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

A Generalization of the Ishida Complex with applications

Laura Felicia Matusevich Department of Mathematics
Texas A&M University
College Station, TX 77843.
matusevich@tamu.edu
Erika Ordog Department of Mathematics
Texas A&M University
College Station, TX 77843.
erika.ordog@tamu.edu
 and  Byeongsu Yu Department of Mathematics
Texas A&M University
College Station, TX 77843.
byeongsu.yu@tamu.edu
Abstract.

We construct a generalize Ishida complex to compute the local cohomology with monomial support of modules over quotients of polynomial rings by cellular binomial ideals. As a consequence, we obtain a combinatorial criterion to determine when such a quotient is Cohen–Macaulay. In particular, this gives a Cohen–Macaulayness criterion for lattice ideals. We also prove a result relating the local cohomology with radical monomial ideal support of an affine semigroup ring to the local cohomology with maximal ideal support of the quotient of the affine semigroup ring by the radical monomial ideal. This requires a combinatorial assumption on the semigroup, which holds for (not necessarily normal) semigroups whose cone is the cone over a simplex.

2020 Mathematics Subject Classification:
Primary 13F65, 13D45; Secondary 13F55.

1. Introduction

Stanley–Reisner rings and affine semigroup rings are a staple of combinatorial commutative algebra. The former are defined by squarefree monomial ideals, while the latter are defined by toric ideals. Both of these special classes of ideals fall under the general umbrella of binomial ideals, that is, ideals defined by polynomials with at most two terms.

Local cohomology is an important tool from homological commutative algebra, which has proved very useful in combinatorial settings. In this article, we study local cohomology of more general binomial ideals, known as cellular binomial ideals. To do this, we generalize and adapt the Ishida complex, which was developed to compute local cohomology for modules over affine semigroup rings with support at the maximal monomial ideal, in two ways: by extending the base ring, and also the supporting monomial ideal.

Extending the base ring is necessary, as quotients by cellular binomial ideals are not modules over affine semigroup rings. The proof that our generalized Ishida complex indeed computes local cohomology follows along the usual lines of checking that it works at homological degree zero, and then proving vanishing for injectives. The combinatorics involved becomes more challenging in the more general context.

Our original motivation for these developments was was twofold. First, we wanted to have a combinatorial criterion for when a lattice ideal is Cohen–Macaulay. Second, we wanted to better understand a duality result for local cohomology in the Stainley–Reisner case.

1.1. Lattice ideals

A lattice is a subgroup of n\mathbb{Z}^{n}. Given a lattice LL, its lattice ideal is

IL:=xuxvuvL𝕜[x1,,xn],I_{L}:=\langle x^{u}-x^{v}\mid u-v\in L\rangle\subset\Bbbk[x_{1},\dots,x_{n}],

where 𝕜\Bbbk is a field. The saturation of a lattice LL is Lsat:=(L)nL_{\text{sat}}:=(\mathbb{Q}\oplus_{\mathbb{Z}}L)\cap\mathbb{Z}^{n}. LL is saturated if it equals its saturation. Lattice ideals corresponding to saturated lattices are known as toric ideals, and are easily seen to be prime. The toric ideal ILsatI_{L_{\text{sat}}} is known to be a minimal prime of ILI_{L}, and if 𝕜\Bbbk is algebraically closed, all associated primes of ILI_{L} are isomorphic to ILsatI_{L_{\text{sat}}} by rescaling the variables [MR1394747].

Quotients by toric ideals are affine semigroup rings. Cohen–Macaulayness of affine semigroup rings is well studied, see for instance [MR304376, MR857437, MY2022]. In the case of general lattice ideals, one can compute Betti numbers using suitable simplicial complexes. In special cases [MR1649322, MR1475887], these simplicial complexes have tractable enough homology to provide combinatorial criteria to determine whether a quotient by a lattice ideal is Cohen–Macaulay. Moreover, there is no clear relationship between Cohen–Macaulayness of ILI_{L} and ILsatI_{L_{\text{sat}}} [MR3957112].

Local cohomology has been very effective to provide such combinatorial criteria in other situations, but was not previously studied for lattice ideals. This is mainly for two reasons: the lack of specialized tools (which are available for toric ideals but not in general), and the fact that the natural grading group for ILI_{L} is n/L\mathbb{Z}^{n}/L which has torsion.

In this case we tackle both of these issues. First, we generalize the Ishida complex, which computes local cohomology for modules over affine semigroup rings to the more general context of lattice ideals. (Actually, we can deal with more general binomial ideals known as cellular binomial ideals.) Then we study torsion gradings. The end result is the desired combinatorial Cohen–Macaulayness criterion in the cellular binomial case.

1.2. Duality for local cohomology

If Δ\Delta is a simplicial complex, IΔI_{\Delta} is the associated squarefree monomial ideal, and 𝕜[Δ]=𝕜[x1,,n]/IΔ\Bbbk[\Delta]=\Bbbk[x_{1},\dots,n]/I_{\Delta} the corresponding Stanley–Reisner ring, a comparison of Hilbert series formulas for local cohomology due to Hochster (for maximal ideal support) and Terai (for radical monomial ideal support) [Terai99] yields the following result (see also [Huneke07]*Theorem 6.8 and the references therein)

HIΔJ(𝕜[x])=0H𝔪nj(𝕜[Δ])=0.H_{I_{\Delta}}^{J}(\Bbbk[x])=0\Longleftrightarrow H^{n-j}_{\mathfrak{m}}(\Bbbk[\Delta])=0. (1)

We asked the question whether there was a relationship in the nonvanishing case, and whether this would hold in the more general context of radical monomial ideals in affine semigroup rings.

This brought us to our second direction of generalization for the Ishida complex, to deal with monomial supports other than the homogeneous maximal ideal.

In the end, we can prove an isomorphism of local cohomology generalizing (1) (Theorem 6.3), but it requires a combinatorial condition on the affine semigroup ring. Nevertheless, our result does hold for affine semigroup rings whose corresponding cone is the cone over a simplex. We remark that we do not require a normality assumption.

Outline

In section 2, we recall known results on binomial ideals. In Section 3, we introduce the generalized Ishida complex, which allows us to compute the local cohomology of a module over a polynomial ring quotient by a cellular binomial ideal with radical monomial ideal support. In section 4, we study local cohomology of cellular binomial ideals, and provide a Cohen–Macaulayness criterion. In Section 5, we focus on the local cohomology of a quotient of a simplicial affine semigroup ring by a radical monomial ideal. In section 6, we present our duality result for local cohomology.

Acknowledgments

We are grateful to Aida Maraj, Aleksandra Sobieska, Catherine Yan, Jaeho Shin, Jennifer Kenkel, Jonathan Montaño, Joseph Gubeladze, Kenny Easwaran, Melvin Hochster, Sarah Witherspoon, Semin Yoo, Serkan Hoşten, Yupeng Li for inspiring conversations we had while working on this project. EO was supported by an NSF Mathematical Sciences Postdoctoral Research Fellowship under award DMS-2103253.

2. Preliminaries

2.1. Affine semigroup rings

An affine semigroup QQ is a finitely generated submonoid of d\mathbb{Z}^{d}. Throughout this article, we denote by AA a d×nd\times n integer matrix of rank dd whose columns generate QQ, so that Q=AQ=\mathbb{N}A. We assume none of the columns of AA is the zero vector. If 𝕜\Bbbk is a field and QQ is the affine semigroup given by AA, we denote by 𝕜[Q]=𝕜[A]\Bbbk[Q]=\Bbbk[\mathbb{N}A] the corresponding affine semigroup ring. Since AA has rank dd, 𝕜[Q]\Bbbk[Q] is a dd-dimensional 𝕜\Bbbk-algebra.

The cone over the affine semigroup QQ (or over AA) is the (rational, polyhedral) cone 0Q(=0A)\mathbb{R}_{\geq 0}Q(=\mathbb{R}_{\geq 0}A), that is, the set of all non-negative real combinations of columns of AA. This cone is pointed if it contains no lines. In this case, the affine semigroup QQ is also called pointed. A face of 0A\mathbb{R}_{\geq 0}A is a subset of this cone where some linear functional on d\mathbb{R}^{d} is maximized over 0A\mathbb{R}_{\geq 0}A. We denote by (0Q)\mathcal{F}(\mathbb{R}_{\geq 0}Q) the collection of all faces of 0Q\mathbb{R}_{\geq 0}Q. This set forms a lattice under inclusion. For ease in the notation, we identify a face of QQ with the set of columns of AA that lie on that face. For a face F(Q)F\in\mathcal{F}(Q), the relative interior RelInt(F)\operatorname{RelInt}(\mathbb{N}F) of the semigroup over FF is the set of elements of F\mathbb{N}F that do not belong to any subsemigroups arising from proper faces of F\mathbb{N}F. A transverse section KK of 0Q\mathbb{R}_{\geq 0}Q is the intersection of 0Q\mathbb{R}_{\geq 0}Q with a hyperplane which meets all unbounded faces of 0Q\mathbb{R}_{\geq 0}Q [Ziegler95]*Exercise 2.19. It is well-known that (K)\mathcal{F}(K) is canonically bijective to (0Q)\mathcal{F}(\mathbb{R}_{\geq 0}Q) as a poset.

Let SpecMon𝕜[Q]\operatorname{Spec}_{\text{Mon}}\Bbbk[Q] be the set of all monomial prime ideals of the affine semigroup ring 𝕜[Q]\Bbbk[Q], ordered by inclusion. There is an order-reversing isomorphism between SpecMon𝕜[Q]\operatorname{Spec}_{\text{Mon}}\Bbbk[Q] and the face lattice (K)\mathcal{F}(K) of the transverse section KK of 0Q\mathbb{R}_{\geq 0}Q given by identifying a face of KK with the (prime) ideal generated by all monomials whose exponents do not belong to the corresponding face of 0Q\mathbb{R}_{\geq 0}Q.

It is known that the set of all radical ideals is in one to one correspondence with the set of all polyhedral subcomplexes of (Q)\mathcal{F}(Q).

2.2. Binomial ideals

Let 𝕜[x]=𝕜[x1,x2,,xd]\Bbbk[x]=\Bbbk[x_{1},x_{2},\dots,x_{d}] be the polynomial ring over a field 𝕜\Bbbk. A binomial is a polynomial having at most two terms. A binomial ideal is an ideal generated by binomials. We are interested in three types of binomial ideals, lattice ideals, toric ideals, and cellular binomial ideals.

Let LρL_{\rho} be a subgroup of d\mathbb{Z}^{d}. A partial character ρ:Lρ𝕜\rho:L_{\rho}\to\Bbbk^{\ast} on d\mathbb{Z}^{d} is a group homomorphism, where 𝕜\Bbbk^{\ast} is the multiplicative group of 𝕜\Bbbk. The lattice ideal I(ρ)I(\rho) corresponding to ρ\rho is the ideal in 𝕜[x]\Bbbk[x] defined as

I(ρ):=x𝐮ρ(𝐮𝐯)xv𝐮𝐯Lρ𝕜[x1,,xn].I(\rho):=\langle x^{\mathbf{u}}-\rho(\mathbf{u}-\mathbf{v})x^{v}\mid\mathbf{u}-\mathbf{v}\in L_{\rho}\rangle\subset\Bbbk[x_{1},\dots,x_{n}]. (2)

We remark that in [MR1394747], these lattice ideals are denoted by I(ρ)+I(\rho)_{+}, while I(ρ)I(\rho) is used for lattice ideals in Laurent polynomial rings. Since we do not need the more general context, we use (2) for economy in the notation.

The saturation (Lρ)sat(L_{\rho})_{\text{sat}} of a lattice LρL_{\rho} is (Lρ)sat:=(Lρ)d(L_{\rho})_{\text{sat}}:=(\mathbb{Q}\otimes_{\mathbb{Z}}L_{\rho})\cap\mathbb{Z}^{d}. A lattice is saturated if Lρ=(Lρ)satL_{\rho}=(L_{\rho})_{\text{sat}}. If 𝕜\Bbbk is algebraically closed, then I(ρ)I(\rho) is prime if and only if LρL_{\rho} is saturated [MR1394747]*Theorem 2.1.c.. Furthermore, the associated primes of a lattice ideal are lattice ideals corresponding to the saturation of the underlying lattice [MR1394747]*Corollary 2.5. In particular, [MR1394747]*Corollary 2.5 implies that if 𝕜\Bbbk is an algebraically closed field of characteristic zero, then lattice ideals are radical, and primary lattice ideals are prime.

An ideal I𝕜[x]I\subset\Bbbk[x] is cellular if all variables are either nonzero divisors modulo II or nilpotent modulo II. The nonzero divisor variables are known as the cellular variables of II. Let ζ[n]\zeta\subset[n] be the set of all indices of cellular variables of a cellular binomial ideal II, then II is ζ\zeta-cellular. Lattice ideal are special cases of cellular binomial ideals, where all variables are cellular. Also, I𝕜[ζ]I\cap\Bbbk[\mathbb{N}^{\zeta}] is a lattice ideal [MR3556446]. The following result can be found in [MR1394747]*Section 6 and also in [MR3556446]*Corollary 3.5.

Theorem 2.1.

Let II be a ζ\zeta-cellular binomial ideal in 𝕜[x]\Bbbk[x]. The associated primes of II are the ideals 𝕜[x]P+xiiζc\Bbbk[x]\cdot P+\langle x_{i}\mid i\in\zeta^{c}\rangle, where P𝕜[ζ]P\subset\Bbbk\left[\mathbb{N}^{\zeta}\right] runs over the associated primes of lattice ideals of the form (I:m)𝕜[ζ](I:m)\cap\Bbbk\left[\mathbb{N}^{\zeta}\right], for monomials m𝕜[ζc]m\in\Bbbk\left[\mathbb{N}^{\zeta^{c}}\right].

A toric ideal is the prime lattice ideal I(χ)I(\chi) corresponding to the trivial character χ:Lχ{1}𝕜\chi:L_{\chi}\to\{1\}\subset\Bbbk^{\ast}. An affine semigroup QQ is the quotient of d\mathbb{N}^{d} by the equivalence relation Lχ\sim_{L_{\chi}} given by 𝐮Lχ𝐯𝐮𝐯Lχ\mathbf{u}\sim_{L_{\chi}}\mathbf{v}\iff\mathbf{u}-\mathbf{v}\in L_{\chi}. A quotient of polynomial ring by corresponding prime lattice ideal 𝕜[x]/I(χ)\Bbbk[x]/I(\chi) is isomorphic to an affine semigroup ring 𝕜[Q]\Bbbk[Q] [CCA]*Theorem 7.3.

From now on, we assume that the base field 𝕜\Bbbk is algebraically closed.

3. Generalized Ishida complex

Given a lattice ideal II (resp. ζ\zeta-cellular binomial ideal II), pick a minimal associated prime ideal JJ of II (resp. of I𝕜[ζ]I\cap\Bbbk[\mathbb{N}^{\zeta}]). Since 𝕜\Bbbk is algebraically closed, JJ is a binomial prime ideal, and the quotient 𝕜[ζ]/J\Bbbk[\mathbb{N}^{\zeta}]/J is isomorphic (by rescaling the variables) to an affine semigroup ring 𝕜[Q]\Bbbk[Q] with Q=AQ=\mathbb{N}A. The natural projection map 𝕜[x]/I𝕜[x]/J𝕜[Q]\Bbbk[x]/I\to\Bbbk[x]/J\cong\Bbbk[Q] (resp. 𝕜[x]/I𝕜[x]/(J+xi:iζc)𝕜[Q]\Bbbk[x]/I\to\Bbbk[x]/(J+\langle x_{i}:i\in\zeta^{c}\rangle)\cong\Bbbk[Q]) induces AA-grading on 𝕜[x]/I\Bbbk[x]/I; in other words, for any monomial x𝐮¯𝕜[x]/I\overline{x^{\mathbf{u}}}\in\Bbbk[x]/I, the AA-degree of x𝐮¯\overline{x^{\mathbf{u}}} is degA(x𝐮¯)=A𝐮\deg_{A}(\overline{x^{\mathbf{u}}})=A\cdot\mathbf{u} (resp. A𝐮ζA\cdot\mathbf{u}^{\zeta}, where 𝐮ζ:=(ui𝐮;iζ)\mathbf{u}^{\zeta}:=(u_{i}\in\mathbf{u};i\in\zeta)). We specify the AA-grading on 𝕜[x]/I\Bbbk[x]/I using the triple (I,J,A)(I,J,A) unless the minimal prime JJ or the generators of affine semigroup A\mathbb{N}A are understood in context.

