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

Local cohomology tables of sequentially almost Cohen-Macaulay modules

Cheng Meng
Abstract.

Let RR be a polynomial ring over a field. We introduce the concept of sequentially almost Cohen-Macaulay modules and describe the extremal rays of the cone of local cohomology tables of finitely generated graded RR-modules which are sequentially almost Cohen-Macaulay, and describe some cases when the local cohomology table of a module of dimension 3 has a nontrivial decomposition.

1. introduction

Let R=k[x1,,xn]R=k[x_{1},\ldots,x_{n}] be a standard graded polynomial ring over a field kk. The graded Betti numbers and the local cohomology modules are important homological data of graded modules over RR. In 2006, Boij and Söderberg [2] formulated two conjectures on the cone of graded Betti tables of finitely generated Cohen-Macaulay modules, which were proved by David Eisenbud, Gunnar Fløystad and Jerzy Weyman in characteristic 0 in [6] and by Eisenbud and Schreyer in arbitrary characteristic in [7]. These conjectures were also extended to the non-Cohen-Macaulay case by Boij and Söderberg in [3].

Denote the Betti table of a finitely generated graded module by β()\beta^{\bullet}(\cdot), then the above results can be restated in the following way:

Theorem 1.1.

Let R=k[x1,,xn]R=k[x_{1},\ldots,x_{n}] be a standard graded polynomial ring. The extremal rays of the cone generated by Betti tables of finitely generated graded RR-modules are given by the modules with a pure resolution, and every Betti table in the cone decomposes in the following way. For every finitely generated graded module MM there exist finitely generated modules N1,,NsN_{1},\ldots,N_{s} with pure resolutions and r1,,rsr_{1},\ldots,r_{s}\in\mathbb{Q}, r1,,rs>0r_{1},\ldots,r_{s}>0 such that

β(M)=i=1sriβ(Ni).\beta^{\bullet}(M)=\sum_{i=1}^{s}r_{i}\beta^{\bullet}(N_{i}).

Eisenbud and Schreyer, in [7], asked for a similar description for the cone of cohomology tables of coherent sheaves, and they proved a result similar to Theorem 1.1 in [8]:

Theorem 1.2.

Let R=k[x1,,xn]R=k[x_{1},\ldots,x_{n}] be a standard graded polynomial ring, X=Proj(R)X=\textup{Proj}(R). The extremal rays of the cone generated by cohomology tables of coherent sheaves are given by those of supernatural vector bundles, and for every coherent sheaf \mathcal{F} there exist possibly infinitely many supernatural vector bundles 1,2,\mathcal{F}_{1},\mathcal{F}_{2},\ldots and r1,r2,r_{1},r_{2},\ldots\in\mathbb{Q}, ri>0r_{i}>0 such that

H(X,)=i1riH(X,i).H^{\bullet}(X,\mathcal{F})=\sum_{i\geq 1}r_{i}H^{\bullet}(X,\mathcal{F}_{i}).

In this paper, we study the problem of writing the local cohomology tables of modules as a finite sum. There are two obstacles between this problem and Eisenbud and Schreyer’s result. The first problem is about finiteness. The second one is that the local cohomology tables of modules and cohomology tables of coherent sheaves differ at the beginning. For instance, let X=Proj(R)X=\textup{Proj}(R), MM be a finitely generated graded RR-module, 𝔪=(x1,,xn)\mathfrak{m}=(x_{1},\ldots,x_{n}) be the graded maximal ideal, =M~\mathcal{F}=\tilde{M} be the corresponding coherent sheaf. Let Γ(M)=tH0(X,(t))\Gamma(M)=\oplus_{t\in\mathbb{Z}}H^{0}(X,\mathcal{F}(t)) be the module of global sections of M. Then we know that Hi(X,)=H𝔪i+1(M)H^{i}(X,\mathcal{F})=H^{i+1}_{\mathfrak{m}}(M) for i1i\geq 1 and there is an exact sequence

0H𝔪0(M)MΓ(M)H𝔪1(M)0.0\to H^{0}_{\mathfrak{m}}(M)\to M\to\Gamma(M)\to H^{1}_{\mathfrak{m}}(M)\to 0.

The sheaf \mathcal{F} only determines Γ(M)\Gamma(M), and in general it is hard to decompose H𝔪1(M)H^{1}_{\mathfrak{m}}(M) using the data of Γ(M)\Gamma(M).

This problem is still open in general, and up to now, the best result is proved by Smirnov and De Stefani in [5], where they gave a complete description of the cone of local cohomology tables of modules of dimension at most 2. By denoting the local cohomology table of a graded module by H()H^{\bullet}(\cdot), Smirnov and De Stefani’s results are as follows:

Theorem 1.3 ([5], Theorem 4.6).

Let R=k[x1,,xn]R=k[x_{1},\ldots,x_{n}] and S=k[x1,x2]S=k[x_{1},x_{2}] be two standard graded polynomial rings. Assume n2n\geq 2. Let AA be the set of SS-modules {k(a),k[x](a),S(a),(x1,x2)t(a),t,a}\{k(a),k[x](a),S(a),(x_{1},x_{2})^{t}(a),t\in\mathbb{N},a\in\mathbb{Z}\}. Identify these SS-modules as RR-modules via the ring map SR/(x3,,xn)S\cong R/(x_{3},\ldots,x_{n}). Then for every finitely generated graded RR-module MM of dimension at most 22, there exist N1,,NsAN_{1},\ldots,N_{s}\in A, r1,,rsr_{1},\ldots,r_{s}\in\mathbb{Q} and r1,,rs>0r_{1},\ldots,r_{s}>0 such that

H(M)=i=1sriH(Ni).H^{\bullet}(M)=\sum_{i=1}^{s}r_{i}H^{\bullet}(N_{i}).

Moreover, set AA describes the vertex set of the cone of local cohomology tables of finitely generated graded modules of dimension at most 22.

In this paper we will analyze this problem in a more general setting than the one considered in [5]. We will focus on the sequentially almost Cohen-Macaulay(saCM) case, which we will define in Section 3. The following is the first main result of this article:

Theorem (See Theorem 4.20).

Let R=k[x1,,xn]R=k[x_{1},\ldots,x_{n}] be a standard graded polynomial ring. Let CC be the cone generated by local cohomology tables of saCM modules. The vertex set of this cone is a set AMA_{M}; we can find a set AMA^{\prime}_{M} containing AMA_{M} such that for every local cohomology table H(M)H^{\bullet}(M) in the cone, there is a finite set of modules N1,,NsAMN_{1},\ldots,N_{s}\in A^{\prime}_{M} and r1,,rsr_{1},\ldots,r_{s}\in\mathbb{Q} and r1,,rs>0r_{1},\ldots,r_{s}>0 such that

H(M)=i=1sriH(Ni).H^{\bullet}(M)=\sum_{i=1}^{s}r_{i}H^{\bullet}(N_{i}).

However, when n3n\geq 3, then there is an saCM module M0M_{0} such that for any finite set of saCM modules N1,,NsAMN_{1},\ldots,N_{s}\in A_{M} and r1,,rsr_{1},\ldots,r_{s}\in\mathbb{Q} and r1,,rs>0r_{1},\ldots,r_{s}>0,

H(M)i=1sriH(Ni).H^{\bullet}(M)\neq\sum_{i=1}^{s}r_{i}H^{\bullet}(N_{i}).

