A Generalization of the Ishida Complex with applications
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 . Given a lattice , its lattice ideal is
where is a field. The saturation of a lattice is . 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 is known to be a minimal prime of , and if is algebraically closed, all associated primes of are isomorphic to 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 and [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 is 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 is a simplicial complex, is the associated squarefree monomial ideal, and 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)
(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 is a finitely generated submonoid of . Throughout this article, we denote by a integer matrix of rank whose columns generate , so that . We assume none of the columns of is the zero vector. If is a field and is the affine semigroup given by , we denote by the corresponding affine semigroup ring. Since has rank , is a -dimensional -algebra.
The cone over the affine semigroup (or over ) is the (rational, polyhedral) cone , that is, the set of all non-negative real combinations of columns of . This cone is pointed if it contains no lines. In this case, the affine semigroup is also called pointed. A face of is a subset of this cone where some linear functional on is maximized over . We denote by the collection of all faces of . This set forms a lattice under inclusion. For ease in the notation, we identify a face of with the set of columns of that lie on that face. For a face , the relative interior of the semigroup over is the set of elements of that do not belong to any subsemigroups arising from proper faces of . A transverse section of is the intersection of with a hyperplane which meets all unbounded faces of [Ziegler95]*Exercise 2.19. It is well-known that is canonically bijective to as a poset.
Let be the set of all monomial prime ideals of the affine semigroup ring , ordered by inclusion. There is an order-reversing isomorphism between and the face lattice of the transverse section of given by identifying a face of with the (prime) ideal generated by all monomials whose exponents do not belong to the corresponding face of .
It is known that the set of all radical ideals is in one to one correspondence with the set of all polyhedral subcomplexes of .
2.2. Binomial ideals
Let be the polynomial ring over a field . 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 be a subgroup of . A partial character on is a group homomorphism, where is the multiplicative group of . The lattice ideal corresponding to is the ideal in defined as
(2) |
We remark that in [MR1394747], these lattice ideals are denoted by , while 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 of a lattice is . A lattice is saturated if . If is algebraically closed, then is prime if and only if 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 is an algebraically closed field of characteristic zero, then lattice ideals are radical, and primary lattice ideals are prime.
An ideal is cellular if all variables are either nonzero divisors modulo or nilpotent modulo . The nonzero divisor variables are known as the cellular variables of . Let be the set of all indices of cellular variables of a cellular binomial ideal , then is -cellular. Lattice ideal are special cases of cellular binomial ideals, where all variables are cellular. Also, 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 be a -cellular binomial ideal in . The associated primes of are the ideals , where runs over the associated primes of lattice ideals of the form , for monomials .
A toric ideal is the prime lattice ideal corresponding to the trivial character . An affine semigroup is the quotient of by the equivalence relation given by . A quotient of polynomial ring by corresponding prime lattice ideal is isomorphic to an affine semigroup ring [CCA]*Theorem 7.3.
From now on, we assume that the base field is algebraically closed.
3. Generalized Ishida complex
Given a lattice ideal (resp. -cellular binomial ideal ), pick a minimal associated prime ideal of (resp. of ). Since is algebraically closed, is a binomial prime ideal, and the quotient is isomorphic (by rescaling the variables) to an affine semigroup ring with . The natural projection map (resp. ) induces -grading on ; in other words, for any monomial , the -degree of is (resp. , where ). We specify the -grading on using the triple unless the minimal prime or the generators of affine semigroup are understood in context.
In this section, we generalize the Ishida complex to compute the local cohomology of a lattice ideal (resp. -cellular binomial ideal) supported at a contraction of a radical monomial ideal of . Recall that the contraction of a radical monomial ideal is also a monomial ideal in .
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 with transverse section , there is a canonical isomorphism given as follows: is the minimal face of such that . Since is pointed, it has a unique zero-dimensional face, namely the origin, which corresponds to the (-1)-dimensional face of . As a CW complex, has an incidence function that has a nonzero value when two faces are incident. We now recall the definition of the Ishida complex [MR977758].