In this section, we generalize the Ishida complex to compute the local cohomology of a lattice ideal (resp. ζ\zeta-cellular binomial ideal) (I,J,A)(I,J,A) supported at a contraction of a radical monomial ideal of 𝕜[A]=𝕜[Q]\Bbbk[\mathbb{N}A]=\Bbbk[Q]. Recall that the contraction of a radical monomial ideal is also a monomial ideal in 𝕜[x]/I\Bbbk[x]/I.

3.1. Ishida complex

The Ishida complex was originally developed for computing the local cohomology of modules over pointed affine semigroup rings supported on the graded maximal ideal [MR977758]. Given a pointed affine semigroup QQ with transverse section KK, there is a canonical isomorphism ^:(K)(Q)\widehat{-}:\mathcal{F}(K)\to\mathcal{F}(Q) given as follows: F^\widehat{F} is the minimal face of QQ such that 0F^0F\mathbb{R}_{\geq 0}\widehat{F}\supseteq\mathbb{R}_{\geq 0}F. Since QQ is pointed, it has a unique zero-dimensional face, namely the origin, which corresponds to the (-1)-dimensional face \varnothing of KK. As a CW complex, KK has an incidence function ϵ:i=1d1(K)i×(K)i+1{0,±1}\epsilon:\bigoplus_{i=-1}^{d-1}\mathcal{F}(K)^{i}\times\mathcal{F}(K)^{i+1}\to\{0,\pm 1\} that has a nonzero value when two faces are incident. We now recall the definition of the Ishida complex [MR977758].

Definition 3.1.

Let 𝔪\mathfrak{m} be the maximal monomial ideal of 𝕜[Q]\Bbbk[Q]. The set of all kk-dimensional faces in (K)\mathcal{F}(K) is denoted by (K)k\mathcal{F}(K)^{k}. Let LL^{\bullet} be the chain complex

L:0{L^{\bullet}:0}L0{L^{0}}L1{L^{1}}{\cdots}Ld{L^{d}}0,Lk:=F(K)k1𝕜[QF^]{0,\qquad L^{k}:=\bigoplus\limits_{F\in\mathcal{F}(K)^{k-1}}\Bbbk[Q-\mathbb{N}\widehat{F}]}\scriptstyle{\partial}\scriptstyle{\partial}\scriptstyle{\partial}\scriptstyle{\partial}

where the differential :LkLk+1\partial:L^{k}\to L^{k+1} is induced by the componentwise map F,G\partial_{F,G} with F(K)k1F\in\mathcal{F}(K)^{k-1}, G(K)kG\in\mathcal{F}(K)^{k} such that

F,G:𝕜[QF^]𝕜[QG^] to be {0 if FGϵ(F,G)nat if FG\partial_{F,G}:\Bbbk[Q-\mathbb{N}\widehat{F}]\to\Bbbk[Q-\mathbb{N}\widehat{G}]\text{ to be }\begin{cases}0&\text{ if }F\not\subset G\\ \epsilon(F,G)\cdot\operatorname{nat}&\text{ if }F\subset G\end{cases}

with nat\operatorname{nat}, the canonical injection 𝕜[QF^]𝕜[QG^]\Bbbk[Q-\mathbb{N}\widehat{F}]\to\Bbbk[Q-\mathbb{N}\widehat{G}] when FGF\subseteq G. We say that L𝕜[Q]ML^{\bullet}\otimes_{\Bbbk[Q]}M is the Ishida complex of a 𝕜[Q]\Bbbk[Q]-module MM supported at the maximal monomial ideal.

The cohomology of the Ishida complex of MM supported at the maximal monomial ideal is isomorphic to the local cohomology of MM supported at the maximal monomial ideal.

Theorem 3.2 ([MR977758]*Theorem 6.2.5).

For any 𝕜[Q]\Bbbk[Q]-module MM, and all k0k\geq 0,

H𝔪k(M)Hk(L𝕜[Q]M).H_{\mathfrak{m}}^{k}(M)\cong H^{k}(L^{\bullet}\mathbin{\mathop{\otimes}\displaylimits_{\Bbbk[Q]}}M).

Let JΔJ_{\Delta} be the radical monomial ideal of 𝕜[Q]\Bbbk[Q] associated to a subcomplex Δ(Q)\Delta\subset\mathcal{F}(Q). Then, JΔ𝕜[x]/I\sqrt{J_{\Delta}\cdot\Bbbk[x]/I} denotes the contraction of JΔJ_{\Delta} via 𝕜[x]/I𝕜[x]/J𝕜[Q]\Bbbk[x]/I\to\Bbbk[x]/J\cong\Bbbk[Q]. To compute the local cohomology of a 𝕜[x]/I\Bbbk[x]/I-module supported on JΔ𝕜[x]/I\sqrt{J_{\Delta}\cdot\Bbbk[x]/I}, we construct a (generalized) Ishida complex below.

Let KJΔK_{J_{\Delta}} be a transverse section of the polyhedron 0{uζxuJΔ}\mathbb{R}_{\geq 0}\{u\in\mathbb{Z}^{\zeta}\mid x^{u}\in J_{\Delta}\} with the canonical isomorphism ^:(KJΔ)(Q)\widehat{-}:\mathcal{F}(K_{J_{\Delta}})\to\mathcal{F}(Q) where F^\widehat{F} is the minimal face of QQ such that 0F^0F\mathbb{R}_{\geq 0}\widehat{F}\supseteq\mathbb{R}_{\geq 0}F. The set of all kk-dimensional faces in (KJΔ)\mathcal{F}(K_{J_{\Delta}}) is denoted by (KJΔ)k\mathcal{F}(K_{J_{\Delta}})^{k}. Also, (𝕜[x]/I)F^(\Bbbk[x]/I)_{\widehat{F}} refers to the localization of 𝕜[x]/I\Bbbk[x]/I by the multiplicative set consisting of all monomials in 𝕜[ζ]\Bbbk[\mathbb{N}^{\zeta}] whose AA-graded degrees are in F^\mathbb{N}\widehat{F}.

Definition 3.3 (Generalized Ishida complex).

Let LL^{\bullet} be the chain complex

L:0{L^{\bullet}:0}L0{L^{0}}L1{L^{1}}{\cdots}Ld{L^{d}}0,Lk:=F(KJΔ)k1(𝕜[x]/I)F^{0,\quad L^{k}:=\bigoplus\limits_{F\in\mathcal{F}(K_{J_{\Delta}})^{k-1}}\left(\Bbbk[x]/I\right)_{\widehat{F}}}\scriptstyle{\partial}\scriptstyle{\partial}\scriptstyle{\partial}\scriptstyle{\partial}

where the differential :LkLk+1\partial:L^{k}\to L^{k+1} is induced by a componentwise map F,G\partial_{F,G} with two faces F(KJΔ)k1F\in\mathcal{F}(K_{J_{\Delta}})^{k-1}, G(KJΔ)kG\in\mathcal{F}(K_{J_{\Delta}})^{k} such that

F,G:(𝕜[x]/I)F^(𝕜[x]/I)G^ to be {0 if FGϵ(F,G)nat if FG\partial_{F,G}:\left(\Bbbk[x]/I\right)_{\widehat{F}}\to\left(\Bbbk[x]/I\right)_{\widehat{G}}\text{ to be }\begin{cases}0&\text{ if }F\not\subset G\\ \epsilon(F,G)\cdot\operatorname{nat}&\text{ if }F\subset G\end{cases}

with nat\operatorname{nat}, the canonical injection (𝕜[x]/I)F^(𝕜[x]/I)G^\left(\Bbbk[x]/I\right)_{\widehat{F}}\to\left(\Bbbk[x]/I\right)_{\widehat{G}} when FGF\subseteq G. We say that L𝕜[x]/IML^{\bullet}\otimes_{\Bbbk[x]/I}M is the Ishida complex of a (𝕜[x]/I)(\Bbbk[x]/I)-module MM supported at the radical monomial ideal JΔ𝕜[x]/I\sqrt{J_{\Delta}\cdot\Bbbk[x]/I}.

The following theorem is the main result in this section. The proof is adapted from [BH_CMrings]. The key ingredients are Lemma 3.5 and Lemma 3.6 which are given later.

Theorem 3.4.

For any 𝕜[x]/I\Bbbk[x]/I-module MM, and all k0k\geq 0,

HJΔ𝕜[x]/Ik(M)HJΔ𝕜[x]/Ik(M)Hk(L𝕜[x]/IM).H_{J_{\Delta}\cdot\Bbbk[x]/I}^{k}(M)\cong H_{\sqrt{J_{\Delta}\cdot\Bbbk[x]/I}}^{k}(M)\cong H^{k}(L^{\bullet}\mathbin{\mathop{\otimes}\displaylimits_{\Bbbk[x]/I}}M).

The first step in our proof is to verify that the zeroth homology of the generalized Ishida complex computes torsion.

Proof.

This follows from Lemma 3.5 and Lemma 3.6 together with the fact that all the summands of the components of the Ishida complex are flat, thus 𝕜[x]/IL-\mathbin{\mathop{\otimes}\displaylimits_{\Bbbk[x]/I}}L^{\bullet} is an exact functor. ∎

Lemma 3.5.

H0(L𝕜[x]/IM)(0:MJΔ)H^{0}(L\otimes_{\Bbbk[x]/I}M)\cong(0:_{M}J_{\Delta}^{\infty}).

Proof.

It suffices to show that

F(KJΔ)0{x𝐮¯𝕜[ζ]/(I𝕜[ζ]):degA(𝐮)RelInt(F^)}=JΔ𝕜[x]/I.\sqrt{\left\langle\bigcup\limits_{F\in\mathcal{F}(K_{J_{\Delta}})^{0}}\{\overline{x^{\mathbf{u}}}\in\Bbbk[\mathbb{N}^{\zeta}]/(I\cap\Bbbk[\mathbb{N}^{\zeta}]):\deg_{A}(\mathbf{u})\in\operatorname{RelInt}(\mathbb{N}\widehat{F})\}\right\rangle}=\sqrt{J_{\Delta}\cdot\Bbbk[x]/I}.

The generators of the left hand-side ideal might be not the same as those of the multiplicative sets inducing localization of components in L1L^{1} when II is not toric. However, they admit the same radical ideal in 𝕜[x]/I\Bbbk[x]/I.

First, let x𝐮¯𝕜[x]/I\overline{x^{\mathbf{u}}}\in\Bbbk[x]/I be an element whose AA-degree is in F^\mathbb{N}\widehat{F} for a vertex FF of KJΔK_{J_{\Delta}}. The canonical map 𝕜[x]/I𝕜[x]/J𝕜[A]\Bbbk[x]/I\to\Bbbk[x]/J\cong\Bbbk[\mathbb{N}A] sends x𝐮¯\overline{x^{\mathbf{u}}} to xdegA(𝐮)𝕜[A]x^{\deg_{A}(\mathbf{u})}\in\Bbbk[\mathbb{N}A] where degA(𝐮)RelInt(F^)\deg_{A}(\mathbf{u})\in\operatorname{RelInt}(\mathbb{N}\widehat{F}). Since F^Δ\widehat{F}\not\in\Delta, xdegA(𝐮)JΔx^{\deg_{A}(\mathbf{u})}\in J_{\Delta}, we have that x𝐮¯JΔ𝕜[x]/I\overline{x^{\mathbf{u}}}\in\sqrt{J_{\Delta}\cdot\Bbbk[x]/I} by the correspondence between polyhedral subcomplexes and the radical monomial ideals.

Conversely, let f𝕜[ζ]f\in\Bbbk[\mathbb{N}^{\zeta}] be a preimage of a monomial in JΔ𝕜[A]J_{\Delta}\subset\Bbbk[\mathbb{N}A] and g𝕜[x]/Ig\in\Bbbk[x]/I be an AA-homogeneous element of 𝕜[x]/I\Bbbk[x]/I. Then, f=x𝐮¯f=\overline{x^{\mathbf{u}}} for some 𝐮ζ\mathbf{u}\in\mathbb{N}^{\zeta} such that A𝐮F¯A\mathbf{u}\in\mathbb{N}\overline{F} for some F(KJΔ)F\in\mathcal{F}(K_{J_{\Delta}}). If dimF=0\dim F=0, fgfg is in the left hand-side of the equation. Suppose dimF>0\dim F>0; then FF has vertices {v1,,vm}\{v_{1},\dots,v_{m}\}. Then, 𝐮\mathbf{u} is a linear combination of elements of vi^\mathbb{N}\widehat{v_{i}} over \mathbb{Q}, say 𝐮=c1𝐮1++cm𝐮m\mathbf{u}=c_{1}\mathbf{u}_{1}+\cdots+c_{m}\mathbf{u}_{m} where 𝐮ivi^\mathbf{u}_{i}\in\mathbb{N}\widehat{v_{i}} and cic_{i}\in\mathbb{Q}. Multiplying 𝐮\mathbf{u} by a suitable number NN, we may assume that cic_{i}\in\mathbb{N}. Then, fN=i=1m(fi)cif^{N}=\prod_{i=1}^{m}(f_{i})^{c_{i}} where fi=x𝐮i¯f_{i}=\overline{x^{\mathbf{u}_{i}}}, which implies that fNgNf^{N}g^{N} is in the left hand-side, thus fgfg is in the left hand-side. ∎

To complete the proof of our main result, we need to check that the generalized Ishida complex is exact on injectives. This is stated in the following lemma, which requires three auxiliary results.

Lemma 3.6.

If MM is an injective 𝕜[x]/I\Bbbk[x]/I-module, then L𝕜[x]/IML^{\bullet}\mathbin{\mathop{\otimes}\displaylimits_{\Bbbk[x]/I}}M is exact.

Proof.

It suffices to check the case when MM is an injective indecomposable module E(𝕜[x]/P)E(\Bbbk[x]/P) over a prime ideal PP containing II. Let AiA_{i} be the ii-th column of AA. The set F:={AixiE(𝕜[x]/P)E(𝕜[x]/P)}F:=\{A_{i}\mid x_{i}E(\Bbbk[x]/P)\cong E(\Bbbk[x]/P)\} is called the face corresponding to PP.  Lemma 3.7 shows that this is indeed a face of A\mathbb{N}A. Lemma 3.9 shows that L𝕜[x]/IE(𝕜[x]/P)L\mathbin{\mathop{\otimes}\displaylimits_{\Bbbk[x]/I}}E(\Bbbk[x]/P) is exact using Lemma 3.8.

We start verifying our proposed face is a face.

Lemma 3.7.

Given a monomial prime ideal PP containing II, F:={AixiE(𝕜[x]/P)E(𝕜[x]/P)}F^{\prime}:=\{A_{i}\mid x_{i}E(\Bbbk[x]/P)\cong E(\Bbbk[x]/P)\} is a face of A\mathbb{N}A.

Proof.