That is, AMA_{M} does not generate the cone.

The sets AM,AMA_{M},A^{\prime}_{M} are defined in Section 4 and their descriptions rely on the concept of the Auslander transpose. For a finitely generated RR-module MM with a presentation matrix ϕ\phi, the Auslander transpose, denoted by Tr(M)\textup{Tr}(M), is the cokernel of ϕ\phi^{*}, where =HomR(,R)*=\textup{Hom}_{R}(\cdot,R) is the dual functor. It is not unique in general; it is only unique up to a free summand. Note that RR is graded local. So throughout this paper we use Tr(M)\textup{Tr}(M) to denote the minimal Auslander transpose, that is, cokerϕ\textup{coker}\phi^{*} where ϕ\phi is a minimal presentation matrix. This choice makes Tr()\textup{Tr}(\cdot) a map from the set of isomorphism classes of finitely generated graded RR-modules to itself. The main idea is that under proper assumptions on MM, there is a \mathbb{Q}-linear transformation that maps β(M)\beta^{\bullet}(M) to H(Tr(M))H^{\bullet}(\textup{Tr}(M)). The existence of such a transformation allows us to use the Boij-Söderberg theory of Betti tables to decompose a local cohomology table. The vertices of the cone of Betti tables are of the form β(M)\beta^{\bullet}(M) where MM has a pure resolution and every such table can be computed using the degree sequence of MM, so we can also compute H(Tr(M))H^{\bullet}(\textup{Tr}(M)) and determine when a local cohomology table decomposes.

To study the decomposition of local cohomology tables of modules of dimension 3, we may also assume that dimR=3\dim R=3 by Lemma 3.7. We first reduce to the case where depth(M)=1(M)=1 and MM has no dimension 1 submodule. Then we relate MM to the module of global sections Γ(M)\Gamma(M); in this case Γ(M)\Gamma(M) is finitely generated with depth(Γ(M))2(\Gamma(M))\geq 2. This means Γ(M)\Gamma(M) is Cohen-Macaulay or almost Cohen-Macaulay, so it is saCM and its cohomology table decomposes according to Theorem 4.20. The key point is whether a decomposition of H(Γ(M))H^{\bullet}(\Gamma(M)) induces a decomposition of H(M)H^{\bullet}(M), and it does in two cases, described by the following two theorems. In the first case, there is a submodule of dimension 2 of MM that induces a decomposition:

Theorem (See Theorem 6.8).

Let MM be a module of depth 11 and assume MM has no dimension 11 submodule. Let Γ=Γ(M)\Gamma=\Gamma(M), Q=Γ/TorΓQ=\Gamma/\textup{Tor}\Gamma. Then H(M)=H(TorM)+H(M/TorM)(0,HS(H𝔪1(Q)),HS(H𝔪1(Q)),0)H^{\bullet}(M)=H^{\bullet}(\textup{Tor}M)+H^{\bullet}(M/\textup{Tor}M)-(0,HS(H^{1}_{\mathfrak{m}}(Q)),HS(H^{1}_{\mathfrak{m}}(Q)),0). In particular, if H𝔪1(Q)=0H^{1}_{\mathfrak{m}}(Q)=0, then H(M)=H(TorM)+H(M/TorM)H^{\bullet}(M)=H^{\bullet}(\textup{Tor}M)+H^{\bullet}(M/\textup{Tor}M).

A description of the module H𝔪1(Q)H^{1}_{\mathfrak{m}}(Q) is given in Proposition 6.10. Note that here we view the vector with power series entries (0,HS(H𝔪1(Q)),HS(H𝔪1(Q)),0)(0,HS(H^{1}_{\mathfrak{m}}(Q)),HS(H^{1}_{\mathfrak{m}}(Q)),0) as a table; this is called a table in series form, which will be explained in Section 2.

In the second case there is a submodule of dimension 3 of MM that induces a decomposition:

Theorem (See Corollary 6.14).

Suppose L=ExtR2(TrM,R)=0L=\textup{Ext}^{2}_{R}(\textup{Tr}M,R)=0, Γ\Gamma be the module of global sections, and Γ=HomR(Γ,R)Tr(L)\Gamma^{*}=\textup{Hom}_{R}(\Gamma,R)\neq\textup{Tr}(L^{\prime}) for any module LL^{\prime} of finite length. Then H(M)=H(MF)+H(M/MF)H^{\bullet}(M)=H^{\bullet}(M\cap F)+H^{\bullet}(M/M\cap F) for some free module FΓF\subset\Gamma.

This paper consists of six sections. Section 1 is an introduction. Section 2 defines a table in series form, reviews some basic definitions on convex cones in a \mathbb{Q}-vector space, and some other basic propositions. Section 3 covers the concept of the dimension filtration introduced by Schenzel, and shows that to decompose the local cohomology table of an saCM module, it suffices to decompose the tables of a module of projective dimension at most 1; Section 4 shows how to decompose these tables. Section 5 introduces some propositions of the Γ\Gamma-functor and prove that in order to decompose H(M)H^{\bullet}(M), it suffices to decompose H(Γ(M))H^{\bullet}(\Gamma(M)) and HS(Γ(M)/M)HS(\Gamma(M)/M) simultaneously. Note that if the module has dimension 3, then we reduce to the case where the ring has dimension 3; in this case the projective dimension of Γ(M)\Gamma(M) is at most 1. Finally in Section 6, we find conditions under which we can decompose H(Γ(M))H^{\bullet}(\Gamma(M)) and HS(Γ(M)/M)HS(\Gamma(M)/M) simultaneously.

2. notations

In all the following sections, we assume R=k[x1,,xn]R=k[x_{1},\ldots,x_{n}] is a standard graded polynomial ring over a field kk, and let 𝔪=(x1,,xn)\mathfrak{m}=(x_{1},\ldots,x_{n}) be its graded maximal ideal. Let MM be a finitely generated graded RR-module. Conventionally, the Betti table of MM is the n×n\times\mathbb{Z}-table β(M)\beta^{\bullet}(M) with entries β(M)i,j=dimkToriR(M,k)i+j\beta^{\bullet}(M)_{i,j}=\dim_{k}\textup{Tor}^{R}_{i}(M,k)_{i+j}. The local cohomology table H(M)H^{\bullet}(M), by definition, is the n×n\times\mathbb{Z}-table defined by H(M)i,j=dimkH𝔪i(M)jH^{\bullet}(M)_{i,j}=\dim_{k}H^{i}_{\mathfrak{m}}(M)_{j}, and the Ext-table E(M)E^{\bullet}(M) is the n×n\times\mathbb{Z}-table defined by E(M)i,j=dimkE^{\bullet}(M)_{i,j}=\dim_{k}Ext(M,R)jRi{}^{i}_{R}(M,R)_{j}. Here is an example of a Betti table:

0 1 2 3
0 1 0 0 0
1 0 3 3 1
2 0 1 1 0
Table 1.