Definition 3.1.
Let be the maximal monomial ideal of . The set of all -dimensional faces in is denoted by . Let be the chain complex
where the differential is induced by the componentwise map with , such that
with , the canonical injection when . We say that is the Ishida complex of a -module supported at the maximal monomial ideal.
The cohomology of the Ishida complex of supported at the maximal monomial ideal is isomorphic to the local cohomology of supported at the maximal monomial ideal.
Theorem 3.2 ([MR977758]*Theorem 6.2.5).
For any -module , and all ,
Let be the radical monomial ideal of associated to a subcomplex . Then, denotes the contraction of via . To compute the local cohomology of a -module supported on , we construct a (generalized) Ishida complex below.
Let be a transverse section of the polyhedron with the canonical isomorphism where is the minimal face of such that . The set of all -dimensional faces in is denoted by . Also, refers to the localization of by the multiplicative set consisting of all monomials in whose -graded degrees are in .
Definition 3.3 (Generalized Ishida complex).
Let be the chain complex
where the differential is induced by a componentwise map with two faces , such that
with , the canonical injection when . We say that is the Ishida complex of a -module supported at the radical monomial ideal .
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 -module , and all ,
The first step in our proof is to verify that the zeroth homology of the generalized Ishida complex computes torsion.
Proof.
Lemma 3.5.
.
Proof.
It suffices to show that
The generators of the left hand-side ideal might be not the same as those of the multiplicative sets inducing localization of components in when is not toric. However, they admit the same radical ideal in .
First, let be an element whose -degree is in for a vertex of . The canonical map sends to where . Since , , we have that by the correspondence between polyhedral subcomplexes and the radical monomial ideals.
Conversely, let be a preimage of a monomial in and be an -homogeneous element of . Then, for some such that for some . If , is in the left hand-side of the equation. Suppose ; then has vertices . Then, is a linear combination of elements of over , say where and . Multiplying by a suitable number , we may assume that . Then, where , which implies that is in the left hand-side, thus 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 is an injective -module, then is exact.
Proof.
It suffices to check the case when is an injective indecomposable module over a prime ideal containing . Let be the -th column of . The set is called the face corresponding to . Lemma 3.7 shows that this is indeed a face of . Lemma 3.9 shows that is exact using Lemma 3.8.
∎
We start verifying our proposed face is a face.
Lemma 3.7.
Given a monomial prime ideal containing , is a face of .
Proof.
Since and the module is indecomposable, is either if or if . Now suppose that the given set is not a face; then there exists a minimal face whose relative interior intersects with the relative interior of . Pick . Then the corresponding variable induces the zero map on . If is not in the relative interior of , let so that for some . Choose a suitable such that for some nonnegative and . By construction, is in the lattice . Thus, there exists such that where , . But then , which is a contradiction. If is in the relative interior of , a similar argument gives another contradiction. ∎
The following is necessary to prove exactness.
Lemma 3.8.
Given a face , is a face of .
Proof.
If , then the statement is clear. Assume . First, we claim that for any , if and only if . One direction follows straight from the definition of . Conversely, assume that . Then, implies . Therefore . This claim shows that is the union of all faces such that . Thus can be regarded as a realized subcomplex of .
Next, we claim that has a unique maximal element. Suppose not; let and be two distinct faces of of maximal dimension. Then, implies that there is a face such that , the join of the two faces. Since and are distinct and the same dimension, or . Therefore . Hence, , which implies . Hence, , therefore , contradicting the maximality of and . We conclude has a unique maximal element, say .
Lastly, we claim that , which implies . For any , let in . Then, . By the maximality of , , which implies . ∎
Lemma 3.9.
Given a monomial prime ideal whose corresponding face is , for ,
where is the reduced chain complex of as a CW complex.
Proof.