Since Ass(E(𝕜[x]/P))={P}\operatorname{Ass}(E(\Bbbk[x]/P))=\{P\} and the module is indecomposable, xiE(𝕜[x]/P)x_{i}E(\Bbbk[x]/P) is either 0 if xiPx_{i}\in P or E(𝕜[x]/P)E(\Bbbk[x]/P) if xiPx_{i}\not\in P. Now suppose that the given set FF^{\prime} is not a face; then there exists a minimal face FF whose relative interior intersects with the relative interior of FF^{\prime}. Pick AjFFA_{j}\in F\setminus F^{\prime}. Then the corresponding variable xjx_{j} induces the zero map on E(𝕜[x]/P)E(\Bbbk[x]/P). If AjA_{j} is not in the relative interior of FF, let 𝐟RelInt(F)\mathbf{f}\in\operatorname{RelInt}(\mathbb{N}F) so that 𝐟=AiFciAi\mathbf{f}=\sum_{A_{i}\in F^{\prime}}c_{i}A_{i} for some cic_{i}\in\mathbb{N}. Choose a suitable N1N_{1}\in\mathbb{N} such that N1𝐟=AiFciAi+dAjN_{1}\mathbf{f}=\sum_{A_{i}\in F^{\prime}}c^{\prime}_{i}A_{i}+dA_{j} for some nonnegative cic^{\prime}_{i}\in\mathbb{Q} and d>0d\in\mathbb{Q}_{>0}. By construction, N1𝐟N_{1}\mathbf{f} is in the lattice (Lρ,ρ)(L_{\rho},\rho). Thus, there exists N2>N1N_{2}>N_{1} such that N2𝐟=AiFdiAi+dAjN_{2}\mathbf{f}=\sum_{A_{i}\in F^{\prime}}d_{i}A_{i}+d^{\prime}A_{j} where di,dd_{i},d\in\mathbb{N}, d>0d>0. But then xN2𝐟E(𝕜[x]/P)=0x^{N_{2}\mathbf{f}}E(\Bbbk[x]/P)=0, which is a contradiction. If AjA_{j} is in the relative interior of FF, a similar argument gives another contradiction. ∎

The following is necessary to prove exactness.

Lemma 3.8.

Given a face F(A)F\in\mathcal{F}(\mathbb{N}A), KJΔF:=KJΔ0FK_{J_{\Delta}}^{\cap F}:=K_{J_{\Delta}}\cap\mathbb{R}_{\geq 0}F is a face of KJΔK_{J_{\Delta}}.

Proof.

If F=AF=\mathbb{N}A, then the statement is clear. Assume dimF<dimA\dim F<\dim\mathbb{N}A. First, we claim that for any G(KJΔ)G\in\mathcal{F}(K_{J_{\Delta}}), G0FG\subseteq\mathbb{R}_{\geq 0}F if and only if G^F\widehat{G}\subseteq F. One direction follows straight from the definition of G^\widehat{G}. Conversely, assume that G^F\widehat{G}\not\subseteq F. Then, RelInt(0G^)0F=\operatorname{RelInt}(\mathbb{R}_{\geq 0}\widehat{G})\cap\mathbb{R}_{\geq 0}F=\emptyset implies RelInt(G)0F=\operatorname{RelInt}(G)\cap\mathbb{R}_{\geq 0}F=\emptyset. Therefore G0FG\not\subseteq\mathbb{R}_{\geq 0}F. This claim shows that KJΔFK_{J_{\Delta}}^{\cap F} is the union of all faces G(KJΔ)G\in\mathcal{F}(K_{J_{\Delta}}) such that G^F\widehat{G}\subseteq F. Thus KJΔFK_{J_{\Delta}}^{\cap F} can be regarded as a realized subcomplex of (KJΔF):={G(KJΔ):G^F}\mathcal{F}(K_{J_{\Delta}}^{\cap F}):=\{G\in\mathcal{F}(K_{J_{\Delta}}):\widehat{G}\subseteq F\}.

Next, we claim that (KJΔF)\mathcal{F}(K_{J_{\Delta}}^{\cap F}) has a unique maximal element. Suppose not; let G1G_{1} and G2G_{2} be two distinct faces of (KJΔF)\mathcal{F}(K_{J_{\Delta}}^{\cap F}) of maximal dimension. Then, FG1^+G2^F\supseteq\widehat{G_{1}}+\widehat{G_{2}} implies that there is a face G1^G2^(A)\widehat{G_{1}}\vee\widehat{G_{2}}\in\mathcal{F}(\mathbb{N}A) such that FG1^G2^F\supseteq\widehat{G_{1}}\vee\widehat{G_{2}}, the join of the two faces. Since G1G_{1} and G2G_{2} are distinct and the same dimension, G1G2G1G_{1}\vee G_{2}\neq G_{1} or G2G_{2}. Therefore G1G2G1G2G_{1}\vee G_{2}\supsetneq G_{1}\cup G_{2}. Hence, 0(G1^G2^)0(G1G2^)\mathbb{R}_{\geq 0}(\widehat{G_{1}}\vee\widehat{G_{2}})\supseteq\mathbb{R}_{\geq 0}(\widehat{G_{1}\vee G_{2}}), which implies FG1^G2^G1G2^F\supseteq\widehat{G_{1}}\vee\widehat{G_{2}}\supseteq\widehat{G_{1}\vee G_{2}}. Hence, G1G20FG_{1}\vee G_{2}\subseteq\mathbb{R}_{\geq 0}F, therefore G1G2(KJΔF)G_{1}\vee G_{2}\in\mathcal{F}(K_{J_{\Delta}}^{\cap F}), contradicting the maximality of G1G_{1} and G2G_{2}. We conclude (KJΔF)\mathcal{F}(K_{J_{\Delta}}^{\cap F}) has a unique maximal element, say HH.

Lastly, we claim that (H)=(KJΔF)\mathcal{F}(H)=\mathcal{F}(K_{J_{\Delta}}^{\cap F}), which implies KJΔF=HK_{J_{\Delta}}^{\cap F}=H. For any G(KJΔF)G\in\mathcal{F}(K_{J_{\Delta}}^{\cap F}), let G:=GHG^{\prime}:=G\vee H in (KJΔ)\mathcal{F}(K_{J_{\Delta}}). Then, RelInt(G^)0(G^H^)RelInt(G^)0FG^FG(KJΔF)\emptyset\neq\operatorname{RelInt}(\widehat{G^{\prime}})\cap\mathbb{R}_{\geq 0}(\widehat{G}\cup\widehat{H})\subseteq\operatorname{RelInt}(\widehat{G^{\prime}})\cap\mathbb{R}_{\geq 0}F\implies\widehat{G^{\prime}}\subseteq F\implies G^{\prime}\in\mathcal{F}(K_{J_{\Delta}}^{\cap F}). By the maximality of HH, G=HG^{\prime}=H, which implies GHG\subseteq H. ∎

Lemma 3.9.

Given a monomial prime ideal PP whose corresponding face is FF, for k1k\geq 1,

Lk𝕜[x]/IE(𝕜[x]/P)=G(KJΔF)k1E(𝕜[x]/P)Hom(𝒞~(KJΔF)(1),E(𝕜[x]/P))L^{k}\mathbin{\mathop{\otimes}\displaylimits_{\Bbbk[x]/I}}E(\Bbbk[x]/P)=\bigoplus_{G\in\mathcal{F}\left(K_{J_{\Delta}}^{\cap F}\right)^{k-1}}E(\Bbbk[x]/P)\cong\operatorname{Hom}_{\mathbb{Z}}\left(\tilde{\mathcal{C}}\left(K_{J_{\Delta}}^{\cap F}\right)(-1),E\left(\Bbbk[x]/P\right)\right)

where 𝒞~(KJΔF)\tilde{\mathcal{C}}\left(K_{J_{\Delta}}^{\cap F}\right) is the reduced chain complex of KJΔFK_{J_{\Delta}}^{\cap F} as a CW complex.

Proof.

Lemma 3.7 shows that for any F,G(A)F,G\in\mathcal{F}(\mathbb{N}A),

E(𝕜[x]/P)𝕜[x]/I(𝕜[x]/I)G={0 if GFE(𝕜[x]/P) if GF.E(\Bbbk[x]/P)\mathbin{\mathop{\otimes}\displaylimits_{\Bbbk[x]/I}}\left(\Bbbk[x]/I\right)_{G}=\begin{cases}0&\text{ if }G\not\subseteq F\\ E(\Bbbk[x]/P)&\text{ if }G\subseteq F\end{cases}.

If FF is not the image of a face in KJΔK_{J_{\Delta}}, then no sub-face of FF is the image of a face of KJΔK_{J_{\Delta}}. Otherwise, there is a face GFG\subseteq F containing an unbounded face of 0{uζxuJΔ}\mathbb{R}_{\geq 0}\{u\in\mathbb{Z}^{\zeta}\mid x^{u}\in J_{\Delta}\}. Then by the correspondence between radical monomial ideals and subcomplexes of (Q)\mathcal{F}(Q), RelInt(F)\operatorname{RelInt}(F) contains an element of an unbounded face of 0{uζxuJΔ}\mathbb{R}_{\geq 0}\{u\in\mathbb{Z}^{\zeta}\mid x^{u}\in J_{\Delta}\}, a contradiction. Therefore, no images of faces are subsets of FF. This implies that KJΔF=0K_{J_{\Delta}}^{\cap F}=0 and Lk𝕜[A]E(𝕜[x]/P)=0L^{k}\mathbin{\mathop{\otimes}\displaylimits_{\Bbbk[\mathbb{N}A]}}E(\Bbbk[x]/P)=0 for k1k\geq 1.

Otherwise, F:=F^F:=\widehat{F^{\prime}} for some F(KJΔ)F^{\prime}\in\mathcal{F}(K_{J_{\Delta}}). Since G^F\widehat{G^{\prime}}\subseteq F if and only if G(KJΔF)G^{\prime}\in\mathcal{F}(K_{J_{\Delta}}^{\cap F}) by Lemma 3.8, so the first equality holds.

Now observe that Hom(,E(𝕜[x]/P))E(𝕜[x]/P)\operatorname{Hom}_{\mathbb{Z}}(\mathbb{Z},E(\Bbbk[x]/P))\cong E(\Bbbk[x]/P) as a 𝕜[A]\Bbbk[\mathbb{N}A]-module. Thus,

L𝕜[A]E(𝕜[x]/P)=G(KJΔF)1E(𝕜[x]/P)G(KJΔF)1Hom(,E(𝕜[x]/P))\displaystyle L^{\bullet}\mathbin{\mathop{\otimes}\displaylimits_{\Bbbk[\mathbb{N}A]}}E(\Bbbk[x]/P)=\bigoplus_{G\in\mathcal{F}(K_{J_{\Delta}}^{\cap F})^{\bullet-1}}E(\Bbbk[x]/P)\cong\bigoplus_{G\in\mathcal{F}(K_{J_{\Delta}}^{\cap F})^{\bullet-1}}\operatorname{Hom}_{\mathbb{Z}}(\mathbb{Z},E(\Bbbk[x]/P))
Hom(G(KJΔF)1,E(𝕜[x]/P))=Hom(𝒞~(KJΔF)(1),E(𝕜[x]/P)).\displaystyle\cong\operatorname{Hom}_{\mathbb{Z}}\left(\bigoplus_{G\in\mathcal{F}(K_{J_{\Delta}}^{\cap F})^{\bullet-1}}\mathbb{Z},E(\Bbbk[x]/P)\right)=\operatorname{Hom}_{\mathbb{Z}}\left(\tilde{\mathcal{C}}\left(K_{J_{\Delta}}^{\cap F}\right)(-1),E(\Bbbk[x]/P)\right).

4. Local cohomology with monomial support for cellular binomial ideals

In this section we express the Hilbert series of the local cohomology with monomial support of 𝕜[x]/I\Bbbk[x]/I as a (formal) finite sum of rational functions when II is a lattice ideal (Theorem 4.2) or a cellular binomial ideal (Theorem 4.3). As a corollary, we provide a generalization of Reisner’s criterion to the context of cellular binomial ideals, which gives a Cohen-Macaulay characterization for 𝕜[x]/I\Bbbk[x]/I in terms of the cohomology of finitely many chain complexes (Corollary 4.5). Let (I,J,A)(I,J,A) be a tuple consisting of a lattice ideal (resp. cellular binomial ideal), a minimal prime ideal JJ of I𝕜[ζ]I\cap\Bbbk[\mathbb{N}^{\zeta}], and the corresponding affine semigroup A=Q\mathbb{N}A=Q. Then JJ is also a prime lattice ideal and we may assume after rescaling the variables that J=I(ξ)J=I(\xi) is toric, with lattice Lξ=(Lρ)satL_{\xi}=(L_{\rho})_{\text{sat}}. Let T:=Lξ/LρT:=L_{\xi}/L_{\rho} be the corresponding torsion abelian group, then

d/LρTA.\mathbb{Z}^{d}/L_{\rho}\cong T\oplus\mathbb{Z}A.

We may induce a fine grading of 𝕜[x]/I\Bbbk[x]/I by TAT\oplus\mathbb{Z}A as follows: for any x𝐮¯𝕜[x]/I\overline{x^{\mathbf{u}}}\in\Bbbk[x]/I for some 𝐮d\mathbf{u}\in\mathbb{Z}^{d}, degT,A(x𝐮¯):=(𝐮+Lρ,A𝐮)\deg_{T,A}(\overline{x^{\mathbf{u}}}):=(\mathbf{u}+L_{\rho},A\cdot\mathbf{u}). Here we use x𝐮¯\overline{x^{\mathbf{u}}} to indicate the image of x𝐮𝕜[x]x^{\mathbf{u}}\in\Bbbk[x] in 𝕜[x]/I\Bbbk[x]/I.

Let \mathcal{I} be a monomial ideal of an affine semigroup ring 𝕜[Q]\Bbbk[Q]. Let deg(𝕜[Q]/):={𝐚d(𝕜[Q]/)𝐚0}\deg(\Bbbk[Q]/\mathcal{I}):=\{\mathbf{a}\in\mathbb{Z}^{d}\mid(\Bbbk[Q]/\mathcal{I})_{\mathbf{a}}\neq 0\}. A proper pair (𝐚,F)(\mathbf{a},F) of II is a pair such that 𝐚+Fdeg(𝕜[Q]/)\mathbf{a}+\mathbb{N}F\subset\deg(\Bbbk[Q]/\mathcal{I}). Given two pairs (𝐚,F)(\mathbf{a},F) and (𝐛,G)(\mathbf{b},G), we say (𝐚,F)<(𝐛,G)(\mathbf{a},F)<(\mathbf{b},G) if 𝐚+F𝐛+G\mathbf{a}+\mathbb{N}F\subset\mathbf{b}+\mathbb{N}G. A degree pair of \mathcal{I} is a maximal element of the set of all proper pairs with the given order. Two pairs (𝐚,F)(\mathbf{a},F) and (𝐛,F)(\mathbf{b},F) with the same face FF overlap if the intersection (𝐚+F)(𝐛+F)(\mathbf{a}+\mathbb{N}F)\cap(\mathbf{b}+\mathbb{N}F) is nonempty. Overlapping is an equivalence relation on pairs; an overlap class [𝐚,F][\mathbf{a},F] is an equivalence class containing the degree pair (𝐚,F)(\mathbf{a},F). Let [𝐚,F]\bigcup[\mathbf{a},F] be the set of all degrees belonging to 𝐚+F\mathbf{a}^{\prime}+\mathbb{N}F for some degree pair (𝐚,F)(\mathbf{a}^{\prime},\mathbb{N}F) in [𝐚,F][\mathbf{a},F]. The original degree space deg(𝕜[Q]/)\bigcup\deg(\Bbbk[Q]/\mathcal{I}) is the union of deg((𝕜[Q]/)P)\deg((\Bbbk[Q]/\mathcal{I})_{P}) for all monomial prime ideals PP of 𝕜[Q]\Bbbk[Q]. The degree pair topology is the smallest topology on deg(𝕜[Q]/)\bigcup\deg(\Bbbk[Q]/\mathcal{I}) such that for any overlap class [𝐚,F][\mathbf{a},F] of any localization (𝕜[Q]/)P(\Bbbk[Q]/\mathcal{I})_{P}, the set [𝐚,F]\bigcup[\mathbf{a},F] is both open and closed. These notions were introduced in [STV95, STDPAIR, MY2022].