Table 1 corresponds to the minimal resolution of R/(x12,x1x2,x1x3,x23)R/(x_{1}^{2},x_{1}x_{2},x_{1}x_{3},x_{2}^{3}) over RR when n3n\geq 3, which is

0R(4)R(4)R(3)3R(3)R(2)3R.0\to R(-4)\to R(-4)\oplus R(-3)^{3}\to R(-3)\oplus R(-2)^{3}\to R.

Throughout this paper we use a different kind of notations, and the tables considered will be in series form. The space of Betti tables or Ext-tables is V=i=0n[[t]][t1]viV=\oplus^{n}_{i=0}\mathbb{Q}[[t]][t^{-1}]v_{i}. It is a free [[t]][t1]\mathbb{Q}[[t]][t^{-1}]-module of rank n+1n+1, and it is also a \mathbb{Q}-vector space. The space of local cohomology tables is V=i=0n[[t1]][t]viV^{*}=\oplus^{n}_{i=0}\mathbb{Q}[[t^{-1}]][t]v^{*}_{i}. The Betti table of MM is an element (β0(M),β1(M),,βn(M))V(\beta_{0}(M),\beta_{1}(M),\ldots,\beta_{n}(M))\in V defined by βi(M)=Σjβi,j(M)tj\beta_{i}(M)=\Sigma_{j\in\mathbb{Z}}\beta_{i,j}(M)t^{j}. Note that this is an unshifted Betti table; this is different from the usual convention. The Ext-table is an element (E0(M),E1(M)(E^{0}(M),E^{1}(M), …,En(M))VE^{n}(M))\in V where Ei(M)=ΣjE^{i}(M)=\Sigma_{j\in\mathbb{Z}}dimkExt(M,R)jRitj{}^{i}_{R}(M,R)_{j}t^{j}. The local cohomology table of MM is (h0(M),h1(M),,hn(M))(h^{0}(M),h^{1}(M),\ldots,h^{n}(M)) V\in V^{*} where hi(M)=Σjh^{i}(M)=\Sigma_{j\in\mathbb{Z}}dimH𝔪ik(M)jtj{}_{k}H^{i}_{\mathfrak{m}}(M)_{j}t^{j}. These two representations of a table are equivalent. For example, table 1 will become (1,3t2+t3,3t3+t4,t4)(1,3t^{2}+t^{3},3t^{3}+t^{4},t^{4}) in series form. The series form has two advantages: first, the entry of each table is a series, and they interact with the Hilbert series of graded modules; and second, the action of taking the difference of a table becomes multiplication by (1t)(1-t), which makes sense as V,VV,V^{*} are [t]\mathbb{Q}[t]-modules.

We want to consider convex cones CC in the vector space VV or VV^{*}, that is, subsets that are closed under multiplication by positive rational numbers and addition. We call the expression 1isaici\sum_{1\leq i\leq s}a_{i}c_{i} with ciC,ai,ai>0c_{i}\in C,a_{i}\in\mathbb{Q},a_{i}>0 a positive linear combination of c1,c2,,csc_{1},c_{2},\ldots,c_{s}. If cCc\in C is a positive linear combination of c1,c2,,csc_{1},c_{2},\ldots,c_{s}, we also say that cc decomposes into c1,c2,,csc_{1},c_{2},\ldots,c_{s}; we say the decomposition is trivial if s=1s=1 and in this case cc and c1c_{1} differ by a positive rational scalar. A generating set of the cone is a subset GG of the cone CC such that every element is a positive linear combination of elements in GG. We also say GG generates the cone CC if GG is a generating set. A vertex is an element that does not decompose nontrivially and the vertex set is the set of all vertices. We say that a ray inside the cone is extremal if it contains a vertex; in this case every element of the ray is a vertex except for the origin. In this paper, we will consider the vertex sets of 3 kinds of cones: the cones generated by the Betti tables, the Ext-tables, and the local cohomology tables. It is easy to see that if GG is a generating set and vv is a vertex, then vv must decompose trivially, which means that a positive multiple of vv is in GG. So to find the vertex set, we may find a generating set GG first and then find elements in GG that decompose trivially. The following lemma about cones is trivial but will be useful in Section 4.

Lemma 2.1.

Let L:W1W2L:W_{1}\to W_{2} be a linear map between vector spaces over \mathbb{Q}. Suppose CW1C\subset W_{1} is a cone with vertex set V1V_{1} and a generating set G1G_{1}. Then L(C)L(C) is a cone in W2W_{2} generated by L(G1)L(G_{1}); suppose the vertex set of this cone is V2V_{2}, then V2L(V1)V_{2}\subset L(V_{1}). Furthermore, V2V_{2} is also the subset of L(V1)L(V_{1}) which is not a positive linear combination of the other elements in L(G1)L(G_{1}). If moreover, LL is an injection, then V2=L(V1)V_{2}=L(V_{1}).

Let E=ER(k)E=E_{R}(k) be the graded injective hull of k=R/𝔪k=R/\mathfrak{m}. Recall that by local duality, H𝔪i(M)=HomR(H^{i}_{\mathfrak{m}}(M)=\textup{Hom}_{R}(Ext(M,R(n))Rni,E){}^{n-i}_{R}(M,R(-n)),E). This implies dim(H𝔪i(M)j)k={}_{k}(H^{i}_{\mathfrak{m}}(M)_{j})= dim(k{}_{k}(Ext(M,R)njRni){}^{n-i}_{R}(M,R)_{-n-j}), hence we have:

Proposition 2.2.

The \mathbb{Q}-linear map L0:VVL_{0}:V\to V^{*}, where

L0(f0(t),f1(t),,fn(t))=tn(fn(t1),fn1(t1),,f0(t1)),L_{0}(f_{0}(t),f_{1}(t),\ldots,f_{n}(t))=t^{-n}(f_{n}(t^{-1}),f_{n-1}(t^{-1}),\ldots,f_{0}(t^{-1})),

is invertible and H(M)=L0(E(M))H^{\bullet}(M)=L_{0}(E^{\bullet}(M)).

By the above proposition, the extremal rays of the cone of local cohomology tables and the cone of Ext-tables are in 1-1 correspondence under L0L_{0}. So to find the extremal rays of the cone generated by all local cohomology tables, it suffices to find those of all Ext-tables.

It is well known that the Betti numbers are nonzero for finitely many entries, and the dimension of the ii-th Ext module has dimension at most nin-i. So actually these tables sit in a proper subspace of VV or VV^{*}.

Proposition 2.3.

The following proposition holds.

(1) β(M)i=0n[t][t1]vi\beta^{\bullet}(M)\in\oplus^{n}_{i=0}\mathbb{Q}[t][t^{-1}]v_{i}.

(2) E(M)i=0n[t][t1]1(1t)niviE^{\bullet}(M)\in\oplus^{n}_{i=0}\mathbb{Q}[t][t^{-1}]\frac{1}{(1-t)^{n-i}}v_{i}.

(3) H(M)i=0n[t1][t]1(1t1)iviH^{\bullet}(M)\in\oplus^{n}_{i=0}\mathbb{Q}[t^{-1}][t]\frac{1}{(1-t^{-1})^{i}}v_{i}.