Lemma 3.7 shows that for any ,
If is not the image of a face in , then no sub-face of is the image of a face of . Otherwise, there is a face containing an unbounded face of . Then by the correspondence between radical monomial ideals and subcomplexes of , contains an element of an unbounded face of , a contradiction. Therefore, no images of faces are subsets of . This implies that and for .
Otherwise, for some . Since if and only if by Lemma 3.8, so the first equality holds.
Now observe that as a -module. Thus,
∎
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 as a (formal) finite sum of rational functions when 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 in terms of the cohomology of finitely many chain complexes (Corollary 4.5). Let be a tuple consisting of a lattice ideal (resp. cellular binomial ideal), a minimal prime ideal of , and the corresponding affine semigroup . Then is also a prime lattice ideal and we may assume after rescaling the variables that is toric, with lattice . Let be the corresponding torsion abelian group, then
We may induce a fine grading of by as follows: for any for some , . Here we use to indicate the image of in .
Let be a monomial ideal of an affine semigroup ring . Let . A proper pair of is a pair such that . Given two pairs and , we say if . A degree pair of is a maximal element of the set of all proper pairs with the given order. Two pairs and with the same face overlap if the intersection is nonempty. Overlapping is an equivalence relation on pairs; an overlap class is an equivalence class containing the degree pair . Let be the set of all degrees belonging to for some degree pair in . The original degree space is the union of for all monomial prime ideals of . The degree pair topology is the smallest topology on such that for any overlap class of any localization , the set 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 is the hyperplane arrangement consisting of the supporting hyperplanes of the facets of . Let be the set of regions of , where a region is a connected components of [Stanley07]. Then, for any region , regard as a space with the trivial topology. The extended degree space of is the disjoint union as a topological space. (As a set, equals .) Lastly, let (resp. ) be the collection of all minimal open sets of the extended (resp. original) degree space of . Since is finite [MY2022]*Lemma 4.3 and is finite, is also finite.
Given , let be the following ideal in
For an open set , let be the graded part of the generalized Ishida complex associated to the element whose torsion degree is and whose -degree is . This is well-defined regardless of choice of , as is stated below.
Lemma 4.1.
If and with the same torsion degree are in the same minimal open set of the extended degree space of , then their corresponding graded parts of the Ishida complex coincide.
Proof.
If neither nor are in the original degree space of , they must be in the same region of the hyperplane arrangement. Hence, for any localization by a face , either both degrees are or they do not belong to the same localization.
If both are in the original degree space, let be an overlap class whose degree set contains and , and such that is minimal with this property. Then and do not appear in the localization of by a multiplicative set generated by variables corresponding to a proper face of . Conversely, if there is no overlap class containing , this means that appears on every localization of by faces containing . This completely determines the graded parts of the Ishida complex. ∎
Theorem 4.2.
Given a lattice ideal , define as before. T he multi-graded Hilbert series for the th local cohomology module of supported on the inverse image of the radical monomial ideal with respect to the -grading is
Proof.
This is a direct consequence of Lemma 4.1. ∎
Since is finite and is finite for all , the sum is finite. Moreover, each minimal open set is the set of lattice points in a convex polyhedron, so that can be written as a rational function [Barvinok99, Barvinok03].
For the case of a -cellular binomial ideal , let be the multiplicative set consisting of monomials on the nilpotent variables of . Then for each , is a lattice ideal containing . Hence, according to Theorem 2.1, we may pick an associated prime ideal of whose extension is the associated prime containing . Moreover, the quotient is isomorphic to an affine semigroup ring for some integer matrix . The canonical projection induces a monoid map . By letting be the quotient of the saturation of the lattice corresponding to by , we have a fine grading of by the abelian group
via Thus, for a fixed and a torsion , let
Using the same arguments as for lattice ideals, we know that two elements whose degrees are in the same minimal open set have the same graded part of the Ishida complex . Then, we have
Theorem 4.3.