In this article, we extend the notion of degree space further. Suppose that 𝒜\mathcal{A} is the hyperplane arrangement consisting of the supporting hyperplanes of the facets of 0Q\mathbb{R}_{\geq 0}Q. Let 𝔯(𝒜)\mathfrak{r}(\mathcal{A}) be the set of regions of 𝒜\mathcal{A}, where a region is a connected components of d𝒜\mathbb{R}^{d}\setminus\bigcup_{\mathscr{H}\in\mathcal{A}}\mathscr{H} [Stanley07]. Then, for any region 𝔯𝔯(𝒜)\mathfrak{r}\in\mathfrak{r}(\mathcal{A}), regard 𝔯(ddeg(𝕜[Q]/))\mathfrak{r}\cap(\mathbb{Z}^{d}\smallsetminus\bigcup\deg(\Bbbk[Q]/\mathcal{I})) as a space with the trivial topology. The extended degree space Q\mathbb{Z}Q of \mathcal{I} is the disjoint union Q=(deg(𝕜[Q]/))(𝔯𝔯(𝒜)𝔯(ddeg(𝕜[Q]/))\mathbb{Z}Q=\left(\bigcup\deg(\Bbbk[Q]/\mathcal{I})\right)\cup\big{(}\bigcup_{\mathfrak{r}\in\mathfrak{r}(\mathcal{A})}\mathfrak{r}\cap(\mathbb{Z}^{d}\smallsetminus\bigcup\deg(\Bbbk[Q]/\mathcal{I})\big{)} as a topological space. (As a set, Q\mathbb{Z}Q equals d\mathbb{Z}^{d}.) Lastly, let 𝒢(Q)\mathcal{G}(\mathbb{Z}Q) (resp. 𝒢(𝕜[Q]/)\mathcal{G}(\Bbbk[Q]/\mathcal{I})) be the collection of all minimal open sets of the extended (resp. original) degree space of \mathcal{I}. Since 𝒢(𝕜[Q]/)\mathcal{G}(\Bbbk[Q]/\mathcal{I}) is finite [MY2022]*Lemma 4.3 and 𝔯(𝒜)\mathfrak{r}(\mathcal{A}) is finite, 𝒢(Q)\mathcal{G}(\mathbb{Z}Q) is also finite.

Given t:=𝐮+Lρt:=\mathbf{u}+L_{\rho}, let t\mathcal{I}_{t} be the following ideal in 𝕜[Q]\Bbbk[Q]

t:=xA𝐮degT,A(x𝐮¯)=(t,A𝐮) and x𝐮¯𝕜[x]/.\mathcal{I}_{t}:=\langle x^{A\cdot\mathbf{u}}\mid\deg_{T,A}(\overline{x^{\mathbf{u}}})=(t,A\cdot\mathbf{u})\text{ and }\overline{x^{\mathbf{u}}}\in\Bbbk[x]/\mathcal{I}\rangle.

For an open set O𝒢(𝕜[x]/t)O\in\mathcal{G}(\Bbbk[x]/\mathcal{I}_{t}), let 𝒞~(O)\tilde{\mathcal{C}}\left(O\right) be the graded part of the generalized Ishida complex associated to the element x𝐮¯\overline{x^{\mathbf{u}}} whose torsion degree is tt and whose AA-degree is A𝐮OA\mathbf{u}\in O. This is well-defined regardless of choice of x𝐮¯\overline{x^{\mathbf{u}}}, as is stated below.

Lemma 4.1.

If x𝐮¯\overline{x^{\mathbf{u}}} and x𝐯¯\overline{x^{\mathbf{v}}} with the same torsion degree are in the same minimal open set of the extended degree space of t\mathcal{I}_{t}, then their corresponding graded parts of the Ishida complex coincide.

Proof.

If neither A𝐮A\cdot\mathbf{u} nor A𝐯A\cdot\mathbf{v} are in the original degree space of t\mathcal{I}_{t}, they must be in the same region of the hyperplane arrangement. Hence, for any localization by a face FF, either both degrees are deg((t)𝕜[AF])\deg((\mathcal{I}_{t})\cdot\Bbbk[\mathbb{N}A-\mathbb{N}F]) or they do not belong to the same localization.

If both are in the original degree space, let [𝐚,F][\mathbf{a},F] be an overlap class whose degree set [𝐚,F]\bigcup[\mathbf{a},F] contains A𝐮A\cdot\mathbf{u} and A𝐯A\cdot\mathbf{v}, and such that FF is minimal with this property. Then x𝐮¯\overline{x^{\mathbf{u}}} and x𝐯¯\overline{x^{\mathbf{v}}} do not appear in the localization of 𝕜[x]/I\Bbbk[x]/I by a multiplicative set generated by variables corresponding to a proper face of FF. Conversely, if there is no overlap class [𝐚,G][\mathbf{a},G] containing A𝐯A\cdot\mathbf{v}, this means that A𝐯A\cdot\mathbf{v} appears on every localization of QQ by faces containing GG. This completely determines the graded parts of the Ishida complex. ∎

Theorem 4.2.

Given a lattice ideal II, define QQ as before. T he multi-graded Hilbert series for the iith local cohomology module of 𝕜[x]/I\Bbbk[x]/I supported on the inverse image of the radical monomial ideal JΔ𝕜[Q]J_{\Delta}\subset\Bbbk[Q] with respect to the TAT\otimes\mathbb{Z}A-grading is

Hilb(HJΔi(𝕜[x]/I),𝐭)=tTO𝒢(𝕜[A]/t)dimHi(𝒞~(O);𝕜)udAuOxu¯.\operatorname{Hilb}(H_{J_{\Delta}}^{i}(\Bbbk[x]/I),\mathbf{t})=\sum_{t\in T}\sum_{O\in\mathcal{G}(\Bbbk[\mathbb{N}A]/\mathcal{I}_{t})}\dim H^{i}(\tilde{\mathcal{C}}\left(O\right);\Bbbk)\sum_{\begin{subarray}{c}u\in\mathbb{Z}^{d}\\ Au\in O\end{subarray}}x^{\overline{u}}.
Proof.

This is a direct consequence of Lemma 4.1. ∎

Since TT is finite and 𝒢(𝕜[x]/t)\mathcal{G}(\Bbbk[x]/\mathcal{I}_{t}) is finite for all tTt\in T, the sum is finite. Moreover, each minimal open set O𝒢(𝕜[x]/t)O\in\mathcal{G}(\Bbbk[x]/\mathcal{I}_{t}) is the set of lattice points in a convex polyhedron, so that udAuOxu¯\sum_{\begin{subarray}{c}u\in\mathbb{Z}^{d}\\ Au\in O\end{subarray}}x^{\overline{u}} can be written as a rational function [Barvinok99, Barvinok03].

For the case of a ζ\zeta-cellular binomial ideal (I,J,A)(I,J,A), let MM be the multiplicative set consisting of monomials on the nilpotent variables of 𝕜[x]/I\Bbbk[x]/I. Then for each mMm\in M, (I:m)𝕜[ζ](I:m)\cap\Bbbk[\mathbb{N}^{\zeta}] is a lattice ideal containing I𝕜[ζ]I\cap\Bbbk[\mathbb{N}^{\zeta}]. Hence, according to Theorem 2.1, we may pick an associated prime ideal JmJ_{m} of I𝕜[ζ]I\cap\Bbbk[\mathbb{N}^{\zeta}] whose extension Jm𝕜[x]+xiiζcJ_{m}\Bbbk[x]+\langle x_{i}\mid i\in\zeta^{c}\rangle is the associated prime containing JJ. Moreover, the quotient 𝕜[x]/(Jm𝕜[x]+xiiζc)\Bbbk[x]/(J_{m}\Bbbk[x]+\langle x_{i}\mid i\in\zeta^{c}\rangle) is isomorphic to an affine semigroup ring 𝕜[Am]\Bbbk[\mathbb{N}A_{m}] for some integer matrix AmA_{m}. The canonical projection 𝕜[x]/J𝕜[x]/(Jm𝕜[x]+xiiζc)\Bbbk[x]/J\to\Bbbk[x]/(J_{m}\Bbbk[x]+\langle x_{i}\mid i\in\zeta^{c}\rangle) induces a monoid map AAm\mathbb{N}A\to\mathbb{N}A_{m}. By letting TmT_{m} be the quotient of the saturation of the lattice LmL_{m} corresponding to (I:m)𝕜[ζ](I:m)\cap\Bbbk[\mathbb{N}^{\zeta}] by LmL_{m}, we have a fine grading of 𝕜[x]/I\Bbbk[x]/I by the abelian group

mM(TmAm)\bigoplus_{m\in M}(T_{m}\oplus\mathbb{Z}A_{m})

via degM,T,A(x𝐮¯)=(𝐮ζc,𝐮ζ+Lm,Am𝐮ζ).\deg_{M,T,A}(\overline{x^{\mathbf{u}}})=(\mathbf{u}^{\zeta^{c}},\mathbf{u}^{\zeta}+L_{m},A_{m}\cdot\mathbf{u}^{\zeta}). Thus, for a fixed mMm\in M and a torsion tTmt\in T_{m}, let

It:=xAm𝐮ζdegM,T,A(x𝐮¯)=(deg(m),t,A𝐮ζ) and x𝐮¯𝕜[x]/I.I_{t}:=\langle x^{A_{m}\cdot\mathbf{u}^{\zeta}}\mid\deg_{M,T,A}(\overline{x^{\mathbf{u}}})=(\deg(m),t,A\cdot\mathbf{u}^{\zeta})\text{ and }\overline{x^{\mathbf{u}}}\in\Bbbk[x]/I\rangle.

Using the same arguments as for lattice ideals, we know that two elements whose degrees are in the same minimal open set O𝒢(𝕜[Am]/It)O\in\mathcal{G}(\Bbbk[\mathbb{N}A_{m}]/I_{t}) have the same graded part of the Ishida complex 𝒞~(O)\tilde{\mathcal{C}}\left(O\right). Then, we have

Theorem 4.3.

Given a cellular binomial ideal II, the multi-graded Hilbert series for the local cohomology module of 𝕜[x]/I\Bbbk[x]/I supported on the image of the radical monomial ideal JΔ𝕜[Q]J_{\Delta}\subset\Bbbk[Q] with respect to the mM(TmAm)\bigoplus_{m\in M}(T_{m}\oplus\mathbb{Z}A_{m})-grading is

Hilb(HJΔi(𝕜[x]/I),𝐭)=mMtTmO𝒢(𝕜[Am]/It)dimHi(𝒞~(O);𝕜)udAmuζOxu¯.\operatorname{Hilb}(H_{J_{\Delta}}^{i}(\Bbbk[x]/I),\mathbf{t})=\sum_{m\in M}\sum_{t\in T_{m}}\sum_{O\in\mathcal{G}(\Bbbk[\mathbb{N}A_{m}]/I_{t})}\dim H^{i}(\tilde{\mathcal{C}}\left(O\right);\Bbbk)\sum_{\begin{subarray}{c}u\in\mathbb{Z}^{d}\\ A_{m}\cdot u^{\zeta}\in O\end{subarray}}\overline{x^{u}}.

Again, this is a finite sum of rational functions. The Corollary 4.5 gives us the equivalent of Reisner’s criterion for cellular binomial ideals, providing a characterization of Cohen-Macaulayness in terms of the cohomology of finitely many polyhedral complexes. First we need an auxilliary result.

Lemma 4.4.

𝒞~(O)\tilde{\mathcal{C}}\left(O\right) is the cochain complex of a polyhedral complex.

Proof.

It suffices to show that the nontrivial top dimensional part of 𝒞~(O)\tilde{\mathcal{C}}\left(O\right) is 𝕜1\Bbbk^{1}; suppose not; then there exists distinct maximal faces F1^\widehat{F_{1}} and F2^\widehat{F_{2}} of QQ such that degA(xu1¯)F1^\deg_{A}(\overline{x^{u_{1}}})\in\widehat{F_{1}} and degA(xu2¯)F2^\deg_{A}(\overline{x^{u_{2}}})\in\widehat{F_{2}} for some distinct xu1¯\overline{x^{u_{1}}} and xu2¯\overline{x^{u_{2}}} with u1,u2Ou_{1},u_{2}\in O. Then, the degree of the product xu1¯xu2¯\overline{x^{u_{1}}}\cdot\overline{x^{u_{2}}} lies in the relative interior of a face G^\widehat{G} in (Q)\mathcal{F}(Q) which is a minimal face containing both F1^\widehat{F_{1}} and F2^\widehat{F_{2}}, contradicting the maximality of F1^\widehat{F_{1}} and F2^\widehat{F_{2}}. ∎

Corollary 4.5.

Let II be a cellular binomial ideal. Then 𝕜[x]/I\Bbbk[x]/I is Cohen–Macaulay if and only if Hi(𝒞~(O);𝕜)=0H^{i}(\tilde{\mathcal{C}}\left(O\right);\Bbbk)=0 for all idim(𝕜[x]/I)i\neq\text{dim}(\Bbbk[x]/I) and for all O𝒢(𝕜[Am]/It)O\in\mathcal{G}(\Bbbk[\mathbb{N}A_{m}]/I_{t}), mM,tTmm\in M,t\in T_{m}. ∎

Example 4.6.

Let LL be the following lattice in 4\mathbb{Z}^{4} and LsatL_{\text{sat}} its saturation.

L:=(2030),(1515),Lsat:=(1111),(0212),(1122),(2030).L:=\left\langle\left(\begin{smallmatrix}2\\ 0\\ -3\\ 0\end{smallmatrix}\right),\left(\begin{smallmatrix}1\\ -5\\ 1\\ 5\end{smallmatrix}\right)\right\rangle,L_{\text{sat}}:=\left\langle\left(\begin{smallmatrix}1\\ -1\\ -1\\ 1\end{smallmatrix}\right),\left(\begin{smallmatrix}0\\ -2\\ 1\\ 2\end{smallmatrix}\right),\left(\begin{smallmatrix}-1\\ -1\\ 2\\ 2\end{smallmatrix}\right),\left(\begin{smallmatrix}-2\\ 0\\ 3\\ 0\end{smallmatrix}\right)\right\rangle.

The torsion group T:=L/LsatT:=L/L_{\text{sat}} is isomorphic to /5\mathbb{Z}/5\mathbb{Z}. We represent TT as follows

T={e=(0000)¯,ξ=(1111)¯,ξ2=(0212)¯,ξ3=(1303)¯,ξ4=(0424)¯}T=\left\{e=\overline{\left(\begin{smallmatrix}0\\ 0\\ 0\\ 0\end{smallmatrix}\right)},\xi=\overline{\left(\begin{smallmatrix}-1\\ 1\\ 1\\ -1\end{smallmatrix}\right)},\xi^{2}=\overline{\left(\begin{smallmatrix}0\\ 2\\ -1\\ -2\end{smallmatrix}\right)},\xi^{3}=\overline{\left(\begin{smallmatrix}-1\\ 3\\ 0\\ -3\end{smallmatrix}\right)},\xi^{4}=\overline{\left(\begin{smallmatrix}0\\ 4\\ -2\\ -4\end{smallmatrix}\right)}\right\}

using GRevLex term order in Macaulay2.

In this case Q:=(4/Lsat)4[31200101]Q:=(\mathbb{Z}^{4}/L_{\text{sat}})\cap\mathbb{N}^{4}\cong\mathbb{N}\left[\begin{smallmatrix}3&1&2&0\\ 0&1&0&1\end{smallmatrix}\right]. As usual Q=4/Lsat\mathbb{Z}Q=\mathbb{Z}^{4}/L_{\text{sat}}. On the polynomial ring 𝕜[a,b,c,d]\Bbbk[a,b,c,d], the lattice ideals corresponding to LL and LsatL_{\text{sat}} are

IL:=a2c3,acd5b5 and Isat:=bcad,cd2b2,c2dab,c3a2.I_{L}:=\langle a^{2}-c^{3},acd^{5}-b^{5}\rangle\text{ and }I_{\text{sat}}:=\langle bc-ad,cd^{2}-b^{2},c^{2}d-ab,c^{3}-a^{2}\rangle.