Suppose we have a short exact sequence of finitely generated graded RR-modules 0MMM"00\to M^{\prime}\to M\to M"\to 0. Then we have a long exact sequence of local cohomology modules. We see from the long exact sequence that H(M)=H(M)+H(M")H^{\bullet}(M)=H^{\bullet}(M^{\prime})+H^{\bullet}(M") if and only if all the connecting maps H𝔪i(M")H𝔪i+1(M)H^{i}_{\mathfrak{m}}(M")\to H^{i+1}_{\mathfrak{m}}(M^{\prime}) are 0; in this case we say that the exact sequence induces a decomposition of local cohomology tables. Finally, the depths of these modules are related by the well-known depth lemma:

Proposition 2.4 (Depth lemma).

Let 0MMM"00\to M^{\prime}\to M\to M"\to 0 be an exact sequence of finitely generated RR-modules, then:

(1) 0pt(M)min{0pt(M),0pt(M")}0pt(M)\geq\textup{min}\{0pt(M^{\prime}),0pt(M")\}.

(2) 0pt(M)min{0pt(M),0pt(M")+1}0pt(M^{\prime})\geq\textup{min}\{0pt(M),0pt(M")+1\}.

(3) 0pt(M")min{0pt(M)1,0pt(M)}0pt(M")\geq\textup{min}\{0pt(M^{\prime})-1,0pt(M)\}.

3. The dimension filtration

Let us recall the concept of the dimension filtration introduced by Schenzel in [9].

Let AA be a Noetherian ring of dimension dd, and MM be a finitely generated AA-module. We define MiM_{i} to be the largest submodule of MM such that dim(Mi)i\dim(M_{i})\leq i; such module exists by the Noetherian property. Then 0M0M1M2Md=M0\subset M_{0}\subset M_{1}\subset M_{2}\subset\ldots\subset M_{d}=M forms a filtration of MM, called the dimension filtration of MM. We say MiM_{i} is the largest submodule of MM of dimension at most ii; if it happens that dim(Mi)=i\dim(M_{i})=i we say it is the largest submodule of MM of dimension ii. We say Ni=Mi/Mi1N_{i}=M_{i}/M_{i-1} the ii-th dimension factor of MM; it is either 0 or of dimension ii, and MM has no nonzero submodule of dimension ii if and only if the ii-th dimension factor is 0.

Proposition 3.1.

Let MiM_{i} be the largest submodule of MM of dimension at most ii. The following holds:

(1) Ass(Mi)={𝔭Ass(M)|dimA/𝔭i}Ass(M_{i})=\{\mathfrak{p}\in Ass(M)|\dim A/\mathfrak{p}\leq i\}.

(2) Ass(M/Mi)={𝔭Ass(M)|dimA/𝔭>i}Ass(M/M_{i})=\{\mathfrak{p}\in Ass(M)|\dim A/\mathfrak{p}>i\}.

(3) Ass(Mi/Mi1)={𝔭Ass(M)|dimA/𝔭=i}Ass(M_{i}/M_{i-1})=\{\mathfrak{p}\in Ass(M)|\dim A/\mathfrak{p}=i\}.

Proof.

See Corollary 2.3 of [9]. The proof of the general case can be carried from that of the local case. ∎

Corollary 3.2.

Let MiM_{i} be the largest submodule of MM of dimension at most ii. The following holds:

(1) M/MiM/M_{i} has no nonzero submodule of dimension at most ii.

(2) Mi=0M_{i}=0 if and only if for any 𝔭Ass(M)\mathfrak{p}\in Ass(M), dim(A/𝔭)>i\dim(A/\mathfrak{p})>i. If AA is a local catenary domain, this is equivalent to ht(𝔭)<dimAi\textup{ht}(\mathfrak{p})<\dim A-i.

For a general module MM the dimension filtration of dimension at most 1 induces a decomposition of local cohomology tables.

Lemma 3.3.

Let MM be a finitely generated graded RR-module. Then M0=H𝔪0(M)M_{0}=H^{0}_{\mathfrak{m}}(M), and the exact sequence 0M0MM/M000\to M_{0}\to M\to M/M_{0}\to 0 induces a decomposition of local cohomology tables, and 0pt(M/M0)10pt(M/M_{0})\geq 1.

The proof is trivial and we omit it.

Lemma 3.4.

Let MM be a finitely generated graded RR-module of depth at least 1. Then the exact sequence 0M1MM/M100\to M_{1}\to M\to M/M_{1}\to 0 induces a decomposition of local cohomology tables.

Proof.

We have depthM1M_{1}\geq 1 because M1M_{1} is a submodule of MM and depthMM\geq 1. This means that M1M_{1} is either 0 or Cohen-Macaulay of dimension 1. Also, M/M1M/M_{1} does not have submodule of dimension at most 1; hence H𝔪0(M/M1)=0H^{0}_{\mathfrak{m}}(M/M_{1})=0. Now the long exact sequence of local cohomology modules breaks up into short exact sequences:

0H𝔪1(M1)H𝔪1(M)H𝔪1(M/M1)00\to H^{1}_{\mathfrak{m}}(M_{1})\to H^{1}_{\mathfrak{m}}(M)\to H^{1}_{\mathfrak{m}}(M/M_{1})\to 0

and

0H𝔪i(M)H𝔪i(M/M1)00\to H^{i}_{\mathfrak{m}}(M)\to H^{i}_{\mathfrak{m}}(M/M_{1})\to 0

for any i2i\geq 2, and all the connecting homomorphisms are 0, so H(M)=H(M1)+H(M/M1)H^{\bullet}(M)=H^{\bullet}(M_{1})+H^{\bullet}(M/M_{1}). ∎

Recall that a module MM is almost Cohen-Macaulay if 0pt(M)=dim(M)10pt(M)=\dim(M)-1. We give the definition of sequentially Cohen-Macaulay introduced by Stanley and generalize to sequentially almost Cohen-Macaulay using the dimension filtration:

Definition 3.5.

Let MM be a finitely generated AA-module. We say MM is sequentially Cohen-Macaulay if all its nonzero dimension factors are Cohen-Macaulay and sequentially almost Cohen-Macaulay (saCM) if all its nonzero dimension factors are Cohen-Macaulay or almost Cohen-Macaulay.

The ii-th dimension factor NiN_{i} of MM must have dimension ii if it is nonzero, so if MM is sequentially Cohen-Macaulay then 0ptNi=i0ptN_{i}=i, and if MM is saCM then 0ptNi=i0ptN_{i}=i or i1i-1. Note that by definition sequentially Cohen-Macaulay implies saCM.

The saCM modules have an important property: their local cohomology tables decompose into local cohomology tables of its dimension factors. More precisely, we have:

Proposition 3.6.

Let MM be saCM, MiM_{i} be its largest submodule of dimension at most ii, Ni=Mi/Mi1N_{i}=M_{i}/M_{i-1} be its dimension factors.

(1) 0ptM/Mii0ptM/M_{i}\geq i for any 0in10\leq i\leq n-1.

(2) 0Mi/Mi1M/Mi1M/Mi00\to M_{i}/M_{i-1}\to M/M_{i-1}\to M/M_{i}\to 0 induces a decomposition in local cohomology tables.

(3) H(M)=i=0nH(Ni)H^{\bullet}(M)=\sum^{n}_{i=0}H^{\bullet}(N_{i}).

Proof.

(1) We prove this by induction from i=n1i=n-1; in this case 0ptM/Mn1n10ptM/M_{n-1}\geq n-1 by the saCM assumption because M/Mn1M/M_{n-1} is just the nn-th dimension factor. Suppose (1) is true for ii. Consider the exact sequence 0Mi/Mi1M/Mi1M/Mi00\to M_{i}/M_{i-1}\to M/M_{i-1}\to M/M_{i}\to 0. We have 0pt(Mi/Mi1)i10pt(M_{i}/M_{i-1})\geq i-1 by the saCM assumption and 0pt(M/Mi)i0pt(M/M_{i})\geq i by the induction hypothesis, so by the depth lemma, 0pt(M/Mi1)i10pt(M/M_{i-1})\geq i-1, which implies that (1) is true for i1i-1, hence by induction (1) is true for any 0in10\leq i\leq n-1.

(2) The boundary maps of the long exact sequence of local cohomology modules are H𝔪k(M/Mi)H𝔪k+1(Mi/Mi1)H^{k}_{\mathfrak{m}}(M/M_{i})\to H^{k+1}_{\mathfrak{m}}(M_{i}/M_{i-1}). We have dim(Mi/Mi1)i\dim(M_{i}/M_{i-1})\leq i and 0pt(M/Mi)i0pt(M/M_{i})\geq i. Thus if ki1k\leq i-1 then the source of the map is 0; if kik\geq i then the target is 0. Therefore, the boundary maps are 0 and the exact sequence induces a decomposition in local cohomology tables.

(3) Apply (2) inductively. ∎

It follows that to find the decomposition of local cohomology tables of saCM modules, we only need to decompose the tables of almost Cohen-Macaulay modules.

There is an important principle for modules of lower dimension: if dim(M)<dim(R)\dim(M)<\dim(R), then to find the decomposition of H(M)H^{\bullet}(M), we may always replace RR with a new polynomial ring SS with dim(S)=dim(M)\dim(S)=\dim(M). This is done by the following proposition:

Lemma 3.7.

Let R=k[x1,,xn]R=k[x_{1},\ldots,x_{n}] and S=k[y1,,yd]S=k[y_{1},\ldots,y_{d}] be two standard graded polynomial rings with ndn\geq d. For a finitely generated graded RR-module MM with dim(M)d\dim(M)\leq d, there is a finitely generated graded SS-module NN such that H(M)H^{\bullet}(M) can be obtained from H(N)H^{\bullet}(N) by multiplying a positive rational scalar and adding 0’s.

Proof.

If kk is infinite, see Lemma 2.2 of [5]. If kk is finite, let l=k¯l=\bar{k} be its algebraic closure. Then by Lemma 2.2 of [5], there is an Sl=l[y1,,yd]S_{l}=l[y_{1},\ldots,y_{d}]-module NN such that H(N)H^{\bullet}(N) and H(M)H^{\bullet}(M) differ by a positive rational scalar. But NN is finitely generated, so there is a finite extension kk^{\prime} of kk in ll such that NN is extended from Sk=k[y1,,yd]S_{k^{\prime}}=k^{\prime}[y_{1},\ldots,y_{d}], that is, there is an SkS_{k^{\prime}}-module NkN_{k^{\prime}} such that N=NkklN=N_{k^{\prime}}\otimes_{k^{\prime}}l. Then H(N)=H(Nk)H^{\bullet}(N)=H^{\bullet}(N_{k^{\prime}}). Since S=k[y1,,yd]S=k[y_{1},\ldots,y_{d}] embeds into SkS_{k^{\prime}} and this is module-finite, we can endow NkN_{k^{\prime}} with an SS-module structure and it becomes a finitely generated graded module, say MM^{\prime}. Then H(Nk)=H(M)[k:k]H^{\bullet}(N_{k^{\prime}})=H^{\bullet}(M^{\prime})[k^{\prime}:k]. Thus, H(M)H^{\bullet}(M) and H(M)H^{\bullet}(M^{\prime}) only differ by a positive rational scalar. ∎

Remark 3.8.

Lemma 3.7 is a variant of Lemma 2.2 of [5] except that we do not change the base field. The table H(M)H^{\bullet}(M) consists of n+1n+1 series and the table H(N)H^{\bullet}(N) consists of d+1d+1 series so we need to add ndn-d 0’s to make their sizes equal.

The 0 entries in H(M)H^{\bullet}(M) remains 0 in H(N)H^{\bullet}(N), so for any choice of NN, we have 0pt(M)=0pt(N)0pt(M)=0pt(N) and dim(M)=dim(N)\dim(M)=\dim(N). If we fix a choice of NN for every MM and view it as a correspondence, then almost Cohen-Macaulay modules correspond to almost Cohen-Macaulay modules of maximal dimension, which are just the modules with maximal dimension and projective dimension 1. We will find how to decompose the local cohomology tables of modules of projective dimension 1 in Section 4.

Remark 3.9.

Let R=k[x1,,xn]R=k[x_{1},\ldots,x_{n}] and S=k[y1,,yd]S=k[y_{1},\ldots,y_{d}] be two standard graded polynomial rings with ndn\geq d. Then SR/(xn,,xd+1)S\cong R/(x_{n},\ldots,x_{d+1}). The local cohomology table of an SS-module MM does not change if we view MM as an RR-module, so there is a natural inclusion from {H(M),M\{H^{\bullet}(M),M is an SS-module}\} to {H(M),M\{H^{\bullet}(M),M is an RR-module}\}. Hence there is a natural inclusion of their cones, that is, 0{H(M),M\mathbb{Q}_{\geq 0}\{H^{\bullet}(M),M is an SS-module}\} injects into 0{H(M),M\mathbb{Q}_{\geq 0}\{H^{\bullet}(M),M is an RR-module}\}. By Lemma 3.7, the image is just 0{H(M),M\mathbb{Q}_{\geq 0}\{H^{\bullet}(M),M is an RR-module, dimMdimS}\dim M\leq\dim S\}. In this sense, when we study the decomposition of local cohomology tables of RR-modules of dimension at most dd, it suffices to study the decomposition of local cohomology tables of SS-modules.

4. Projective dimension 1 case

This section describes the cone generated by H(M)H^{\bullet}(M) where MM is an RR-module with projdim(MM) \leq 1, how the tables in this cone decompose, and how the decomposition of such tables leads to the decomposition of tables of saCM modules.

First, let us recall the definition of Auslander transpose introduced by Auslander and Bridger in [1]. Let =HomR(,R)\cdot^{*}=\textup{Hom}_{R}(\cdot,R) be the dual functor.

Definition 4.1.

Let MM be a finitely generated module over a RR. Consider a finite presentation F1ϕF0M0F_{1}\xrightarrow{\phi}F_{0}\to M\to 0. Taking dual yields an exact sequence 0MF0ϕF1N00\to M^{*}\to F_{0}^{*}\xrightarrow{\phi^{*}}F_{1}^{*}\to N\to 0. Then the Auslander transpose of MM is Tr(M)=N=Coker(ϕ)\textup{Tr}(M)=N=\textup{Coker}(\phi^{*}). If ϕ\phi is a minimal presentation, then Tr(M)\textup{Tr}(M) is called the minimal Auslander transpose.

Remark 4.2.

In general, the Auslander transpose is unique up to a projective summand. However, the minimal Auslander transpose is unique up to isomorphism; so in the following sections, when we mention the module Tr(M)\textup{Tr}(M), we always mean the minimal Auslander transpose.

Here are some basic properties of the Auslander transpose.

Proposition 4.3.

(1) The Auslander transpose of a graded module is also graded.

(2) If projdim(M)1(M)\leq 1, then Tr(M)=ExtR1(M,R)\textup{Tr}(M)=\textup{Ext}^{1}_{R}(M,R).

(3) Tr(M)=0\textup{Tr}(M)=0 if and only if projdim(M)=0(M)=0, that is, MM is free.

(4) Tr(MM)=Tr(M)Tr(M)\textup{Tr}(M\oplus M^{\prime})=\textup{Tr}(M)\oplus\textup{Tr}(M^{\prime}).

(5) M=Tr(Tr(M))FM=\textup{Tr}(\textup{Tr}(M))\oplus F, FF is free and Tr(Tr(M))\textup{Tr}(\textup{Tr}(M)) does not have a free summand.

Let MM be a finitely generated graded RR-module. To study the decomposition of H(M)H^{\bullet}(M), we may assume MM is indecomposable without loss of generality. If projdimM=0M=0 then MM is free, and we must have MR(i)M\cong R(-i) for some ii. In this case H(M)H^{\bullet}(M) is clear. So we may assume that projdim(M)=1(M)=1. Here is an important observation about the properties of MM and Tr(M)\textup{Tr}(M).

Proposition 4.4.

(1) Let MM be a finitely generated RR-module of projective dimension 1 with no free summand. Let 0F1ϕF0M00\to F_{1}\xrightarrow{\phi}F_{0}\to M\to 0 be a minimal presentation which is also a minimal resolution. Let N=Tr(M)0N=\textup{Tr}(M)\neq 0, then dimNN \leq n1n-1, and F0ϕF1N0F_{0}^{*}\xrightarrow{\phi^{*}}F_{1}^{*}\to N\to 0 is a minimal presentation of NN.

(2) Let NN be a nonzero module with dim(N)(N) \leq n1n-1. Take a minimal presentation G1𝜓G0N0G_{1}\xrightarrow{\psi}G_{0}\to N\to 0. Then G0ψG1G_{0}^{*}\xrightarrow{\psi^{*}}G_{1}^{*} is injective and its image lies in 𝔪G1\mathfrak{m}G_{1}^{*}, so if M=Tr(N)=Coker(ψ)M=\textup{Tr}(N)=\textup{Coker}(\psi^{*}), then projdimM=1M=1 and MM has a minimal resolution 0G0ψG1M00\to G_{0}^{*}\xrightarrow{\psi^{*}}G_{1}^{*}\to M\to 0. Also, MM does not have a free summand.

(3) Taking the Auslander transpose Tr induces a 1-1 correspondence between the isomorphism classes of finitely generated graded RR-modules MM of projective dimension 1 without a free summand and finitely generated graded RR-modules NN of dimension at most n1n-1.

(4) Under the assumption in (3), β0,j(M)=β1,j(N)\beta_{0,j}(M)=\beta_{1,-j}(N), β1,j(M)=β0,j(N)\beta_{1,j}(M)=\beta_{0,-j}(N), or equivalently, β0(M)(t)=β1(N)(t1),β1(M)(t)=β0(N)(t1)\beta_{0}(M)(t)=\beta_{1}(N)(t^{-1}),\beta_{1}(M)(t)=\beta_{0}(N)(t^{-1}).

Proof.

(1) Suppose projdimMM = 1 with minimal resolution 0F1ϕF0M00\to F_{1}\xrightarrow{\phi}F_{0}\to M\to 0. Then N=N= Ext(M,R)R1{}^{1}_{R}(M,R), so dimNN \leq n1n-1. Now consider the exact sequence F0ϕF1N0F_{0}^{*}\xrightarrow{\phi^{*}}F_{1}^{*}\to N\to 0. Since ϕ\phi has entries in 𝔪\mathfrak{m}, so does ϕ\phi^{*}, so the image of F1F_{1}^{*} generates NN minimally. Now F0F_{0}^{*} surjects onto Syz(N)1{}_{1}(N); if the image of one basis of F0F_{0}^{*} is not a minimal generating set of the image of ϕ\phi^{*}, then some basis element eie_{i} of F0F_{0}^{*} is mapped to 0 under ϕ\phi^{*}. Then taking the dual again, ReiRe_{i}^{*} will become a free summand of MM, contradicting with our assumption. So the image of the basis of F0F_{0}^{*} is a minimal generating set of Syz(N)1{}_{1}(N). This means that F0F1N0F_{0}^{*}\to F_{1}^{*}\to N\to 0 is a minimal presentation of NN.

(2) Take a minimal presentation G1𝜓G0N0G_{1}\xrightarrow{\psi}G_{0}\to N\to 0. Let M=Tr(N)M=\textup{Tr}(N), then G0ψG1M0G_{0}^{*}\xrightarrow{\psi^{*}}G_{1}^{*}\to M\to 0 is exact. By minimality ψ\psi has entries in 𝔪\mathfrak{m}, hence so does ψ\psi^{*}. Let KK = Quot(RR) the quotient field of RR. Then NK=0N\otimes K=0, hence ψK\psi\otimes K is surjective and it is also a KK-linear map where KK is a field. So (ψK)=ψK(\psi\otimes K)^{*}=\psi^{*}\otimes K is injective. This implies that Ker(ψ)(\psi^{*}) is torsion, but it is a submodule of G0G_{0}^{*}, so it must be 0. In other words, ψ\psi^{*} is injective and this means 0G0G1M00\to G_{0}^{*}\to G_{1}^{*}\to M\to 0 is a minimal resolution of MM. If MM has a free summand, it must be generated by the image of some basis elements of G1G_{1}^{*}. Pick one of these basis elements eie_{i} and expand it to a basis of G1G_{1}^{*}, then we know that the eie_{i}-coefficient of all elements in ψ(G0)\psi(G_{0}^{*}) is 0. Taking dual again, we get ψ(ei)=0\psi(e_{i}^{*})=0, which means that G1G_{1} is not mapped to Syz(N)1{}_{1}(N) minimally. This is a contradiction. Hence MM has no free summands.

(3) Obvious by (1) and (2).

(4) Obvious by (1), (2) and (3). ∎

For a graded module MM, let HS(M)=i0dimk(Mi)tiHS(M)=\sum_{i\geq 0}\dim_{k}(M_{i})t^{i} denote its Hilbert series. Then the Betti table of the Auslander transpose describes the local cohomology table under the same assumption as above.

Proposition 4.5.

Let MM be a finitely generated graded RR-module without a free summand, projdim(M)=1(M)=1, N=Tr(M)N=\textup{Tr}(M), then E(M)=(1t)n(i=2n(1)iβi(N)E^{\bullet}(M)=(1-t)^{-n}(\sum_{i=2}^{n}(-1)^{i}\beta_{i}(N), i=0n(1)iβi(N),0,,0)\sum_{i=0}^{n}(-1)^{i}\beta_{i}(N),0,\ldots,0).

Proof.

Let 0F1F0M00\to F_{1}\to F_{0}\to M\to 0 be the minimal resolution of MM. This induces an exact sequence 0MF0F1N00\to M^{*}\to F_{0}^{*}\to F_{1}^{*}\to N\to 0 where F0F1N0F_{0}^{*}\to F_{1}^{*}\to N\to 0 is a minimal representation. By definition, E(M)=(HS(M),HS(N),0,,0)E^{\bullet}(M)=(HS(M^{*}),HS(N),0,\ldots,0). Now F0F1N0F_{0}^{*}\to F_{1}^{*}\to N\to 0 is a minimal presentation, so HS(F1)=β0(N)HS(R)=β0(N)(1t)n,HS(F0)=β1(N)HS(R)=β1(N)(1t)nHS(F_{1}^{*})=\beta_{0}(N)HS(R)=\beta_{0}(N)(1-t)^{-n},HS(F_{0}^{*})=\beta_{1}(N)HS(R)=\beta_{1}(N)(1-t)^{-n}. Now HS(N)=(i=0n(1)iβi(N))(1t)nHS(N)=(\sum_{i=0}^{n}(-1)^{i}\beta_{i}(N))(1-t)^{-n} and by the long exact sequence, HS(M)=HS(F0)+HS(N)HS(F1)=(i=2n(1)iβi(N))(1t)nHS(M^{*})=HS(F_{0}^{*})+HS(N)-HS(F_{1}^{*})=(\sum_{i=2}^{n}(-1)^{i}\beta_{i}(N))(1-t)^{-n}. ∎

Corollary 4.6.

Let VV be the space of Betti tables and Ext tables, MM, NN be two modules satisfying the assumption in Proposition 4.5. Define L1:VVL_{1}:V\to V to be the \mathbb{Q}-linear map (β0,β1,,βn)(1t)n(i=2n(1)iβi,i=0n(1)iβi,0,,0)(\beta_{0},\beta_{1},\ldots,\beta_{n})\to(1-t)^{-n}(\sum_{i=2}^{n}(-1)^{i}\beta_{i},\sum_{i=0}^{n}(-1)^{i}\beta_{i},0,\ldots,0), then E(M)=L1(β(N))E^{\bullet}(M)=L_{1}(\beta^{\bullet}(N)).

Corollary 4.7.

Let CwfC_{wf} be the cone in VV generated by the Ext-tables of modules of projective dimension 11 which does not have a free summand. Then if E(M)E^{\bullet}(M) is an extremal ray and N=Tr(M)N=\textup{Tr}(M), then NN has a pure resolution of length at least 11, and every element in CwfC_{wf} is a positive linear combination of elements of the form E(M)E^{\bullet}(M), where Tr(M)\textup{Tr}(M) has a pure resolution of length at least 11.

Proof.

Let CbC_{b} be the cone generated by all Betti tables of modules of dimension at most n1n-1. Then by Proposition 4.4 (3) and Corollary 4.6, Cwf=L1(Cb)C_{wf}=L_{1}(C_{b}). Applying the Boij-Söderberg theory for Betti tables we know that the extremal rays of CbC_{b} is the Betti tables of modules with pure resolutions of length ss, where 1sn1\leq s\leq n and CbC_{b} is generated by these elements as a cone. Now apply Lemma 2.1. ∎

By the proposition above, we already know how to decompose E(M)E^{\bullet}(M) when N=Tr(M)N=\textup{Tr}(M) is not pure, so to find the vertices of the cone of Ext-tables it suffices to analyze when L1(β(N))L_{1}(\beta^{\bullet}(N)) is decomposable, where NN has a pure resolution. First, we need two lemmas that allow us to compute β(N)\beta^{\bullet}(N) when NN is pure in terms of its degree sequence d=(d0,,ds)\textbf{d}=(d_{0},\ldots,d_{s}). By Corollary 4.7 we may always assume s1s\geq 1.

Lemma 4.8.

Let s1ss_{1}\leq s be two nonnegative integers, d=(d0,,ds)\textbf{d}=(d_{0},\ldots,d_{s}) be a degree sequence, and Vd,s1V_{\textbf{d},s_{1}} be the vector space {f[t,t1]such thatf=i=0sπd,ditdi,(1t)s1|f}\{f\in\mathbb{Q}[t,t^{-1}]\textup{such that}f=\sum_{i=0}^{s}\pi_{\textbf{d},d_{i}}t^{d_{i}},(1-t)^{s_{1}}|f\}. Then dimVd,s1=ss1+1\dim_{\mathbb{Q}}V_{\textbf{d},s_{1}}=s-s_{1}+1.

Proof.

Multiplying td0t^{-d_{0}} does not affect the order of the pole at t=1t=1, so we may assume d0=0d_{0}=0 without loss of generality. In this case, every element in Vd,s1V_{\textbf{d},s_{1}} will be a polynomial. For a polynomial ff, (1t)s1(1-t)^{s_{1}} divides ff if and only if

djf/dtj(1)=0,j=0,1,,s11.d^{j}f/dt^{j}(1)=0,\forall j=0,1,\ldots,s_{1}-1.

Assume f=i=0sπd,ditdif=\sum_{i=0}^{s}\pi_{\textbf{d},d_{i}}t^{d_{i}}, then we have

i=0sπd,di(dij)=0,j=0,1,,s11.\sum_{i=0}^{s}\pi_{\textbf{d},d_{i}}{{d_{i}}\choose{j}}=0,\forall j=0,1,\ldots,s_{1}-1.

On the other hand, the set {dij,j=0,1,,s11}\{{d_{i}}^{j},j=0,1,\ldots,s_{1}-1\} can be mapped to {\{(dij){d_{i}}\choose{j}, j=0,1,,s11}j=0,1,\ldots,s_{1}-1\} using an invertible linear map. Thus the above equation is equivalent to

i=0sπd,didij=0,j=0,1,,s11.\sum_{i=0}^{s}\pi_{\textbf{d},d_{i}}{d_{i}}^{j}=0,\forall j=0,1,\ldots,s_{1}-1.

Note that the matrix (dij)(i,j),0is,1js1({d_{i}}^{j})_{(i,j)},0\leq i\leq s,1\leq j\leq s_{1} has full rank s1s_{1} because it has a Vandermonde submatrix of rank s1s_{1} and we have s+1s+1 variables, so the dimension of the solution space is ss1+1s-s_{1}+1. ∎

In the case where s=s1s=s_{1}, dimVd,s=1{}_{\mathbb{Q}}V_{\textbf{d},s}=1, so there is a unique vector up to a scalar. Denote the sign function by sgnsgn, then this vector has the alternating sign property, described as below.

Lemma 4.9.

(1) For each degree sequence d=(d0,,ds)\textbf{d}=(d_{0},\ldots,d_{s}), dimVd,s=1{}_{\mathbb{Q}}V_{\textbf{d},s}=1, hence there exists a unique polynomial πd(t)[t,t1]\pi_{\textbf{d}}(t)\in\mathbb{Q}[t,t^{-1}] up to multiplying by a nonzero rational number inside Vd,sV_{\textbf{d},s}, denoted by πd(t)=i=0sπd,ditdi\pi_{\textbf{d}}(t)=\sum_{i=0}^{s}\pi_{\textbf{d},d_{i}}t^{d_{i}}.

(2) If we rescale these coefficients so that πd,d0=1\pi_{\textbf{d},d_{0}}=1, then πd,di=Πj0(djd0)Πji(djdi)\pi_{\textbf{d},d_{i}}=\frac{\Pi_{j\neq 0}(d_{j}-d_{0})}{\Pi_{j\neq i}(d_{j}-d_{i})}, and sgn(πd,di)=(1)isgn(\pi_{\textbf{d},d_{i}})=(-1)^{i}, that is, the coefficients are nonzero and have alternating signs.

(3) Under the assumption in (2), πd(t)/(1t)s|t=1>0\pi_{\textbf{d}}(t)/(1-t)^{s}|_{t=1}>0.

(4) Up to multiplying by a scalar, β(N)=((1)iπd,ditdi)\beta^{\bullet}(N)=((-1)^{i}\pi_{\textbf{d},d_{i}}t^{d_{i}}) if NN is pure of type d.

The proof is shown in the beginning of Section 2.1 of [2] up to Definition 2.3. For example, we have π0,1,2,3(t)=13t+3t2t3\pi_{0,1,2,3}(t)=1-3t+3t^{2}-t^{3}, and π0,1,3,4(t)=12t+2t3t4\pi_{0,1,3,4}(t)=1-2t+2t^{3}-t^{4}.

To analyze the positive relation, we need to introduce more notions. We define another invertible linear map L2:VVL_{2}:V\to V, L2(f0,f1,f2,,fn)=(1t)n(f1f0,f0,f2,,fn)L_{2}(f_{0},f_{1},f_{2},\ldots,f_{n})=(1-t)^{n}(f_{1}-f_{0},f_{0},f_{2},\ldots,f_{n}). Since L2L_{2} is an isomorphism of \mathbb{Q}-vector spaces, it induces a bijection between the vertex set of a cone with the vertex set of the image of the cone by Lemma 2.1. Under this notation, L2L1(β(N))=(β0(N)β1(N),i=2n(1)iβi(N),0,,0)L_{2}L_{1}(\beta^{\bullet}(N))=(\beta_{0}(N)-\beta_{1}(N),\sum_{i=2}^{n}(-1)^{i}\beta_{i}(N),0,\ldots,0). So this element separates the polynomial πd(t)\pi_{\textbf{d}}(t) into two parts, the first part is the sum of the first two terms and the second part is the sum of the rest. Hence it is natural to introduce the following notation for a degree sequence d: let αd(t)=πd,d0td0+πd,d1td1\alpha_{\textbf{d}}(t)=\pi_{\textbf{d},d_{0}}t^{d_{0}}+\pi_{\textbf{d},d_{1}}t^{d_{1}} and αd(t)=i=2sπd,ditdi=πd(t)αd(t)\alpha_{\textbf{d}}^{\prime}(t)=\sum_{i=2}^{s}\pi_{\textbf{d},d_{i}}t^{d_{i}}=\pi_{\textbf{d}}(t)-\alpha_{\textbf{d}}(t). More generally, for an integer dd, define τd\tau_{\leq d} to be a map that sends a Laurent polynomial f[t,t1]f\in\mathbb{Q}[t,t^{-1}] to the sum of terms of ff of degree less than dd and τd\tau_{\geq d}, that sends a Laurent polynomial f[t,t1]f\in\mathbb{Q}[t,t^{-1}] to the sum of terms of ff of degree at least dd. It is easy to see that for a degree sequence d=(d0<d1<<ds)\textbf{d}=(d_{0}<d_{1}<\ldots<d_{s}), αd(t)=τd1πd(t)\alpha_{\textbf{d}}(t)=\tau_{\leq d_{1}}\pi_{\textbf{d}}(t) and αd(t)=τd2πd(t)\alpha_{\textbf{d}}^{\prime}(t)=\tau_{\geq d_{2}}\pi_{\textbf{d}}(t). We have:

Proposition 4.10.

L2L1(β(N))=(αd(t),αd(t))L_{2}L_{1}(\beta^{\bullet}(N))=(\alpha_{\textbf{d}}(t),\alpha_{\textbf{d}}^{\prime}(t)).

We need to check whether L2L1(β(N))L_{2}L_{1}(\beta^{\bullet}(N)) can be decomposed for various d’s. The next proposition shows that if there is a space between did_{i} and di+1d_{i+1} for i1i\neq 1, then we can decompose L2L1(β(N))L_{2}L_{1}(\beta^{\bullet}(N)).

Proposition 4.11.

Let s1s\geq 1 be a positive integer. Let d=(d0<d1<<ds)\textbf{d}=(d_{0}<d_{1}<\ldots<d_{s}) be a degree sequence. Assume di<di+11d_{i}<d_{i+1}-1 for some ii and pick an integer aa such that di<a<di+1d_{i}<a<d_{i+1}. Define two degree sequences d’=(d0<d1<<di<a<di+2<ds)\textbf{d'}=(d_{0}<d_{1}<\ldots<d_{i}<a<d_{i+2}<d_{s}) and d”=(d0<d1<<di1<a<di+1<ds)\textbf{d''}=(d_{0}<d_{1}<\ldots<d_{i-1}<a<d_{i+1}<d_{s}). Then πd(t)=c1πd’(t)+c2πd”(t)\pi_{\textbf{d}}(t)=c_{1}\pi_{\textbf{d'}}(t)+c_{2}\pi_{\textbf{d''}}(t) where c1,c2>0c_{1},c_{2}>0. Moreover, if i1i\neq 1, then we also have αd(t)=c1αd’(t)+c2αd”(t)\alpha_{\textbf{d}}(t)=c_{1}\alpha_{\textbf{d'}}(t)+c_{2}\alpha_{\textbf{d''}}(t) and αd(t)=c1αd’(t)+c2αd”(t)\alpha_{\textbf{d}}^{\prime}(t)=c_{1}\alpha_{\textbf{d'}}^{\prime}(t)+c_{2}\alpha_{\textbf{d''}}^{\prime}(t), so in particular, let NN be a pure module of type d, then L2L1(β(N))=c1L2L1(β(N))+c2L2L1(β(N′′))L_{2}L_{1}(\beta^{\bullet}(N))=c^{\prime}_{1}L_{2}L_{1}(\beta^{\bullet}(N^{\prime}))+c^{\prime}_{2}L_{2}L_{1}(\beta^{\bullet}(N^{\prime\prime})) where NN^{\prime} is pure of type d’, N′′N^{\prime\prime} is pure of type d”, and c1,c2>0c^{\prime}_{1},c^{\prime}_{2}>0 are elements in \mathbb{Q}.

Proof.

If s=1s=1, then πd(t)=td0td1\pi_{\textbf{d}}(t)=t^{d_{0}}-t^{d_{1}}, πd’(t)=td0ta\pi_{\textbf{d'}}(t)=t^{d_{0}}-t^{a}, πd”(t)=tatd1\pi_{\textbf{d''}}(t)=t^{a}-t^{d_{1}}, so πd(t)=πd’(t)+πd”(t)\pi_{\textbf{d}}(t)=\pi_{\textbf{d'}}(t)+\pi_{\textbf{d''}}(t). