{1}|\leq 2 and thus prove Theorem 1.8.

We use the notation introduced in the introduction. Let Δ\Delta be a subcomplex of lkΓ(E)\operatorname{lk}_{\Gamma}(E). For any k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)]-module MM, the restriction of MM to Δ\Delta is M|Δ:=Mk[lkΓ(E)]k[Δ]M|_{\Delta}:=M\otimes_{k[\operatorname{lk}_{\Gamma}(E)]}k[\Delta], where k[Δ]k[\Delta] is a k[lkΓ(E)]k[\operatorname{lk}_{\Gamma}(E)]-module via the restriction map. By the resolution of L(Γ,E)L(\Gamma,E) in Theorem 1.2 and the right exactness of tensoring with k[Δ]k[\Delta], we have an exact sequence

(9) |S|=1IS|Δ[1]I|ΔL(Γ,E)|Δ0.\bigoplus_{|S|=1}I_{S}|_{\Delta}[-1]\to I|_{\Delta}\to L(\Gamma,E)|_{\Delta}\to 0.

Recall from Corollary 1.3 that L(Γ,E)I/JL(\Gamma,E)\cong I/J, where JJ is the ideal generated by {θixF:FE is interior}\{\theta_{i}x^{F}:F\sqcup E\mbox{ is interior}\} and {θjxG:σ(GE)={vj}c}\{\theta_{j}x^{G}:\sigma(G\sqcup E)=\{v_{j}\}^{c}\}. Hence, L(Γ,E)|ΔI|Δ/J|ΔL(\Gamma,E)|_{\Delta}\cong I|_{\Delta}/J|_{\Delta}, where I|Δ,J|ΔI|_{\Delta},J|_{\Delta} are the k[Δ]k[\Delta]-ideals

(10) I|Δ=(xH:HΔ,σ(HE)=V),I|_{\Delta}=(x^{H}:H\subset\Delta,\sigma(H\sqcup E)=V),
(11) J|Δ=(θ1|Δ,,θd|Δ)I|Δ+(θj|ΔxG:GΔ,σ(GE)={vj}c).J|_{\Delta}=(\theta_{1}|_{\Delta},\ldots,\theta_{d}|_{\Delta})\cdot I|_{\Delta}+(\theta_{j}|_{\Delta}x^{G}:\enskip G\subset\Delta,\enskip\sigma(G\sqcup E)=\{v_{j}\}^{c}).

For example, if FF is a face of lkΓ(E)\operatorname{lk}_{\Gamma}(E), then k[F]k[F] is a polynomial ring with variables indexed by the vertices of FF, and L(Γ,E)|FL(\Gamma,E)|_{F} is identified with a quotient of ideals in this polynomial ring.

Lemma 5.1.

Let σ:Γ2V\sigma\colon\Gamma\to 2^{V} be a quasi-geometric homology triangulation of a simplex, and let EE be a face of Γ\Gamma. Let FlkΓ(E)F\in\operatorname{lk}_{\Gamma}(E) be a face with FEF\sqcup E interior. Assume that FF is not a UU-pyramid. Then there is a surjective graded k[F]k[F]-module homomorphism

L(Γ,E)|FL(Γ,𝒜FE)|F𝒜F[|𝒜F|],L(\Gamma,E)|_{F}\to L(\Gamma,\mathcal{A}_{F}\sqcup E)|_{F\smallsetminus\mathcal{A}_{F}}[-|\mathcal{A}_{F}|],

where the second term is a k[F]k[F]-module via the restriction map k[F]k[F𝒜F]k[F]\mapsto k[F\smallsetminus\mathcal{A}_{F}].

Proof.

If Δ\Delta is a subcomplex of lkΓ(E)\operatorname{lk}_{\Gamma}(E) contained in the closed star of 𝒜F\mathcal{A}_{F}, then x𝒜Fx^{\mathcal{A}_{F}} is a non-zero divisor in k[Δ]k[\Delta]. In particular, x𝒜Fx^{\mathcal{A}_{F}} is a non-zero divisor in k[F]k[F] (this is also clear since k[F]k[F] is a polynomial ring). Note that every face of FF with carrier codimension at most 11 contains 𝒜F\mathcal{A}_{F}. Thus I|F=x𝒜FMI|_{F}=x^{\mathcal{A}_{F}}\cdot M and J|F=x𝒜FNJ|_{F}=x^{\mathcal{A}_{F}}\cdot N, where MM and NN are the ideals in k[F]k[F]

M=(xH:HF𝒜F,σ(H𝒜FE)=V),M=(x^{H}:\enskip H\subset F\smallsetminus\mathcal{A}_{F},\enskip\sigma(H\sqcup\mathcal{A}_{F}\sqcup E)=V),
N=(θ1|F,,θd|F)M+(θj|FxG:GF𝒜F,σ(G𝒜FE)={vj}c).N=(\theta_{1}|_{F},\ldots,\theta_{d}|_{F})\cdot M+(\theta_{j}|_{F}x^{G}:\enskip G\subset F\smallsetminus\mathcal{A}_{F},\enskip\sigma(G\sqcup\mathcal{A}_{F}\sqcup E)=\{v_{j}\}^{c}).