Given a cellular binomial ideal , the multi-graded Hilbert series for the local cohomology module of supported on the image of the radical monomial ideal with respect to the -grading is
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.
is the cochain complex of a polyhedral complex.
Proof.
It suffices to show that the nontrivial top dimensional part of is ; suppose not; then there exists distinct maximal faces and of such that and for some distinct and with . Then, the degree of the product lies in the relative interior of a face in which is a minimal face containing both and , contradicting the maximality of and . ∎
Corollary 4.5.
Let be a cellular binomial ideal. Then is Cohen–Macaulay if and only if for all and for all , . ∎
Example 4.6.
Let be the following lattice in and its saturation.
The torsion group is isomorphic to . We represent as follows
using GRevLex term order in Macaulay2.
In this case . As usual . On the polynomial ring , the lattice ideals corresponding to and are
Hence, the Ishida complex supported on the maximal monomial ideal is
Here, we see that for any ,
when and
when for any . Since there is no non-top cohomology, we conclude that is Cohen-Macaulay.
5. Local cohomology for affine semigroups over simplices
In this section, we consider affine semigroup rings such that is a cone over a simplex. In this case the combinatorics simplifies and becomes more explicit.
5.1. Hyperplane arrangements
The hyperplane arrangement of a polyhedron is the collection of supporting hyperplanes of the facets of in . In this case, , where 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 of a hyperplane arrangement is a connected component of . refers to the collection of all regions of .
Suppose consists of a minimal number of hyperplanes which generate a rational polyhedral cone . Then is linear and all regions in are unbounded rational polyhedral cones. Moreover, every region can be expressed as
for a subset where is the complement of . In other words, a region is labeled by the collection of hyperplanes whose positive half space contains it. It follows that is partially ordered by reverse inclusion on the set of labels; We call the poset of regions of . Our notation here is consistent with that of [Edelman84, BEZ90] regarding as the base region. Moreover, the natural embedding sending a face to the set of indices of hyperplanes containing exists [Edelman84]*Lemma 1.3. Since we are interested in regions partitioning along with the set of degrees of standard monomials of localizations, we modify the definition of as follows, to include boundaries:
A cumulative region is the union of all regions less then . The poset of cumulative regions is a set of all cumulative regions ordered by inclusion. By definition, as posets. We follow the conventions of [Stanley07].
5.2. Sections of polyhedra
If is a zero-dimensional face of a polyhedral complex , then we define the vertex figure of at as follows. Assume is realized in , Pick a sphere centered at such that every nonempty face of except is not completely contained in the sphere. Then, the vertex figure is the polyhedral complex generated by the intersections of all faces of containing on . Likewise, when is positive dimensional face, we define the section of at , denoted , to be the polytope generated by taking vertex figures iteratively over the vertices arising from . As a new abstract polytope, is combinatorially equivalent to the link of over as a sub-polyhedral complex of . Moreover, if we regard the face lattice of the vertex figure as a collection of faces of , then its closure in the usual Euclidean topology agrees with the star of .
Given a polyhedral complex , let be the set of maximal elements of , and let be a set of faces which are intersections of maximal faces of as a poset. We say is -combinatorially connected for . Note that this is a much finer notion of connectedness than the usual topological -connectedness. For example, a simplicial complex consisting of two triangles sharing an edge is -combinatorially connected but contractible (infinitely-connected) in the sense of topological -connectedness.
Lemma 5.1.
Any vertex figure of an -combinatorially connected polyhedral complex is at least -combinatorially connected. Moreover, when , for any vertex , the vertex figure is contractible. Hence, for any face whose dimension is less than , is contractible.
Proof.
As the case holds vacuously, we may let . Pick a vertex . Let be the intersection of all maximal faces of containing . Then, All maximal faces of the vertex figure are inherited from those maximal faces of containing , hence is -combinatorially connected, and .
For the second statement, let and be the same as above. Let as a set of faces of . Since is homotopic to ; we may collapse each maximal face in containing continuously to . Moreover, is homotopic to . To see this, let be a sphere centered at and generating the vertex figure of and on its surface. The homotopy from to restricted on gives the homotopy between vertex figures of and over if . The last statement follows by applying the second statement iteratively. ∎
Lemma 5.2.
Given a -dimensional polyhedral complex homeomorphic to a disk , let be a vertex. If is in the interior of , then is homeomorphic to . Otherwise, if is in the boundary of , then is homeomorphic to .
Proof.
If is in the boundary of , the result is clear since the sphere containing that defines the vertex figure cannot be contained in the polyhedral complex. One can apply a similar argument to the case when is in the relative interior. ∎
5.3. Local cohomology
Let be a prime lattice ideal whose corresponding affine semigroup is simplicial, i.e., the transverse section of the polyhedral cone is a -simplex. Let be the minimal hyperplane arrangement of . Label the faces of by their supporting facets; for a facet , use the label . Hence, the zero-dimensional face (of the transverse section) that does not lie in the hyperplane is indexed by
From the natural isomorphism , we may label faces of using .
In this notation, for a face , defined in Section 5.1 by regarding as a subset of , is a region generated by the positive half spaces containing as a face of . This labeling is induced by the observation that the poset of regions is equal to the face lattice . This also agrees with the labeling of in Section 5.1; is the region contained in the positive half spaces of where . For example, if is the zero-dimensional face corresponding to , then .
Let be a radical monomial ideal of corresponding to a proper subcomplex of and a proper face . In this setting, we always label all minimal open sets of the original degree space using a pair of faces as in the following result. We denote the collection of facets of .
Lemma 5.3.
For every minimal open set of the degree space there exists a unique pair of faces such that and . We use this pair to label .
Proof.
Given a face , let be the set of minimal open sets of the degree space contained in . Recall that is the collection of faces in containing . We claim that and are in bijection.
Recall that all overlap classes (for ) are of the form for some face . Hence, when the corresponding minimal open set inside of is obtained by intersection and complement of faces. If , then the localization is zero, thus the statement is vacuously true. Suppose . The extension of the ideal on is still radical and its overlap classes are of form where is an overlap class of for some face . Thus, every minimal open set in is obtained by intersecting open sets of the form . Thus is labeled by . ∎
We are now ready to study graded pieces of local cohomology modules.
Theorem 5.4.
Let be a degree in a minimal open set indexed by with . Then, for all . If is a degree in a minimal open set indexed by a pair of the same face ,
where 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 that corresponds to . We know that , where . We may assume for some containing . Then is contained in an overlap class of whose face is . From Definition 3.3, the -graded part of the Ishida complex is equal to the (shifted) chain complex of where is a subcomplex of such that its maximal faces are . When , Lemma 5.1 shows that is contractible. Otherwise, . ∎
We remark that this theorem holds more generally, for any affine semigroup whose poset of regions is in bijection with .
Corollary 5.5.
Lemma 5.3 and Theorem 5.4 hold when is an affine semigroup such that is in bijection to the poset of regions of the hyperplane arrangement of .
Proof.
The property we used in the proof of Lemma 5.3 is that for any , there exists a unique minimal face such that . This property holds if and only if is in bijection to . ∎
Example 5.6 (Counterexample: Segre Embedding).
We consider the affine semigroup, which is depicted in Fig. 1. Let be the -th column of Also denote the facets and the hyperplane arrangement as below.
For any face , label by the subset of whose corresponding facet contains . For example, is indexed by . Then we have the desired injection from to by sending a face to . This relationship is depicted in Fig. 2. Note that this is not a bijection; for example,
is in both and but not in . Hence, . 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 .
6. Local cohomology duality for simplicial affine semigroup rings
In this section, we relate the local cohomology of an affine semigroup ring supported on a radical monomial ideal to the local cohomology supported on the maximal ideal of the quotient of . 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 is a polytope embedded in . Let be a face of and let be an outer normal vector for a supporting hyperplane of . We denote this situation . Let be a translation of in the direction of which separates vertices of and all other vertices of . The separated polytope is the polytope defined as the intersection of and the outer half space of . For example, if is a vertex, then the separated polytope is combinatorially equivalent to the vertex figure . We remark that does not depend on the choice of . We also refer to the construction of as cutting the face from .
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 .
Theorem 6.1 (Alexander duality).
For any compact locally contractible nonempty proper topological subspace of a -dimensional sphere , . For any polyhedral subcomplex of the boundary of a polytope , where is the Alexander dual of , which is a subcomplex of the dual polytope of .
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 and a proper polyhedral subcomplex , the abstract dual polyhedral complex is the set with the partial order reversed. We use this name because is an abstract polyhedral complex [Ziegler95]*Corollary 2.14. Let (resp. ) be the reduced cochain complex of for the dual polytope. Let be a chain complex which deletes components (and componentwise maps) corresponding to faces not in from . We call the unmoved Alexander dual chain complex of . This is still a well-defined chain complex, since it coincides with a reduced chain complex of the Alexander dual of (with reversed indices) for homology or cohomology. Moreover, for any ,
Corollary 6.2.
and .
Proof.
From Theorem 6.1, observe that by comparing degrees of their chain complexes. The cohomology case is similar. ∎
6.2. Duality of graded local cohomologies
Let be an affine semigroup ring of dimension as defined in Section 5. Suppose that has no hidden regions, i.e., the hyperplane arrangment consisting of minimal supporting hyperplanes of has poset of regions canonically bijective to [Edelman84]*Lemma 1.3. For example, this is the case when is a cone over a simplex. Let be a monomial radical ideal. Let be the transverse section of with its index sets defined in Section 5. Then, there is a duality between the local cohomology of with the maximal ideal support and the local cohomology of with -support as follow.
Theorem 6.3.
Given a face , let be the minimal open set in the original degree space indexed by from Lemma 5.3 if it exists. Otherwise, let be . For any and ,
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 . It is a radical monomial ideal such that whose corresponding subcomplex has and as facets. Then, for the grade , notes that and . This shows that the graded piece of , the Ishida complex of supported at the monomial maximal ideal , is
On the other hand, the transverse section of is also a rectangle whose vertices are embedded into or respectively. Because , when only. The graded piece of , the Ishida complex of supported at , is therefore
Consequently,
This contradicts the duality when the affine semigroup contains hidden regions.
As a corollary of Theorem 6.3,
Corollary 6.5.
If is a simplicial affine semigroup, then Theorem 6.3 holds.
Note that is a well-defined face since it either exists on both and simultaneously or on neither.
To proceed, assume that is neither nor . consists of for each face . Given the transverse section of , let be the set of faces of whose corresponding faces in contain . Then, form a polyhedral complex since implies . Consequently, the graded part of the local cohomology of with -support is determined as below.
Lemma 6.6.
For a degree ,
Proof.
The -graded part of the Ishida complex with -support consists of components whose localizations are by faces in . Hence, the -graded part of the Ishida complex is equal to the unmoved Alexander dual chain complex of . Therefore, the first isomorphism is from Corollary 6.2. The second isomorphism is from the difference between homological degrees of at the Ishida complex and those of at the CW chain complex. ∎
By construction, is a polyhedral complex consisting of faces of whose corresponding faces of induce localizations not containing the region . From a topological viewpoint, is obtained in two steps; first, cutting out all faces of from to yield . Next, cut out all faces whose corresponding faces of inducing localizations containing from to get . Recall that cut is defined rigorously in Section 5.2. We claim that interchanging these two procedures results in a topological space homotopic to .
Let be a simplicial complex whose dual is on the simplex . In other words, . is the result of cutting out all faces from whose corresponding faces of induces localization containing . Now, we cut out all the maximal faces of from if they exist. We claim that the union of those faces cut by this process is the closure defined by . This is because faces contained in the relative interior of as a topological space are faces containing . Hence, cutting a maximal face of not containing does not change the combinatorial connectedness of the . This argument proves the lemma below.
Lemma 6.7.
If , is homotopic to the .
Furthermore,
Lemma 6.8.
If , then for any .
Proof.
From Lemma 6.7, it suffices to show that is contractible for any . First, suppose that belongs to but not to . Then there exists a minimal face containing . Now, contains the boundary of , thus cutting from does not change its contractibility. Next, suppose that . Then, no faces of contain , thus if a face of intersects , then the intersection lies on the boundary of as a topological space. This keeps contractible. ∎
Note that in the case when or , is homeomorphic to the ball since it excludes all interior elements of and “punctures" the boundary of . If , then is homeomorphic to a sphere . If , then is an empty set as a -dimensional polyhedral complex.
Now we are ready to show that there is a homotopic image of which is the dual of in a sphere . Recall that as a section can be seen as a subspace of sphere by taking vertex figures iteratively as mentioned in Section 5.2.
Lemma 6.9.
is homotopic to where as a -dimensional polyhedral complex realized in .
Proof.
If , then is combinatorially equivalent to the empty set as a polytope and . Also, if , then , thus and . Hence the statement holds for these two cases.
If is nonempty, not maximal nor minimal in , recall that is a simplicial complex homotopic to having in its relative interior. We claim that is homotopic to its image on . To see this, pick a vertex in and take a sphere centered at but not containing any other vertices. By translation, assume is the origin of embedded in . This induces a canonical homotopy map from punctured to the sphere restricted to giving the desired homotopy. Lastly, use Lemma 6.7 and Lemma 5.2 to conclude that taking vertex figure on preserves its image over a polyhedral complex homeomorphic to if , or to if . Iterate this for the other vertices in to complete the argument. ∎
Corollary 6.10.
For any , .
Proof.
Lemma 6.9 shows . Hence, from the topological Alexander duality and the isomorphism between simplicial homology and cohomology,
∎
Corollary 6.11.
Let where indexed by . Then, for any ,
We are finally ready to prove the main result of this section.
Proof of Theorem 6.3.
If , then . Hence except for , also except for . Moreover, by computation, and . Therefore, the desired duality holds.
If , then . Hence, from the generalized Ishida complex except for ; in case of , . Also, except for , and . Thus, the duality holds for this case.
For all other cases, is a proper subcomplex of . Theorem 5.4 shows that minimal open sets with index for all may have nonzero local cohomology. Also, Lemma 6.8 shows that minimal open sets for any may have nonzero local cohomology. Lastly, Corollary 6.11 provides the desired local cohomology. ∎
1 | 1-sim | 0 | ||
---|---|---|---|---|
1 | ||||
2 | 2-sim | 0 | ||
2 | pt | |||
2 | pt | |||
2 | 2 pts | |||
2 | ||||
3 | 3-sim | 0 | ||
3 | 2-sim(yz) | |||
3 | 2-sim(xz) | |||
3 | 2-sim(xy) | |||
3 | (xz,yz) | |||
3 | (xy,yz) | |||
3 | (xy,xz) | |||
3 | (xy,yz,xz) | |||
3 | (x,yz) | |||
3 | (y,xz) | |||
3 | (z,xy) | |||
3 | (y,z) | |||
3 | (x,z) | |||
3 | (x,y) | |||
3 | (z) | |||
3 | (x) | |||
3 | (y) | |||
3 | (x,y,z) | |||
3 |
Example 6.12.
In Fig. 3, we summarize the degrees of local cohomology over a simplicial affine semigroup ring (with or ) and with a radical monomial ideal from a simplicial complex generated by (resp. or ) corresponding to the variables of .
For example, the 12th row of the table illustrates that, when is a 3-dimensional simplicial affine semigroup ring, and is a simplicial complex consisting of two edges and , then its corresponding radical monomial ideal is , thus for any with and with
and for all other and ,