Hence, the Ishida complex supported on the maximal monomial ideal is

0𝕜[x]/IL(𝕜[x]/IL)a,c(𝕜[x]/IL)d(𝕜[x]/IL)a,b,c,d0.0\to\Bbbk[x]/I_{L}\to(\Bbbk[x]/I_{L})_{a,c}\oplus(\Bbbk[x]/I_{L})_{d}\to(\Bbbk[x]/I_{L})_{a,b,c,d}\to 0.

Here, we see that for any 0i0\leq i,

Ct,[ij]:0𝕂j+1𝕂j+1+5𝕂50C_{t,\left[\begin{smallmatrix}i\\ j\end{smallmatrix}\right]}^{\bullet}:0\to\mathbb{K}^{j+1}\to\mathbb{K}^{j+1+5}\to\mathbb{K}^{5}\to 0

when 0j<50\leq j<5 and

Ct,[ij]:0𝕂5𝕂10𝕂50C_{t,\left[\begin{smallmatrix}i\\ j\end{smallmatrix}\right]}^{\bullet}:0\to\mathbb{K}^{5}\to\mathbb{K}^{10}\to\mathbb{K}^{5}\to 0

when 5j5\leq j for any tTt\in T. Since there is no non-top cohomology, we conclude that 𝕜[x]/IL\Bbbk[x]/I_{L} is Cohen-Macaulay.

5. Local cohomology for affine semigroups over simplices

In this section, we consider affine semigroup rings such that 0Q\mathbb{R}_{\geq 0}Q is a cone over a simplex. In this case the combinatorics simplifies and becomes more explicit.

5.1. Hyperplane arrangements

The hyperplane arrangement 𝒜:={1,,m}\mathcal{A}:=\{\mathscr{H}_{1},\cdots,\mathscr{H}_{m}\} of a polyhedron 𝒫\mathcal{P} is the collection of supporting hyperplanes of the facets of 𝒫\mathcal{P} in d\mathbb{R}^{d}. In this case, i=1m+=𝒫\bigcap_{i=1}^{m}\mathscr{H}^{+}=\mathcal{P}, where +\mathscr{H}^{+} denotes the positive half-space associated to a hyperplane (the positive side is the side that contains our polyhedron). A hyperplane arrangement is linear if all hyperplanes in the arrangement contain the origin. A region 𝔯\mathfrak{r} of a hyperplane arrangement 𝒜\mathcal{A} is a connected component of d𝒜\mathbb{R}^{d}-\bigcup_{\mathscr{H}\in\mathcal{A}}\mathscr{H}. 𝔯(𝒜)\mathfrak{r}(\mathcal{A}) refers to the collection of all regions of 𝒜\mathcal{A}.

Suppose 𝒜\mathcal{A} consists of a minimal number of hyperplanes which generate a rational polyhedral cone 𝒫\mathcal{P}. Then 𝒜\mathcal{A} is linear and all regions in 𝔯(𝒜)\mathfrak{r}(\mathcal{A}) are unbounded rational polyhedral cones. Moreover, every region 𝔯\mathfrak{r} can be expressed as

𝔯S:=(i[m]Si+)(iSi)i=1mi\mathfrak{r}_{S}:=\left(\bigcap_{i\in[m]\smallsetminus S}\mathscr{H}_{i}^{+}\right)\cap\left(\bigcap_{i\in S}\mathscr{H}_{i}^{-}\right)\smallsetminus\bigcup_{i=1}^{m}\mathscr{H}_{i}

for a subset S[m]S\subseteq[m] where i\mathscr{H}_{i}^{-} is the complement of i+\mathscr{H}_{i}^{+}. In other words, a region is labeled by the collection of hyperplanes whose positive half space contains it. It follows that 𝔯(𝒜)\mathfrak{r}(\mathcal{A}) is partially ordered by reverse inclusion on the set of labels; 𝔯S1𝔯S2 if S1S2.\mathfrak{r}_{S_{1}}\leq\mathfrak{r}_{S_{2}}\text{ if }S_{1}\supseteq S_{2}. We call 𝔯(𝒜)\mathfrak{r}(\mathcal{A}) the poset of regions of 𝒜\mathcal{A}. Our notation here is consistent with that of  [Edelman84, BEZ90] regarding 𝒫\mathcal{P} as the base region. Moreover, the natural embedding (𝒫)𝔯(𝒜)\mathcal{F}(\mathcal{P})\to\mathfrak{r}(\mathcal{A}) sending a face FF to the set of indices of hyperplanes containing FF exists [Edelman84]*Lemma 1.3. Since we are interested in regions partitioning d\mathbb{R}^{d} along with the set of degrees of standard monomials of localizations, we modify the definition of 𝔯S\mathfrak{r}_{S} as follows, to include boundaries:

𝔯S:=(i[m]Si+)(iSi),\mathfrak{r}_{S}:=\left(\bigcap_{i\in[m]\smallsetminus S}\mathscr{H}_{i}^{+}\right)\cap\left(\bigcap_{i\in S}\mathscr{H}_{i}^{-}\right),

A cumulative region S=(i[m]Si+)\mathfrak{R}_{S}=\left(\bigcap_{i\in[m]\setminus S}\mathscr{H}_{i}^{+}\right) is the union of all regions less then 𝔯S\mathfrak{r}_{S}. The poset of cumulative regions (𝒜)\mathfrak{R}(\mathcal{A}) is a set of all cumulative regions ordered by inclusion. By definition, (𝒜)𝔯(𝒜)\mathfrak{R}(\mathcal{A})\cong\mathfrak{r}(\mathcal{A}) as posets. We follow the conventions of [Stanley07].

5.2. Sections of polyhedra

If FF is a zero-dimensional face of a polyhedral complex 𝒫\mathcal{P}, then we define the vertex figure of 𝒫\mathcal{P} at FF as follows. Assume 𝒫\mathcal{P} is realized in d\mathbb{R}^{d}, Pick a sphere Sd1S^{d-1} centered at FF such that every nonempty face of 𝒫\mathcal{P} except FF is not completely contained in the sphere. Then, the vertex figure 𝒫/F\mathcal{P}/F is the polyhedral complex generated by the intersections of all faces of 𝒫\mathcal{P} containing FF on Sd1S^{d-1}. Likewise, when FF is positive dimensional face, we define the section of 𝒫\mathcal{P} at FF, denoted 𝒫/F\mathcal{P}/F, to be the polytope generated by taking vertex figures iteratively over the vertices arising from FF. As a new abstract polytope, G/FG/F is combinatorially equivalent to the link of FF over GG as a sub-polyhedral complex of 𝒫\mathcal{P}. Moreover, if we regard the face lattice of the vertex figure as a collection of faces of 𝒫\mathcal{P}, then its closure in the usual Euclidean topology agrees with the star of FF.

Given a polyhedral complex Δ\Delta, let max(Δ)\max(\Delta) be the set of maximal elements of Δ\Delta, and let max(Δ)\bigcap\max(\Delta) be a set of faces which are intersections of maximal faces of Δ\Delta as a poset. We say Δ\Delta is mm-combinatorially connected for m:=minFmax(Δ)dimFm:=\min_{F\in\bigcap\max(\Delta)}\dim F. Note that this is a much finer notion of connectedness than the usual topological nn-connectedness. For example, a simplicial complex consisting of two triangles sharing an edge is 11-combinatorially connected but contractible (infinitely-connected) in the sense of topological nn-connectedness.

Lemma 5.1.

Any vertex figure of an mm-combinatorially connected polyhedral complex Δ\Delta is at least (m1)(m-1)-combinatorially connected. Moreover, when m1m\geq 1, for any vertex FΔF\in\Delta, the vertex figure (Δ/F)(\Delta/F) is contractible. Hence, for any face FΔF\in\Delta whose dimension is less than mm, (K/F)Δ(K/F)\cap\Delta is contractible.

Proof.

As the m=0m=0 case holds vacuously, we may let m1m\geq 1. Pick a vertex FF. Let GG be the intersection of all maximal faces of Δ\Delta containing FF. Then, dimGm.\dim G\geq m. All maximal faces of the vertex figure Δ/F\Delta/F are inherited from those maximal faces of Δ\Delta containing FF, hence Δ/F\Delta/F is (dimG1)(\dim G-1)-combinatorially connected, and dimG1m1\dim G-1\geq m-1.

For the second statement, let FF and GG be the same as above. Let X={HΔHF}X=\{H\in\Delta\mid H\supseteq F\} as a set of faces of Δ\Delta. Since XX is homotopic to GG; we may collapse each maximal face in XX containing GG continuously to GG. Moreover, X{F}X\smallsetminus\{F\} is homotopic to G{F}G\smallsetminus\{F\}. To see this, let SFS_{F} be a sphere centered at FF and generating the vertex figure of Δ\Delta and GG on its surface. The homotopy from XX to GG restricted on SFS_{F} gives the homotopy between vertex figures of XX and GG over FF if dimG1\dim G\geq 1. The last statement follows by applying the second statement iteratively. ∎

Lemma 5.2.

Given a dd-dimensional polyhedral complex 𝒫\mathcal{P} homeomorphic to a disk DdD^{d}, let FF be a vertex. If FF is in the interior of 𝒫\mathcal{P}, then 𝒫/F\mathcal{P}/F is homeomorphic to Sd1S^{d-1}. Otherwise, if FF is in the boundary of 𝒫\mathcal{P}, then 𝒫/F\mathcal{P}/F is homeomorphic to Dd1D^{d-1}.

Proof.

If FF is in the boundary of 𝒫\mathcal{P}, the result is clear since the sphere containing FF that defines the vertex figure cannot be contained in the polyhedral complex. One can apply a similar argument to the case when FF is in the relative interior. ∎

5.3. Local cohomology

Let (I,I,A)(I,I,A) be a prime lattice ideal II whose corresponding affine semigroup Q=AQ=\mathbb{N}A is simplicial, i.e., the transverse section KK of the polyhedral cone 0Q\mathbb{R}_{\geq 0}Q is a dd-simplex. Let 𝒜={i}i=1d\mathcal{A}=\{\mathscr{H}_{i}\}_{i=1}^{d} be the minimal hyperplane arrangement of 0Q\mathbb{R}_{\geq 0}Q. Label the faces of KK by their supporting facets; for a facet iK\mathscr{H}_{i}\cap K, use the label [d]{i}[d]\smallsetminus\{i\}. Hence, the zero-dimensional face (of the transverse section) that does not lie in the hyperplane i\mathscr{H}_{i} is indexed by {i}.\{i\}.

From the natural isomorphism (Q)(K)\mathcal{F}(Q)\cong\mathcal{F}(K), we may label faces of QQ using 2[d]=(K)2^{[d]}=\mathcal{F}(K).

In this notation, for a face F(Q)F\in\mathcal{F}(Q), 𝔯F\mathfrak{r}_{F} defined in Section 5.1 by regarding FF as a subset of [d][d], is a region generated by the positive half spaces containing FF as a face of QQ. This labeling is induced by the observation that the poset of regions 𝔯(𝒜)\mathfrak{r}(\mathcal{A}) is equal to the face lattice (K)\mathcal{F}(K). This also agrees with the labeling of 𝔯F\mathfrak{r}_{F} in Section 5.1; 𝔯F\mathfrak{r}_{F} is the region contained in the positive half spaces of i\mathscr{H}_{i} where iFi\in F. For example, if vv is the zero-dimensional face corresponding to xix_{i}, then 𝔯v=𝔯[d]{i}\mathfrak{r}_{v}=\mathfrak{r}_{[d]\smallsetminus\{i\}}.

Let JΔJ_{\Delta} be a radical monomial ideal of 𝕜[Q]𝕜[x]/I\Bbbk[Q]\cong\Bbbk[x]/I corresponding to a proper subcomplex Δ\Delta of KK and a proper face F(Q)F\in\mathcal{F}(Q). In this setting, we always label all minimal open sets of the original degree space deg(𝕜[Q]/JΔ)\bigcup\deg(\Bbbk[Q]/J_{\Delta}) using a pair of faces as in the following result. We denote max(Δ)\max(\Delta) the collection of facets of Δ\Delta.

Lemma 5.3.

For every minimal open set S𝒢(𝕜[Q]/I)S\in\mathcal{G}(\Bbbk[Q]/I) of the degree space deg(𝕜[Q]/JΔ)\bigcup\deg(\Bbbk[Q]/J_{\Delta}) there exists a unique pair of faces (F,G)(F,G) such that S(QF)𝔯FS\subseteq(Q-\mathbb{N}F)\cap\mathfrak{r}_{F} and G{Gmax(Δ)GGF}G\in\{G^{\prime}\in\bigcap\max(\Delta)\mid G^{\prime}\supseteq G\supseteq F\}. We use this pair to label SS.

Proof.

Given a face F(Q)F\in\mathcal{F}(Q), let 𝒢(Q/JΔ)F\mathcal{G}(Q/J_{\Delta})_{F} be the set of minimal open sets of the degree space deg(𝕜[x]/JΔ)\bigcup\deg(\Bbbk[x]/J_{\Delta}) contained in 𝔯F\mathfrak{r}_{F}. Recall that max(Δ)F\bigcap\max(\Delta)_{F} is the collection of faces in max(Δ)\bigcap\max(\Delta) containing FF. We claim that 𝒢(Q/JΔ)F\mathcal{G}(Q/J_{\Delta})_{F} and max(Δ)F\bigcap\max(\Delta)_{F} are in bijection.

Recall that all overlap classes (for JΔJ_{\Delta}) are of the form [0,G][0,G] for some face Gmax(Δ)G\in\max(\Delta). Hence, when F=0~,F=\tilde{0}, the corresponding minimal open set inside of 𝔯Q=Q\mathfrak{r}_{\emptyset}\cap Q=Q is obtained by intersection and complement of faces. If FΔF\not\in\Delta, then the localization is zero, thus the statement is vacuously true. Suppose F0~ΔF\gneq\tilde{0}\in\Delta. The extension (JΔ)F(J_{\Delta})_{F} of the ideal on 𝕜[QF]\Bbbk[Q-\mathbb{N}F] is still radical and its overlap classes are of form [0,G(F)][0,G\cup(-F)] where [0,G][0,G] is an overlap class of 𝕜[Q]/JΔ\Bbbk[Q]/J_{\Delta} for some face GFG\supseteq F. Thus, every minimal open set in 𝒢(Q/JΔ)F\mathcal{G}(Q/J_{\Delta})_{F} is obtained by intersecting open sets of the form [0,G(F)](FG𝔯F)\bigcup[0,G\cup(-F)]\smallsetminus\left(\bigcup_{F^{\prime}\subset G}\mathfrak{r}_{F^{\prime}}\right). Thus 𝒢(Q/JΔ)F\mathcal{G}(Q/J_{\Delta})_{F} is labeled by max(Δ)F\bigcap\max(\Delta)_{F}. ∎

We are now ready to study graded pieces of local cohomology modules.

Theorem 5.4.

Let 𝐮\mathbf{u} be a degree in a minimal open set SS indexed by (F,G)(F,G) with FGF\neq G. Then, H𝔪i(𝕜[Q]/JΔ)𝐮=0H_{\mathfrak{m}}^{i}(\Bbbk[Q]/J_{\Delta})_{\mathbf{u}}=0 for all ii. If 𝐮\mathbf{u} is a degree in a minimal open set indexed by a pair of the same face (F,F)(F,F),

H𝔪i(𝕜[Q]/JΔ)𝐮H~simpidimF1((K/F)Δ)H_{\mathfrak{m}}^{i}(\Bbbk[Q]/J_{\Delta})_{\mathbf{u}}\cong\tilde{H}_{\text{simp}}^{i-\dim F-1}((K/F)\cap\Delta)

where H~simp()\tilde{H}_{\text{simp}}^{\bullet}(-) means the reduced simplicial cohomology.

When our affine semigroup ring is the polynomial ring, the above reduces to the very well known formulas for local cohomology of Stanley–Reisner rings using homology of links.

Proof of Theorem 5.4.

Pick a minimal open set SS that corresponds to (F,G)(F,G). We know that deg(S)𝕜[QH]\deg(S)\subset\Bbbk[Q-H], where H(F/G)H\in(F/G). We may assume G=i=1mGiG=\bigcap_{i=1}^{m}G_{i} for some Gimax(Δ)G_{i}\in\max(\Delta) containing GG. Then SS is contained in an overlap class of JΔJ_{\Delta} whose face is GiG_{i}. From Definition 3.3, the 𝐮\mathbf{u}-graded part of the Ishida complex is equal to the (shifted) chain complex of Δ(G1,,Gm)/F\Delta(G_{1},\cdots,G_{m})/F where Δ(G1,,Gm)\Delta(G_{1},\cdots,G_{m}) is a subcomplex of Δ\Delta such that its maximal faces are G1,,GmG_{1},\cdots,G_{m}. When FGF\neq GLemma 5.1 shows that Δ(G1,,Gm)/F\Delta(G_{1},\cdots,G_{m})/F is contractible. Otherwise, Δ(G1,,Gm)/F=Δ(K/F)\Delta(G_{1},\cdots,G_{m})/F=\Delta\cap(K/F). ∎

We remark that this theorem holds more generally, for any affine semigroup QQ whose poset of regions 𝔯(𝒜)\mathfrak{r}(\mathcal{A}) is in bijection with (Q)\mathcal{F}(Q).

Corollary 5.5.

Lemma 5.3 and Theorem 5.4 hold when QQ is an affine semigroup such that (Q)\mathcal{F}(Q) is in bijection to the poset of regions 𝔯(𝒜)\mathfrak{r}(\mathcal{A}) of the hyperplane arrangement of 0Q\mathbb{R}_{\geq 0}Q.

Proof.

The property we used in the proof of Lemma 5.3 is that for any 𝐮\mathbf{u}, there exists a unique minimal face FF such that 𝐮QF\mathbf{u}\in Q-F. This property holds if and only if 𝔯(𝒜)\mathfrak{r}(\mathcal{A}) is in bijection to (Q)\mathcal{F}(Q). ∎

xxyyzz
Figure 1. Degrees of Q=[010100111111]Q=\mathbb{N}\left[\begin{smallmatrix}0&1&0&1\\ 0&0&1&1\\ 1&1&1&1\end{smallmatrix}\right]
Q\mathbb{Z}QQF1Q-F_{1}QF2Q-F_{2}QF3Q-F_{3}QF4Q-F_{4}Q𝐚1Q-\langle\mathbf{a}_{1}\rangleQ𝐚2Q-\langle\mathbf{a}_{2}\rangleQ𝐚3Q-\langle\mathbf{a}_{3}\rangleQ𝐚4Q-\langle\mathbf{a}_{4}\rangleQQQ\mathbb{Z}Q𝔯1\mathfrak{r}_{1}𝔯2\mathfrak{r}_{2}𝔯3\mathfrak{r}_{3}𝔯4\mathfrak{r}_{4}𝔯1,4\mathfrak{r}_{1,4}𝔯1,2\mathfrak{r}_{1,2}𝔯2,3\mathfrak{r}_{2,3}𝔯3,4\mathfrak{r}_{3,4}𝔯1,3,4\mathfrak{r}_{1,3,4}𝔯1,2,4\mathfrak{r}_{1,2,4}𝔯1,2,3\mathfrak{r}_{1,2,3}𝔯2,3,4\mathfrak{r}_{2,3,4}QQ
Figure 2. Hasse diagrams of CatQ{\textbf{Cat}}_{Q} and 𝔯(𝒜)\mathfrak{r}(\mathcal{A}) of the Segre embedding
Example 5.6 (Counterexample: Segre Embedding).

We consider the affine semigroup, Q=[011000111111],Q=\mathbb{N}\left[\begin{smallmatrix}0&1&1&0\\ 0&0&1&1\\ 1&1&1&1\end{smallmatrix}\right], which is depicted in Fig. 1. Let 𝐚i\mathbf{a}_{i} be the ii-th column of [011000111111].\left[\begin{smallmatrix}0&1&1&0\\ 0&0&1&1\\ 1&1&1&1\end{smallmatrix}\right]. Also denote the facets FiF_{i} and the hyperplane arrangement 𝒜={1,2,3,4}\mathcal{A}=\{\mathscr{H}_{1},\mathscr{H}_{2},\mathscr{H}_{3},\mathscr{H}_{4}\} as below.

F1\displaystyle F_{1} :=𝐚1,𝐚2,\displaystyle:=\langle\mathbf{a}_{1},\mathbf{a}_{2}\rangle, F2\displaystyle F_{2} :=𝐚2,𝐚3,\displaystyle:=\langle\mathbf{a}_{2},\mathbf{a}_{3}\rangle, F3\displaystyle F_{3} :=𝐚3,𝐚4,\displaystyle:=\langle\mathbf{a}_{3},\mathbf{a}_{4}\rangle, F4\displaystyle F_{4} :=𝐚4,𝐚1\displaystyle:=\langle\mathbf{a}_{4},\mathbf{a}_{1}\rangle
1(+)\displaystyle\mathscr{H}_{1}^{(+)} :={y>0},\displaystyle:=\{y>0\}, 2(+)\displaystyle\mathscr{H}_{2}^{(+)} :={z>x},\displaystyle:=\{z>x\}, 3(+)\displaystyle\mathscr{H}_{3}^{(+)} :={z>y},\displaystyle:=\{z>y\}, 4(+)\displaystyle\mathscr{H}_{4}^{(+)} :={x>0}.\displaystyle:=\{x>0\}.

For any face FF, label FF by the subset of {1,2,3,4}\{1,2,3,4\} whose corresponding facet contains FF. For example, 𝐚1\langle\mathbf{a}_{1}\rangle is indexed by {1,4}\{1,4\}. Then we have the desired injection from (Q)\mathcal{F}(Q) to 𝔯(𝒜)\mathfrak{r}(\mathcal{A}) by sending a face FF to 𝔯F:=0(QF)G<F0(QG)\mathfrak{r}_{F}:=\mathbb{R}_{\geq 0}(Q-\mathbb{N}F)\setminus\bigcup_{G<F}\mathbb{R}_{\geq 0}(Q-\mathbb{N}G). This relationship is depicted in Fig. 2. Note that this is not a bijection; for example,

(0,1,0)t=(1,1,1)t(1,0,1)t=(0,1,1)t(0,0,1)t(0,1,0)^{t}=(1,1,1)^{t}-(1,0,1)^{t}=(0,1,1)^{t}-(0,0,1)^{t}

is in both Q𝐚1Q-\langle\mathbf{a}_{1}\rangle and Q𝐚2Q-\langle\mathbf{a}_{2}\rangle but not in QQ. Hence, (0,1,0)t𝔯1,2,4(0,1,0)^{t}\in\mathfrak{r}_{1,2,4}. Therefore, we may not directly apply Corollary 5.5. However, still we may apply Corollary 4.5 to calculate its local cohomology by investigating the graded parts of the generalized Ishida complex corresponding to those “hidden" regions, i.e., regions in the cokernel of the map (Q)𝔯(𝒜)\mathcal{F}(Q)\to\mathfrak{r}(\mathcal{A}).

6. Local cohomology duality for simplicial affine semigroup rings

In this section, we relate the local cohomology of an affine semigroup ring 𝕜[Q]\Bbbk[Q] supported on a radical monomial ideal JΔJ_{\Delta} to the local cohomology supported on the maximal ideal of the quotient of 𝕜[Q]/JΔ\Bbbk[Q]/J_{\Delta}. In the case of Stanley–Reisner rings, it is straightforward to determine the vanishing of such cohomologies using available formulas for (Hilbert series of) local cohomology due to Hochster and to Terai [Terai99].

Throughout this section, we use the same notation as in Section 5.

6.1. Separated polytope and Alexander duality

First of all, we rigorously introduce the notion of cutting a face of a polytope. Suppose 𝒫\mathcal{P} is a polytope embedded in d\mathbb{R}^{d}. Let FF be a face of 𝒫\mathcal{P} and let 𝐜d\mathbf{c}\in\mathbb{R}^{d} be an outer normal vector for a supporting hyperplane F\mathscr{H}_{F} of FF. We denote this situation F=face𝐜(𝒫)F=\operatorname{face}_{\mathbf{c}}(\mathcal{P}). Let F\mathscr{H}_{F}^{\prime} be a translation of F\mathscr{H}_{F}^{\prime} in the direction of 𝐜-\mathbf{c} which separates vertices of FF and all other vertices of 𝒫\mathcal{P}. The separated polytope 𝒫F\mathcal{P}\smallsetminus F is the polytope defined as the intersection of 𝒫\mathcal{P} and the outer half space of F\mathscr{H}_{F}^{\prime}. For example, if FF is a vertex, then the separated polytope 𝒫\F\mathcal{P}\backslash F is combinatorially equivalent to the vertex figure 𝒫/F\mathcal{P}/F. We remark that (PF)\mathcal{F}(P\smallsetminus F) does not depend on the choice of F\mathscr{H}_{F}^{\prime}. We also refer to the construction of PFP\smallsetminus F as cutting the face FF from PP.

We recall the statements for both combinatorial and topological Alexander duality below for reference. Note that all simplicial, CW, and singular (co)homology in this article is reduced and with coefficients over \mathbb{Z}.

Theorem 6.1 (Alexander duality).

For any compact locally contractible nonempty proper topological subspace KK of a (d+1)(d+1)-dimensional sphere SdS^{d}, H~i(SdK)H~di1(K)\tilde{H}_{i}(S^{d}\smallsetminus K)\cong\tilde{H}^{d-i-1}(K). For any polyhedral subcomplex Δ\Delta of the boundary of a polytope PP, H~i(Δ)H~di3(Δ)\tilde{H}_{i}(\Delta)\cong\tilde{H}^{d-i-3}(\Delta^{\ast}) where Δ:=((P)Δ)op\Delta^{\ast}:=(\mathcal{F}(P)\smallsetminus\Delta)^{\text{op}} is the Alexander dual of Δ\Delta, which is a subcomplex of the dual polytope PopP^{\text{op}} of PP.

Proof.

For the topological Alexander duality, refer to [MR1867354]*Theorem 3.44. For the combinatorial Alexander duality, refer to [MR2556456, MR1119198]. ∎

For a given polytope PP and a proper polyhedral subcomplex Δ\Delta, the abstract dual polyhedral complex is the set Δ,op:=((P)Δ)op\Delta^{\ast,\text{op}}:=(\mathcal{F}(P)\smallsetminus\Delta)^{\text{op}} with the partial order reversed. We use this name because Δ\Delta^{\ast} is an abstract polyhedral complex [Ziegler95]*Corollary 2.14. Let 𝒞~(P)\tilde{\mathcal{C}}\left(P\right) (resp. 𝒞~(Δ)\tilde{\mathcal{C}}\left(\Delta\right)) be the reduced cochain complex of PP for the dual polytope. Let 𝒞~(Δ,op)unmoved\tilde{\mathcal{C}}\left(\Delta^{\ast,\text{op}}\right)_{\text{unmoved}} be a chain complex which deletes components (and componentwise maps) corresponding to faces not in Δ,op\Delta^{\ast,\text{op}} from 𝒞~(P)\tilde{\mathcal{C}}\left(P\right). We call 𝒞~(Δ,op)unmoved\tilde{\mathcal{C}}\left(\Delta^{\ast,\text{op}}\right)_{\text{unmoved}} the unmoved Alexander dual chain complex of Δ,op\Delta^{\ast,\text{op}}. This is still a well-defined chain complex, since it coincides with a reduced chain complex of the Alexander dual of Δ\Delta (with reversed indices) for homology or cohomology. Moreover, for any ii\in\mathbb{Z},

Corollary 6.2.

Hi(𝒞~(Δ))Hi+1(𝒞~(Δ,op)unmoved)H_{i}(\tilde{\mathcal{C}}\left(\Delta\right))\cong H_{i+1}(\tilde{\mathcal{C}}\left(\Delta^{\ast,\text{op}}\right)_{\text{unmoved}}) and Hi(𝒞~(Δ))Hi+1(𝒞~(Δ,op)unmoved)H^{i}(\tilde{\mathcal{C}}\left(\Delta\right))\cong H^{i+1}(\tilde{\mathcal{C}}\left(\Delta^{\ast,\text{op}}\right)_{\text{unmoved}}).

Proof.

From Theorem 6.1, observe that Hdi3(𝒞~(Δ))Hi+1(𝒞~(Δ,op)unmoved)H^{d-i-3}(\tilde{\mathcal{C}}\left(\Delta^{\ast}\right))\cong H_{i+1}(\tilde{\mathcal{C}}\left(\Delta^{\ast,\text{op}}\right)_{\text{unmoved}}) by comparing degrees of their chain complexes. The cohomology case is similar. ∎

6.2. Duality of graded local cohomologies

Let 𝕜[Q]\Bbbk[Q] be an affine semigroup ring of dimension d:=dim𝕜[Q]d:=\dim\Bbbk[Q] as defined in Section 5. Suppose that QQ has no hidden regions, i.e., the hyperplane arrangment 𝒜\mathcal{A} consisting of minimal supporting hyperplanes of 0Q\mathbb{R}_{\geq 0}Q has poset of regions 𝔯(𝒜)\mathfrak{r}(\mathcal{A}) canonically bijective to (Q)\mathcal{F}(Q) [Edelman84]*Lemma 1.3. For example, this is the case when 0Q\mathbb{R}_{\geq 0}{Q} is a cone over a simplex. Let JΔJ_{\Delta} be a monomial radical ideal. Let KK be the transverse section of Q\mathbb{R}Q with its index sets defined in Section 5. Then, there is a duality between the local cohomology of 𝕜[Q]/JΔ\Bbbk[Q]/J_{\Delta} with the maximal ideal 𝔪\mathfrak{m} support and the local cohomology of 𝕜[Q]\Bbbk[Q] with JΔJ_{\Delta}-support as follow.

Theorem 6.3.

Given a face F(Q)F\in\mathcal{F}(Q), let SFS_{F} be the minimal open set in the original degree space indexed by (F,F)(F,F) from Lemma 5.3 if it exists. Otherwise, let SFS_{F} be 𝔯F(deg(𝕜[A]/JΔ))\mathfrak{r}_{F}\cap\left(\bigcup\deg(\Bbbk[\mathbb{N}A]/J_{\Delta})\right). For any 𝐮SF\mathbf{u}\in S_{F} and 𝐯𝔯FcQ\mathbf{v}\in\mathfrak{r}_{F^{c}}\cap\mathbb{Z}Q,

(H𝔪i(𝕜[Q]/JΔ))𝐮(HJΔdi(𝕜[Q]))𝐯(H_{\mathfrak{m}}^{i}(\Bbbk[Q]/J_{\Delta}))_{\mathbf{u}}\cong(H_{J_{\Delta}}^{d-i}(\Bbbk[Q]))_{\mathbf{v}}

Some duality still holds in the case when hidden regions exist (only degrees outside of hidden regions are involved), but the statement is complicated and not very enlightening. The result is false for degrees corresponding to hidden regions, as seen in the following example.

Example 6.4 (Continuation of Example 5.6).

Let JΔ=x[012],x[112],x[212]J_{\Delta}=\langle x^{\left[\begin{smallmatrix}0\\ 1\\ 2\end{smallmatrix}\right]},x^{\left[\begin{smallmatrix}1\\ 1\\ 2\end{smallmatrix}\right]},x^{\left[\begin{smallmatrix}2\\ 1\\ 2\end{smallmatrix}\right]}\rangle. It is a radical monomial ideal such that whose corresponding subcomplex Δ\Delta has F1F_{1} and F3F_{3} as facets. Then, for the grade (0,1,0)t(0,1,0)^{t}, notes that x(0,1,0)tJΔ𝕜[Q𝐚1]JΔ𝕜[Q𝐚2]x^{(0,1,0)^{t}}\in J_{\Delta}\cdot\Bbbk[Q-\mathbb{N}\langle\mathbf{a}_{1}\rangle]\cap J_{\Delta}\cdot\Bbbk[Q-\mathbb{N}\langle\mathbf{a}_{2}\rangle] and x(0,1,0)t𝕜[Q𝐚3]𝕜[Q𝐚4]𝕜[QF3]x^{(0,1,0)^{t}}\not\in\Bbbk[Q-\mathbb{N}\langle\mathbf{a}_{3}\rangle]\cup\Bbbk[Q-\mathbb{N}\langle\mathbf{a}_{4}\rangle]\cup\Bbbk[Q-\mathbb{N}F_{3}]. This shows that the graded piece of L𝔪𝕜[Q]𝕜[Q]/JΔL^{\bullet}_{\mathfrak{m}}\mathbin{\mathop{\otimes}\displaylimits_{\Bbbk[Q]}}\Bbbk[Q]/J_{\Delta}, the Ishida complex of 𝕜[Q]/JΔ\Bbbk[Q]/J_{\Delta} supported at the monomial maximal ideal 𝔪\mathfrak{m}, is

L𝔪𝕜[Q]𝕜[Q]/JΔ:00000\displaystyle L^{\bullet}_{\mathfrak{m}}\mathbin{\mathop{\otimes}\displaylimits_{\Bbbk[Q]}}\Bbbk[Q]/J_{\Delta}:0\to 0\to 0\to 0\to 0

On the other hand, the transverse section of 𝕜[Q]JΔ\Bbbk[Q]J_{\Delta} is also a rectangle whose vertices are embedded into F2F_{2} or F4F_{4} respectively. Because (0,1,0)t𝔯1,2,4(0,1,0)^{t}\in\mathfrak{r}_{1,2,4}, x(0,1,0)t𝕜[QFi]x^{(0,1,0)^{t}}\in\Bbbk[Q-\mathbb{N}F_{i}] when i=1,2,4i=1,2,4 only. The graded piece of LJΔ𝕜[Q]𝕜[Q]L^{\bullet}_{J_{\Delta}}\mathbin{\mathop{\otimes}\displaylimits_{\Bbbk[Q]}}\Bbbk[Q], the Ishida complex of 𝕜[Q]\Bbbk[Q] supported at JΔJ_{\Delta}, is therefore

LJΔ𝕜[Q]𝕜[Q]:0𝕜4𝕜3𝕜0\displaystyle L^{\bullet}_{J_{\Delta}}\mathbin{\mathop{\otimes}\displaylimits_{\Bbbk[Q]}}\Bbbk[Q]:0\to\Bbbk^{4}\to\Bbbk^{3}\to\Bbbk\to 0

Consequently,

(HJΔj(𝕜[Q]))deg=(0,1,0)t={𝕜j=1,20o.w.,(H𝔪j(𝕜[Q]/JΔ))deg=(0,1,0)t=0 for all j.\left(H_{J_{\Delta}}^{j}(\Bbbk[Q])\right)_{\deg=(0,1,0)^{t}}=\begin{cases}\Bbbk&j=1,2\\ 0&o.w.\end{cases},\quad\left(H_{\mathfrak{m}}^{j}(\Bbbk[Q]/J_{\Delta})\right)_{\deg=(0,1,0)^{t}}=0\text{ for all }j.

This contradicts the duality when the affine semigroup contains hidden regions.

As a corollary of Theorem 6.3,

Corollary 6.5.

If QQ is a simplicial affine semigroup, then Theorem 6.3 holds.

Note that Fc:={i[n]iF}F^{c}:=\{i\in[n]\mid i\not\in F\} is a well-defined face since it either exists on both (Q)\mathcal{F}(Q) and 𝔯(𝒜)\mathfrak{r}(\mathcal{A}) simultaneously or on neither.

To proceed, assume that JΔJ_{\Delta} is neither 0 nor 𝕜[Q]\Bbbk[Q]. 𝒢(𝕜[Q])\mathcal{G}(\Bbbk[Q]) consists of 𝔯FQ\mathfrak{r}_{F}\cap\mathbb{Z}Q for each face F(Q)F\in\mathcal{F}(Q). Given the transverse section KJΔK_{J_{\Delta}} of 0JΔ\mathbb{R}_{\geq 0}J_{\Delta}, let (KJΔ)𝔯FcQ={FKJΔF^Fc}(K_{J_{\Delta}})_{\mathfrak{r}_{F^{c}}\cap\mathbb{Z}Q}=\{F\in K_{J_{\Delta}}\mid\widehat{F}\supset F^{c}\} be the set of faces of KJΔK_{J_{\Delta}} whose corresponding faces in (Q)\mathcal{F}(Q) contain FcF^{c}. Then, PFc:=(KJΔ)(KJΔ)𝔯FcQP_{F^{c}}:=\mathcal{F}(K_{J_{\Delta}})\smallsetminus(K_{J_{\Delta}})_{\mathfrak{r}_{F^{c}}\cap\mathbb{Z}Q} form a polyhedral complex since GG(KJΔ)𝔯FcQG\supseteq G^{\prime}\in(K_{J_{\Delta}})_{\mathfrak{r}_{F^{c}}\cap\mathbb{Z}Q} implies G(KJΔ)𝔯FcQG\in(K_{J_{\Delta}})_{\mathfrak{r}_{F^{c}}\cap\mathbb{Z}Q}. Consequently, the graded part of the local cohomology of 𝕜[Q]\Bbbk[Q] with JΔJ_{\Delta}-support is determined as below.

Lemma 6.6.

For a degree 𝐮𝔯FcQ\mathbf{u}\in\mathfrak{r}_{F^{c}}\cap\mathbb{Z}Q,

HJΔi(𝕜[Q])𝐮(HIshidai1(𝒞~(PFc)unmoved))(H~CWi2(PFc)).H_{J_{\Delta}}^{i}(\Bbbk[Q])_{\mathbf{u}}\cong(H_{\text{Ishida}}^{i-1}(\tilde{\mathcal{C}}\left(P_{F^{c}}\right)_{\text{unmoved}}))\cong(\tilde{H}_{\text{CW}}^{i-2}(P_{F^{c}})).
Proof.

The 𝐮\mathbf{u}-graded part of the Ishida complex with JΔJ_{\Delta}-support consists of components whose localizations are by faces in (KJΔ)𝔯FcQ(K_{J_{\Delta}})_{\mathfrak{r}_{F^{c}}\cap\mathbb{Z}Q}. Hence, the 𝐮\mathbf{u}-graded part of the Ishida complex is equal to the unmoved Alexander dual chain complex of PFcP_{F^{c}}. Therefore, the first isomorphism is from Corollary 6.2. The second isomorphism is from the difference between homological degrees of PFcP_{F^{c}} at the Ishida complex and those of PFcP_{F^{c}} at the CW chain complex. ∎

By construction, PFcP_{F^{c}} is a polyhedral complex consisting of faces of KJΔK_{J_{\Delta}} whose corresponding faces of KK induce localizations not containing the region 𝔯FcQ\mathfrak{r}_{F^{c}}\cap\mathbb{Z}Q. From a topological viewpoint, PFcP_{F^{c}} is obtained in two steps; first, cutting out all faces of max(Δ)\max(\Delta) from KK to yield KJΔK_{J_{\Delta}}. Next, cut out all faces whose corresponding faces of KK inducing localizations containing 𝔯Fc\mathfrak{r}_{F^{c}} from KJΔK_{J_{\Delta}} to get PFcP_{F^{c}}. Recall that cut is defined rigorously in Section 5.2. We claim that interchanging these two procedures results in a topological space homotopic to |PFc||P_{F^{c}}|.

Let KFcK^{F^{c}} be a simplicial complex whose dual is (2[d]/Fc)(2^{[d]}/F^{c}) on the simplex KK. In other words, KFc:=2[d](2[d]/Fc)K^{F^{c}}:=2^{[d]}\smallsetminus(2^{[d]}/F^{c}). KFcK^{F^{c}} is the result of cutting out all faces from KK whose corresponding faces of KK induces localization containing 𝔯Fc\mathfrak{r}_{F^{c}}. Now, we cut out all the maximal faces of Δ\Delta from |KFc||K^{F^{c}}| if they exist. We claim that the union of those faces cut by this process is the closure ((K/F)Δ)¯\overline{((K/F)\cap\Delta)} defined by ((K/F)Δ)¯:=|{σ2[n]στ for some τ(K/F)Δ}|\overline{((K/F)\cap\Delta)}:=|\{\sigma\in 2^{[n]}\mid\sigma\in\tau\text{ for some }\tau\in(K/F)\cap\Delta\}|. This is because faces contained in the relative interior of KFcK^{F^{c}} as a topological space are faces containing FF. Hence, cutting a maximal face of max(Δ)\max(\Delta) not containing FF does not change the combinatorial connectedness of the KFcK^{F^{c}}. This argument proves the lemma below.

Lemma 6.7.

If FKF\neq K, |PFc||P_{F^{c}}| is homotopic to the |KFc|((K/F)Δ)¯|K^{F^{c}}|\smallsetminus\overline{((K/F)\cap\Delta)}.

Furthermore,

Lemma 6.8.

If Fmax(Δ)F\not\in\bigcap\max(\Delta), then (HJΔ(𝕜[Q]))𝐯=0(H_{J_{\Delta}}^{\bullet}(\Bbbk[Q]))_{\mathbf{v}}=0 for any 𝐯𝔯FcQ\mathbf{v}\in\mathfrak{r}_{F^{c}}\cap\mathbb{Z}Q.

Proof.

From Lemma 6.7, it suffices to show that |KFc|((K/F)Δ)¯|K^{F^{c}}|\smallsetminus\overline{((K/F)\cap\Delta)} is contractible for any Fmax(Δ)F\not\in\bigcap\max(\Delta). First, suppose that FF belongs to Δ\Delta but not to max(Δ)\bigcap\max(\Delta). Then there exists a minimal face Gmax(Δ)G\in\bigcap\max(\Delta) containing FF. Now, GG contains the boundary of KFcK^{F^{c}}, thus cutting GG from KFcK^{F^{c}} does not change its contractibility. Next, suppose that FΔF\not\in\Delta. Then, no faces of max(Δ)\bigcap\max(\Delta) contain FF, thus if a face of max(Δ)\bigcap\max(\Delta) intersects KFcK^{F^{c}}, then the intersection lies on the boundary of KFcK^{F^{c}} as a topological space. This keeps |KFc|((K/F)Δ)¯|K^{F^{c}}|\smallsetminus\overline{((K/F)\cap\Delta)} contractible. ∎

Note that in the case when F0~F\neq\tilde{0} or KK, KFcK^{F^{c}} is homeomorphic to the ball Dd2D^{d-2} since it excludes all interior elements of K=2[d]K=2^{[d]} and “punctures" the boundary of KK. If F=0~F=\tilde{0}, then KFcK^{F^{c}} is homeomorphic to a sphere Sd2S^{d-2}. If F=KF=K, then KFcK^{F^{c}} is an empty set as a (1)(-1)-dimensional polyhedral complex.

Now we are ready to show that there is a homotopic image of PFcP_{F^{c}} which is the dual of (K/F)Δ(K/F)\cap\Delta in a sphere S(d1)dimF1S^{(d-1)-\dim F-1}. Recall that (K/F)Δ(K/F)\cap\Delta as a section can be seen as a subspace of sphere S(d1)dimF1S^{(d-1)-\dim F-1} by taking vertex figures iteratively as mentioned in Section 5.2.

Lemma 6.9.

PFcP_{F^{c}} is homotopic to S(d1)dimF1(K/F)ΔS^{(d-1)-\dim F-1}\smallsetminus(K/F)\cap\Delta where (K/F)Δ(K/F)\cap\Delta as a ((d1)dimF)((d-1)-\dim F)-dimensional polyhedral complex realized in S(d1)dimF1S^{(d-1)-\dim F-1}.

Proof.

If F=0~F=\tilde{0}, then PFcP_{F^{c}} is combinatorially equivalent to the empty set as a polytope and (K/F)ΔSd2(K/F)\cap\Delta\cong S^{d-2}. Also, if F=KF=K, then Δ=K\Delta=K, thus PFc=(K){K}P_{F^{c}}=\mathcal{F}(K)\smallsetminus\{K\} and (K/F)Δ={K}(K/F)\cap\Delta=\{K\}. Hence the statement holds for these two cases.

If FF is nonempty, not maximal nor minimal in (K)\mathcal{F}(K), recall that KFcK^{F^{c}} is a simplicial complex homotopic to Dd2D^{d-2} having FF in its relative interior. We claim that KFc((K/F)Δ)¯K^{F^{c}}\smallsetminus\overline{((K/F)\cap\Delta)} is homotopic to its image on Sd3S^{d-3}. To see this, pick a vertex vv in FF and take a sphere Sd3S^{d-3} centered at vv but not containing any other vertices. By translation, assume vv is the origin of d2\mathbb{R}^{d-2} embedded in PFcP_{F^{c}}. This induces a canonical homotopy map from punctured d2\mathbb{R}^{d-2} to the sphere Sd3S^{d-3} restricted to KFc((K/F)Δ)¯K^{F^{c}}\smallsetminus\overline{((K/F)\cap\Delta)} giving the desired homotopy. Lastly, use Lemma 6.7 and Lemma 5.2 to conclude that taking vertex figure on KFcK^{F^{c}} preserves its image over a polyhedral complex homeomorphic to Dd2D^{d-2} if dimF1\dim F\geq 1, or to Sd3S^{d-3} if dimF=0\dim F=0. Iterate this for the other vertices in FF to complete the argument. ∎

Corollary 6.10.

For any ii\in\mathbb{Z}, H~CWi(PFc)H~simp((d1)dimF1)i1((K/F)Δ)\tilde{H}_{\text{CW}}^{i}(P_{F^{c}})\cong\tilde{H}_{\text{simp}}^{((d-1)-\dim F-1)-i-1}((K/F)\cap\Delta).

Proof.

Lemma 6.9 shows H~CWi(PFc)H~CWi(S(d1)dimF1(K/F)Δ)\tilde{H}_{\text{CW}}^{i}(P_{F^{c}})\cong\tilde{H}_{\text{CW}}^{i}(S^{(d-1)-\dim F-1}\smallsetminus(K/F)\cap\Delta). Hence, from the topological Alexander duality and the isomorphism between simplicial homology and cohomology,

H~CWi(PFc)\displaystyle\tilde{H}_{\text{CW}}^{i}(P_{F^{c}}) H~CW,((d1)dimF1)i1((K/F)Δ)H~simp,((d1)dimF1)i1((K/F)Δ)\displaystyle\cong\tilde{H}_{\text{CW},((d-1)-\dim F-1)-i-1}((K/F)\cap\Delta)\cong\tilde{H}_{\text{simp},((d-1)-\dim F-1)-i-1}((K/F)\cap\Delta)
H~simp((d1)dimF1)i1((K/F)Δ).\displaystyle\cong\tilde{H}_{\text{simp}}^{((d-1)-\dim F-1)-i-1}((K/F)\cap\Delta).

Corollary 6.11.

Let 𝐮SF\mathbf{u}\in S_{F} where SF𝒢(𝕜[x]/JΔ)S_{F}\in\mathcal{G}(\Bbbk[x]/J_{\Delta}) indexed by (F,F)(F,F). Then, for any 𝐯𝔯FcQ\mathbf{v}\in\mathfrak{r}_{F^{c}}\cap\mathbb{Z}Q,

H𝔪i(𝕜[Q]/JΔ)𝐮HJΔdi(Q)𝐯.H_{\mathfrak{m}}^{i}(\Bbbk[Q]/J_{\Delta})_{\mathbf{u}}\cong H_{J_{\Delta}}^{d-i}(Q)_{\mathbf{v}}.
H𝔪i(𝕜[Q]/JΔ)𝐮\displaystyle H_{\mathfrak{m}}^{i}(\Bbbk[Q]/J_{\Delta})_{\mathbf{u}} Theorem 5.4H~simpidimF1((K/F)Δ)\displaystyle\underbrace{\cong}_{\lx@cref{creftype~refnum}{thm:non_acyclic_chains_of_local_coho_of_maximal_id}}\tilde{H}_{\text{simp}}^{i-\dim F-1}((K/F)\cap\Delta)
Corollary 6.10H~CW(d1)dimF1(idimF1)1(PFc)\displaystyle\underbrace{\cong}_{\lx@cref{creftype~refnum}{cor:duality_between_complexes}}\tilde{H}_{\text{CW}}^{(d-1)-\dim F-1-(i-\dim F-1)-1}(P_{F^{c}})
Lemma 6.6HJΔ(d1)dimF1(idimF1)1+2(𝕜[Q])𝐯HJΔdi(𝕜[Q])𝐯.\displaystyle\underbrace{\cong}_{\lx@cref{creftype~refnum}{lem:non_acyclic_chains_of_local_coho_over_radical}}H_{J_{\Delta}}^{(d-1)-\dim F-1-(i-\dim F-1)-1+2}(\Bbbk[Q])_{\mathbf{v}}\cong H_{J_{\Delta}}^{d-i}(\Bbbk[Q])_{\mathbf{v}}.

We are finally ready to prove the main result of this section.

Proof of Theorem 6.3.

If Δ=K\Delta=K, then JΔ=0J_{\Delta}=0. Hence H𝔪i(𝕜[Q]/JΔ)=0H_{\mathfrak{m}}^{i}(\Bbbk[Q]/J_{\Delta})=0 except for i=di=d, also H0i(𝕜[Q])=0H_{0}^{i}(\Bbbk[Q])=0 except for i=0i=0. Moreover, by computation, H𝔪d(𝕜[Q]/JΔ)=𝕜[𝔯deg(𝕜[x])]H_{\mathfrak{m}}^{d}(\Bbbk[Q]/J_{\Delta})=\Bbbk[\mathfrak{r}_{\emptyset}\cap\bigcup\deg(\Bbbk[x])] and H00(𝕜[Q])=𝕜[Q]=𝕜[𝔯[d]deg(𝕜[x])]H_{0}^{0}(\Bbbk[Q])=\Bbbk[Q]=\Bbbk[\mathfrak{r}_{[d]}\cap\bigcup\deg(\Bbbk[x])]. Therefore, the desired duality holds.

If Δ=0\Delta=0, then JΔ=𝔪J_{\Delta}=\mathfrak{m}. Hence, from the generalized Ishida complex H𝔪i(𝕜[Q]/𝔪)=0H_{\mathfrak{m}}^{i}(\Bbbk[Q]/\mathfrak{m})=0 except for i=0i=0; in case of i=0i=0, H𝔪i(𝕜[Q]/JΔ)=𝕜=𝕜[{0}deg(𝕜[x])]H_{\mathfrak{m}}^{i}(\Bbbk[Q]/J_{\Delta})=\Bbbk=\Bbbk[\{0\}\cap\bigcup\deg(\Bbbk[x])]. Also, H𝔪i(𝕜[Q])=0H_{\mathfrak{m}}^{i}(\Bbbk[Q])=0 except for i=di=d, and H𝔪d(𝕜[Q])=𝕜[deg(𝕜[x])]H_{\mathfrak{m}}^{d}(\Bbbk[Q])=\Bbbk[\mathfrak{R}_{\emptyset}\cap\bigcup\deg(\Bbbk[x])]. Thus, the duality holds for this case.

For all other cases, Δ\Delta is a proper subcomplex of KKTheorem 5.4 shows that minimal open sets with index (F,F)(F,F) for all Fmax(Δ)F\in\bigcap\max(\Delta) may have nonzero local cohomology. Also, Lemma 6.8 shows that minimal open sets FcQ\mathfrak{R}_{F^{c}}\cap\mathbb{Z}Q for any Fmax(Δ)F\in\bigcap\max(\Delta) may have nonzero local cohomology. Lastly, Corollary 6.11 provides the desired local cohomology. ∎

dd Δ\Delta JΔJ_{\Delta} degdH𝔪i(𝕜[Q]/JΔ)\deg_{\mathbb{Z}^{d}}H_{\mathfrak{m}}^{i}(\Bbbk[Q]/J_{\Delta}) degdHJΔi(𝕜[Q])\deg_{\mathbb{Z}^{d}}H_{J_{\Delta}}^{i}(\Bbbk[Q])
1 1-sim 0 (,𝔯)(\emptyset,\mathfrak{r}_{\emptyset}) (𝔯1,)(\mathfrak{r}_{1},\emptyset)
1 \emptyset 𝔪\mathfrak{m} ({0},)(\{0\},\emptyset) (,𝔯)(\emptyset,\mathfrak{r}_{\emptyset})
2 2-sim 0 (,,𝔯)(\emptyset,\emptyset,\mathfrak{r}_{\emptyset}) (𝔯1,2,,)(\mathfrak{r}_{1,2},\emptyset,\emptyset)
2 pt x\langle x\rangle (,𝔯2,)(\emptyset,\mathfrak{r}_{2},\emptyset) (,𝔯1,)(\emptyset,\mathfrak{r}_{1},\emptyset)
2 pt y\langle y\rangle (,𝔯1,)(\emptyset,\mathfrak{r}_{1},\emptyset) (,𝔯2,)(\emptyset,\mathfrak{r}_{2},\emptyset)
2 2 pts xy\langle xy\rangle (,𝔯1𝔯2{0},)(\emptyset,\mathfrak{r}_{1}\cup\mathfrak{r}_{2}\cup\{0\},\emptyset) (,𝔯1𝔯2𝔯,)(\emptyset,\mathfrak{r}_{1}\cup\mathfrak{r}_{2}\cup\mathfrak{r}_{\emptyset},\emptyset)
2 \emptyset 𝔪\mathfrak{m} ({0},,)(\{0\},\emptyset,\emptyset) (,,𝔯)(\emptyset,\emptyset,\mathfrak{r}_{\emptyset})
3 3-sim 0 (,,,𝔯)(\emptyset,\emptyset,\emptyset,\mathfrak{r}_{\emptyset}) (𝔯1,2,3,,,)(\mathfrak{r}_{1,2,3},\emptyset,\emptyset,\emptyset)
3 2-sim(yz) x\langle x\rangle (,,𝔯1,)(\emptyset,\emptyset,\mathfrak{r}_{1},\emptyset) (,𝔯2,3,,)(\emptyset,\mathfrak{r}_{2,3},\emptyset,\emptyset)
3 2-sim(xz) y\langle y\rangle (,,𝔯2,)(\emptyset,\emptyset,\mathfrak{r}_{2},\emptyset) (,𝔯1,3,,)(\emptyset,\mathfrak{r}_{1,3},\emptyset,\emptyset)
3 2-sim(xy) z\langle z\rangle (,,𝔯3,)(\emptyset,\emptyset,\mathfrak{r}_{3},\emptyset) (,𝔯1,2,,)(\emptyset,\mathfrak{r}_{1,2},\emptyset,\emptyset)
3 (xz,yz) xy\langle xy\rangle (,,𝔯1𝔯2𝔯1,2,)(\emptyset,\emptyset,\mathfrak{r}_{1}\cup\mathfrak{r}_{2}\cup\mathfrak{r}_{1,2},\emptyset) (,𝔯2,3𝔯1,3𝔯3,,)(\emptyset,\mathfrak{r}_{2,3}\cup\mathfrak{r}_{1,3}\cup\mathfrak{r}_{3},\emptyset,\emptyset)
3 (xy,yz) xz\langle xz\rangle (,,𝔯1𝔯3𝔯1,3,)(\emptyset,\emptyset,\mathfrak{r}_{1}\cup\mathfrak{r}_{3}\cup\mathfrak{r}_{1,3},\emptyset) (,𝔯2,3𝔯1,2𝔯2,,)(\emptyset,\mathfrak{r}_{2,3}\cup\mathfrak{r}_{1,2}\cup\mathfrak{r}_{2},\emptyset,\emptyset)
3 (xy,xz) yz\langle yz\rangle (,,𝔯2𝔯3𝔯2,3,)(\emptyset,\emptyset,\mathfrak{r}_{2}\cup\mathfrak{r}_{3}\cup\mathfrak{r}_{2,3},\emptyset) (,𝔯1,3𝔯1,2𝔯1,,)(\emptyset,\mathfrak{r}_{1,3}\cup\mathfrak{r}_{1,2}\cup\mathfrak{r}_{1},\emptyset,\emptyset)
3 (xy,yz,xz) xyz\langle xyz\rangle (,,(i={1},{2},{3},{1,2},{1,3},{2,3}𝔯i){0}),)(\emptyset,\emptyset,(\bigcup_{\begin{subarray}{c}i=\{1\},\{2\},\{3\},\\ \{1,2\},\{1,3\},\{2,3\}\end{subarray}}\mathfrak{r}_{i})\cup\{0\}),\emptyset) (,(i{1,2,3}𝔯i),,)(\emptyset,(\bigcup_{i\neq\{1,2,3\}}\mathfrak{r}_{i}),\emptyset,\emptyset)
3 (x,yz) xy,xz\langle xy,xz\rangle (,𝔯2,3{0},𝔯1,)(\emptyset,\mathfrak{r}_{2,3}\cup\{0\},\mathfrak{r}_{1},\emptyset) (,𝔯2,3,𝔯1𝔯,)(\emptyset,\mathfrak{r}_{2,3},\mathfrak{r}_{1}\cup\mathfrak{r}_{\emptyset},\emptyset)
3 (y,xz) xy,yz\langle xy,yz\rangle (,𝔯1,3{0},𝔯2,)(\emptyset,\mathfrak{r}_{1,3}\cup\{0\},\mathfrak{r}_{2},\emptyset) (,𝔯1,3,𝔯2𝔯,)(\emptyset,\mathfrak{r}_{1,3},\mathfrak{r}_{2}\cup\mathfrak{r}_{\emptyset},\emptyset)
3 (z,xy) xz,yz\langle xz,yz\rangle (,𝔯1,2{0},𝔯3,)(\emptyset,\mathfrak{r}_{1,2}\cup\{0\},\mathfrak{r}_{3},\emptyset) (,𝔯1,2,𝔯3𝔯,)(\emptyset,\mathfrak{r}_{1,2},\mathfrak{r}_{3}\cup\mathfrak{r}_{\emptyset},\emptyset)
3 (y,z) x,yz\langle x,yz\rangle (,𝔯2,3𝔯1,2{0},,)(\emptyset,\mathfrak{r}_{2,3}\cup\mathfrak{r}_{1,2}\cup\{0\},\emptyset,\emptyset) (,,𝔯1𝔯3𝔯,)(\emptyset,\emptyset,\mathfrak{r}_{1}\cup\mathfrak{r}_{3}\cup\mathfrak{r}_{\emptyset},\emptyset)
3 (x,z) y,xz\langle y,xz\rangle (,𝔯1,3𝔯1,2{0},,)(\emptyset,\mathfrak{r}_{1,3}\cup\mathfrak{r}_{1,2}\cup\{0\},\emptyset,\emptyset) (,,𝔯2𝔯3𝔯,)(\emptyset,\emptyset,\mathfrak{r}_{2}\cup\mathfrak{r}_{3}\cup\mathfrak{r}_{\emptyset},\emptyset)
3 (x,y) z,yz\langle z,yz\rangle (,𝔯1,3𝔯2,3{0},,)(\emptyset,\mathfrak{r}_{1,3}\cup\mathfrak{r}_{2,3}\cup\{0\},\emptyset,\emptyset) (,,𝔯2𝔯1𝔯,)(\emptyset,\emptyset,\mathfrak{r}_{2}\cup\mathfrak{r}_{1}\cup\mathfrak{r}_{\emptyset},\emptyset)
3 (z) x,y\langle x,y\rangle (,𝔯1,2,,)(\emptyset,\mathfrak{r}_{1,2},\emptyset,\emptyset) (,,𝔯3,)(\emptyset,\emptyset,\mathfrak{r}_{3},\emptyset)
3 (x) y,z\langle y,z\rangle (,𝔯2,3,,)(\emptyset,\mathfrak{r}_{2,3},\emptyset,\emptyset) (,,𝔯1,)(\emptyset,\emptyset,\mathfrak{r}_{1},\emptyset)
3 (y) x,z\langle x,z\rangle (,𝔯1,3,,)(\emptyset,\mathfrak{r}_{1,3},\emptyset,\emptyset) (,,𝔯2,)(\emptyset,\emptyset,\mathfrak{r}_{2},\emptyset)
3 (x,y,z) xy,xz,yz\langle xy,xz,yz\rangle (,𝔯2,3𝔯1,3𝔯1,2{0}2,,)(\emptyset,\mathfrak{r}_{2,3}\cup\mathfrak{r}_{1,3}\cup\mathfrak{r}_{1,2}\cup\{0\}^{2},\emptyset,\emptyset) (,,𝔯1𝔯2𝔯3𝔯2,)(\emptyset,\emptyset,\mathfrak{r}_{1}\cup\mathfrak{r}_{2}\cup\mathfrak{r}_{3}\cup\mathfrak{r}_{\emptyset}^{2},\emptyset)
3 \emptyset 𝔪\mathfrak{m} ({0},,,)(\{0\},\emptyset,\emptyset,\emptyset) (,,,𝔯)(\emptyset,\emptyset,\emptyset,\mathfrak{r}_{\emptyset})
Figure 3. Table of local cohomologies over a simplicial affine semigroup ring 𝕜[Q]\Bbbk[Q] when whose dimension dd is 1,2, or 3.
Example 6.12.

In Fig. 3, we summarize the degrees of local cohomology over a simplicial affine semigroup ring 𝕜[Q]\Bbbk[Q] (with dim𝕜[Q]=1,2,\dim\Bbbk[Q]=1,2, or 33) and with a radical monomial ideal JΔJ_{\Delta} from a simplicial complex generated by {x}\{x\} (resp. {x,y}\{x,y\} or {x,y,z}\{x,y,z\}) corresponding to the variables of 𝕜[Q]\Bbbk[Q].

For example, the 12th row of the table illustrates that, when 𝕜[Q]\Bbbk[Q] is a 3-dimensional simplicial affine semigroup ring, and Δ\Delta is a simplicial complex consisting of two edges xy¯\overline{xy} and yz¯\overline{yz}, then its corresponding radical monomial ideal is xy\langle xy\rangle, thus for any xu¯𝕜[Q]/JΔ\overline{x^{\vec{u}}}\in\Bbbk[Q]/J_{\Delta} with u𝔯1𝔯2𝔯1,2\vec{u}\in\mathfrak{r}_{1}\cup\mathfrak{r}_{2}\cup\mathfrak{r}_{1,2} and xv𝕜[Q]x^{\vec{v}}\in\Bbbk[Q] with v𝔯2,3𝔯1,3𝔯3,\vec{v}\in\mathfrak{r}_{2,3}\cup\mathfrak{r}_{1,3}\cup\mathfrak{r}_{3},

H𝔪2(𝕜[Q]/JΔ)uHJΔ32(𝕜[Q]/JΔ)vH_{\mathfrak{m}}^{2}(\Bbbk[Q]/J_{\Delta})_{\vec{u}}\cong H_{J_{\Delta}}^{3-2}(\Bbbk[Q]/J_{\Delta})_{\vec{v}}

and for all other u\vec{u} and v\vec{v},

H𝔪i(𝕜[Q]/JΔ)u=HJΔi(𝕜[Q]/JΔ)v=0.H_{\mathfrak{m}}^{i}(\Bbbk[Q]/J_{\Delta})_{\vec{u}}=H_{J_{\Delta}}^{i}(\Bbbk[Q]/J_{\Delta})_{\vec{v}}=0.

References