Then we have surjective graded k[F]k[F]-module homomorphisms

I|F/J|FM/N[|𝒜F|]M|F𝒜F/N|F𝒜F[|𝒜F|],I|_{F}/J|_{F}\rightarrow M/N[-|\mathcal{A}_{F}|]\rightarrow M|_{F\smallsetminus\mathcal{A}_{F}}/N|_{F\smallsetminus\mathcal{A}_{F}}[-|\mathcal{A}_{F}|],

where the first map is the isomorphism taking x𝒜FxHxHx^{\mathcal{A}_{F}}x^{H}\mapsto x^{H} and the second map is restriction. Finally the right hand term can be identified with L(Γ,𝒜FE)|F𝒜F[|𝒜F|]L(\Gamma,\mathcal{A}_{F}\sqcup E)|_{F\smallsetminus\mathcal{A}_{F}}[-|\mathcal{A}_{F}|]. ∎

We will derive Theorem 1.8 from the following more technical statement.

Theorem 5.2.

Let σ:Γ2V\sigma\colon\Gamma\to 2^{V} be a quasi-geometric homology triangulation, and let EE be a face. Let FlkΓ(E)F\in\operatorname{lk}_{\Gamma}(E) be a face with FEF\sqcup E interior. Suppose 𝒜F=\mathcal{A}_{F}=\emptyset and FF admits an interior partition F=F1F2F=F_{1}\sqcup F_{2}. Assume that FF has no faces GG with GEG\sqcup E interior and |G|<|F1||G|<|F_{1}|. If |F1|2|F_{1}|\leq 2, then L(Γ,E)|FL(\Gamma,E)|_{F} is non-zero in degree |F1||F_{1}|.

Example 5.3.

The conclusion of Theorem 5.2 can fail when |F1|3|F_{1}|\geq 3, even for 𝒜F=E=\mathcal{A}_{F}=E=\emptyset. Consider a geometric triangulation σ:Γ2V\sigma\colon\Gamma\to 2^{V}, where V={v1,,v6}V=\{v_{1},\ldots,v_{6}\} with a face F={w1,,w6}F=\{w_{1},\ldots,w_{6}\} such that

σ(w1)={v1,v3,v6}σ(w2)={v1,v4,v5}σ(w3)={v2,v3,v5}σ(w4)={v2,v4,v6}σ(w5)={v3,v4,v5}σ(w6)={v3,v5,v6}\begin{array}[]{lll}\sigma(w_{1})=\{v_{1},v_{3},v_{6}\}&\sigma(w_{2})=\{v_{1},v_{4},v_{5}\}&\sigma(w_{3})=\{v_{2},v_{3},v_{5}\}\\ \sigma(w_{4})=\{v_{2},v_{4},v_{6}\}&\sigma(w_{5})=\{v_{3},v_{4},v_{5}\}&\sigma(w_{6})=\{v_{3},v_{5},v_{6}\}\end{array}

Then 𝒜F=\mathcal{A}_{F}=\emptyset, and FF admits an interior partition given by F1={w1,w4,w5}F_{1}=\{w_{1},w_{4},w_{5}\}, F2={w2,w3,w6}F_{2}=\{w_{2},w_{3},w_{6}\}. Then (9) gives generators and relations for L(Γ,)|FL(\Gamma,\emptyset)|_{F}, and a linear algebra computation shows that L(Γ,)|F=0L(\Gamma,\emptyset)|_{F}=0.

Before proceeding with the proof of Theorem 5.2, we show how Theorem 1.8 follows from it.

Proof of Theorem 1.8.

We may assume that F=F1F2𝒜FF=F_{1}^{\prime}\sqcup F_{2}^{\prime}\sqcup\mathcal{A}_{F} is an interior partition of FF with |F1||F_{1}^{\prime}| minimal among all possible interior partitions of FF. In particular, if |F1|=2|F_{1}^{\prime}|=2, then there is no vertex vF𝒜Fv\in F\smallsetminus\mathcal{A}_{F} such that {v}𝒜FE\{v\}\sqcup\mathcal{A}_{F}\sqcup E is interior, as then {v}(F1F2{v})𝒜F\{v\}\sqcup(F_{1}^{\prime}\sqcup F_{2}^{\prime}\smallsetminus\{v\})\sqcup\mathcal{A}_{F} would be an interior partition. Hence there are no faces GG of F𝒜FF\smallsetminus\mathcal{A}_{F} with G𝒜FEG\sqcup\mathcal{A}_{F}\sqcup E interior and with cardinality smaller than |F1||F_{1}^{\prime}|. By Theorem 5.2, L(Γ,𝒜FE)|F1F2L(\Gamma,\mathcal{A}_{F}\sqcup E)|_{F_{1}^{\prime}\sqcup F_{2}^{\prime}} is non-zero in degree |F1||F_{1}^{\prime}|. Then, by Lemma 5.1, L(Γ,E)L(\Gamma,E) is nonzero in degree |F1|+|𝒜F||F_{1}^{\prime}|+|\mathcal{A}_{F}|. ∎

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 E=E=\emptyset, we consider the internal edge graph of a subcomplex ΔlkΓ(E)\Delta\subset\operatorname{lk}_{\Gamma}(E). This is the graph contained in the 11-skeleton of lkΓ(E)\operatorname{lk}_{\Gamma}(E) consisting of edges eΔe\subset\Delta with eEe\sqcup E interior.

Lemma 5.4.

Assume σ(E)\sigma(E) has codimension at least 22. Let Δ\Delta be a subcomplex of lkΓ(E)\operatorname{lk}_{\Gamma}(E), and assume Δ\Delta has no vertices vv with {v}E\{v\}\sqcup E interior. If L(Γ,E)|ΔL(\Gamma,E)|_{\Delta} is zero in degree 22, then each connected component of the internal edge graph of Δ\Delta satisfies one of the following.

  1. (1)

    The component is a tree, and it has at most one vertex vv with {v}E\{v\}\sqcup E having carrier codimension more than 11.

  2. (2)

    The component has a unique cycle, and the carrier codimension of {w}E\{w\}\sqcup E is equal to 11 for every vertex ww in the component.

Proof.

From (9), we have the following exact sequence for the degree 22 part of L(Γ,E)|ΔL(\Gamma,E)|_{\Delta}.

|S|=1(IS)1k[lkΓ(E)]k[Δ]I2k[lkΓ(E)]k[Δ](L(Γ,E)|Δ)20.\bigoplus_{|S|=1}(I_{S})_{1}\otimes_{k[\operatorname{lk}_{\Gamma}(E)]}k[\Delta]\to I_{2}\otimes_{k[\operatorname{lk}_{\Gamma}(E)]}k[\Delta]\to(L(\Gamma,E)|_{\Delta})_{2}\to 0.

Because (L(Γ,E)|Δ)2=0(L(\Gamma,E)|_{\Delta})_{2}=0, the first map in the above complex is surjective. As Δ\Delta has no vertices vv with {v}E\{v\}\sqcup E interior, we see that

(12) (xe:eΔ,eE is interior )2=(x{v}θi:vΔ,σ({v}E)={vi}c)2.(x^{e}:e\subset\Delta,\enskip e\sqcup E\text{ is interior })_{2}=(x^{\{v\}}\theta_{i}:v\subset\Delta,\enskip\sigma(\{v\}\sqcup E)=\{v_{i}\}^{c})_{2}.

Thus the number of edges ee with eEe\sqcup E interior is less than or equal to the number of vertices ww with the carrier codimension of {w}E\{w\}\sqcup E equal to 11. If σ({v}E)={vi}c\sigma(\{v\}\sqcup E)=\{v_{i}\}^{c} and θi=wjai,jx{wj}\theta_{i}=\sum_{w_{j}}a_{i,j}x^{\{w_{j}\}}, then

x{v}θi={v,wj}E interior ai,jx{v,wj}.x^{\{v\}}\theta_{i}=\sum_{\{v,w_{j}\}\sqcup E\text{ interior }}a_{i,j}x^{\{v,w_{j}\}}.

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 ee with eEe\sqcup E interior is less than or equal to the number of vertices vv with {v}E\{v\}\sqcup E of carrier codimension 11. As the only connected graphs (V,E)(V,E) where |E||V||E|\leq|V| are either trees or contain a unique cycle, the result follows. ∎

Lemma 5.5.

Assume σ(E)\sigma(E) has codimension at least 22. Let FlkΓ(E)F\subset\operatorname{lk}_{\Gamma}(E) be a face. Assume FF has no vertices vv with {v}E\{v\}\sqcup E interior. If L(Γ,E)|FL(\Gamma,E)|_{F} is zero in degree 22, then no component of the internal edge graph of FF contains a cycle of length 44.

Proof.

Suppose a component of the internal edge graph contains a 44-cycle of vertices F={t1,t2,u1,u2}F=\{t_{1},t_{2},u_{1},u_{2}\}. By Lemma 5.4, this is the unique cycle in this component and every vertex wFw\in F has {w}E\{w\}\sqcup E of carrier codimension 11. Because FF is a face and there are no 33-cycles in this component of the internal edge graph, we may assume that σ({ti}E)={v1}c\sigma(\{t_{i}\}\sqcup E)=\{v_{1}\}^{c} and σ({ui}E)={v2}c\sigma(\{u_{i}\}\sqcup E)=\{v_{2}\}^{c}. Restricting to FF and using that (L(Γ,E)|F)2=0(L(\Gamma,E)|_{F})_{2}=0, we have that

(x{t1,u1},x{u1,t2},x{t2,u2},x{u2,t1})=(x{t1}θ2,x{t2}θ2,x{u1}θ1,x{u2}θ1).(x^{\{t_{1},u_{1}\}},x^{\{u_{1},t_{2}\}},x^{\{t_{2},u_{2}\}},x^{\{u_{2},t_{1}\}})=(x^{\{t_{1}\}}\theta_{2},x^{\{t_{2}\}}\theta_{2},x^{\{u_{1}\}}\theta_{1},x^{\{u_{2}\}}\theta_{1}).

The relation θ1θ2θ2θ1=0\theta_{1}\theta_{2}-\theta_{2}\theta_{1}=0 expands into a relation between the generators of the right-hand side. But the left-hand side is 44-dimensional, a contradiction. ∎

Lemma 5.6.

Assume σ(E)\sigma(E) has codimension 11. Let ΔlkΓ(E)\Delta\subset\operatorname{lk}_{\Gamma}(E) be a subcomplex. Then

dim(L(Γ,E)|Δ)1|{vΔ:{v}E interior}|1.\dim(L(\Gamma,E)|_{\Delta})_{1}\geq|\{v\in\Delta:\{v\}\sqcup E\text{ interior}\}|-1.
Proof.

By considering the degree 11 part of (9), as the codimension of σ(E)\sigma(E) is 11, we get the following exact sequence.

k{k}wΔ{w}E interior kxw{\bigoplus\limits_{\mathclap{\begin{subarray}{c}w\in\Delta\\ \{w\}\sqcup E\text{ interior }\end{subarray}}}k\cdot x^{w}}(L(Γ,E)|Δ)1{(L(\Gamma,E)|_{\Delta})_{1}}0,{0,}

and the result follows. ∎

Proof of Theorem 5.2.

We must show that L(Γ,E)|FL(\Gamma,E)|_{F} is non-zero in degree |F1||F_{1}|. Recall that L(Γ,E)|FL(\Gamma,E)|_{F} is isomorphic to I|F/J|FI|_{F}/J|_{F}, where I|FI|_{F} and J|FJ|_{F} are described in (10) and (11) respectively. First we handle the cases when |F1|1|F_{1}|\leq 1. If F1=F_{1}=\emptyset, then EE is interior and x=1x^{\emptyset}=1, but J|FJ|_{F} is a proper ideal as it is generated by elements of positive degree, so xF1J|Fx^{F_{1}}\not\in J|_{F}. If F1={v}F_{1}=\{v\}, then we assume that EE is not an interior face. Then J|FJ|_{F} is generated by elements of degree at least 22, so xF1J|Fx^{F_{1}}\not\in J|_{F}.

Suppose |F1|=2|F_{1}|=2. We assume that there are no vertices vv with {v}E\{v\}\sqcup E interior and EE is not interior. If σ(E)\sigma(E) has codimension 11, then both F1F_{1} and F2F_{2} must have a vertex vv with {v}E\{v\}\sqcup E interior. Then by Lemma 5.6, we see that dimL(Γ,E)|F1\dim L(\Gamma,E)|_{F}\geq 1. Hence we may assume that σ(E)\sigma(E) has codimension at least 22.

Let F1={u,t}F_{1}=\{u,t\} and assume that L(Γ,E)|FL(\Gamma,E)|_{F} has no non-zero elements in degree 22. Consider the connected component of the internal edge graph containing F1F_{1}. By Lemma 5.4, we may assume that σ({u}E)={v1}c\sigma(\{u\}\sqcup E)=\{v_{1}\}^{c}. Note that v1σ(t)v_{1}\in\sigma(t). There is a vertex tF2t^{\prime}\in F_{2} such that v1σ(t)v_{1}\in\sigma(t^{\prime}), so {u,t}E\{u,t^{\prime}\}\sqcup E is interior. Therefore either {t}E\{t\}\sqcup E or {t}E\{t^{\prime}\}\sqcup E has carrier codimension 11.

If σ({t}E)={v2}c\sigma(\{t\}\sqcup E)=\{v_{2}\}^{c}, then there is a vertex uF2u^{\prime}\in F_{2} such that v2σ(u)v_{2}\in\sigma(u^{\prime}). First assume uu^{\prime} and tt^{\prime} are distinct. Since at least one of {u}E\{u^{\prime}\}\sqcup E and {t}E\{t^{\prime}\}\sqcup E has carrier codimension 11, it follows that {u,t}E\{u^{\prime},t^{\prime}\}\sqcup E is interior. Then {u,t,u,t}\{u,t,u^{\prime},t^{\prime}\} forms a 44-cycle, contradicting Lemma 5.5.

If u=tu^{\prime}=t^{\prime}, then the internal edge graph contains a cycle and hence every vertex ww in it (including tt) has {w}E\{w\}\sqcup E of carrier codimension 11. As F2F_{2} is interior and {u}E\{u^{\prime}\}\sqcup E has carrier codimension 11, there is a vertex wF2w\in F_{2} such that {u,w}E\{u^{\prime},w\}\sqcup E is interior. But then either {u,w}E\{u,w\}\sqcup E or {t,w}E\{t,w\}\sqcup E is interior, contradicting the uniqueness of the cycle in Lemma 5.4.

If {t}E\{t\}\sqcup E does not have carrier codimension 11, then we may assume that σ({t}E)={v2}c\sigma(\{t^{\prime}\}\sqcup E)=\{v_{2}\}^{c}. Choose a vertex uF2u^{\prime}\in F_{2} with v2σ(u)v_{2}\in\sigma(u^{\prime}). Then {t,u}E\{t^{\prime},u^{\prime}\}\sqcup E is interior, so {u}E\{u^{\prime}\}\sqcup E has carrier codimension 11. If v1σ(u)v_{1}\in\sigma(u^{\prime}), then {u,u}E\{u,u^{\prime}\}\sqcup E is interior. If v1σ(u)v_{1}\not\in\sigma(u^{\prime}), then {t,u}E\{t,u^{\prime}\}\sqcup E is interior. In either case, there is a cycle and a vertex vv with {v}E\{v\}\sqcup E of carrier codimension more than 11 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 hh-vectors. For instance, suppose V={v1,,v6}V=\{v_{1},\ldots,v_{6}\} and σ:Γ2V\sigma\colon\Gamma\to 2^{V} is a geometric triangulation with a facet F={w1,,w6}F=\{w_{1},\ldots,w_{6}\} such that

σ(w1)={v1}σ(w2)={v2}σ(w3)={v3}σ(w4)={v1,v4,v5}σ(w5)={v2,v4,v6}σ(w6)={v3,v5,v6}\begin{array}[]{lll}\sigma(w_{1})=\{v_{1}\}&\sigma(w_{2})=\{v_{2}\}&\sigma(w_{3})=\{v_{3}\}\\ \sigma(w_{4})=\{v_{1},v_{4},v_{5}\}&\sigma(w_{5})=\{v_{2},v_{4},v_{6}\}&\sigma(w_{6})=\{v_{3},v_{5},v_{6}\}\end{array}

Then the interior 22-faces of FF are {w1,w5,w6}\{w_{1},w_{5},w_{6}\}, {w2,w4,w6}\{w_{2},w_{4},w_{6}\}, {w3,w4,w5}\{w_{3},w_{4},w_{5}\}, and {w4,w5,w6}\{w_{4},w_{5},w_{6}\}. But FF has no interior vertices or edges, and it has only three edges with carrier codimension one, namely {w4,w5},{w4,w6},\{w_{4},w_{5}\},\{w_{4},w_{6}\}, and {w5,w6}\{w_{5},w_{6}\}. Thus L(Γ,)|FL(\Gamma,\emptyset)|_{F} is non-zero in degree three. Note that FF is not a pyramid and does not admit an interior partition.

References

  • [Ath12a] C. Athanasiadis, Cubical subdivisions and local hh-vectors, Ann. Comb. 16 (2012), no. 3, 421–448.
  • [Ath12b] by same author, Flag subdivisions and γ\gamma-vectors, Pacific J. Math. 259 (2012), no. 2, 257–278.
  • [Ath16] by same author, A survey of subdivisions and local hh-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 hh-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 hh-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 hh-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 hh-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.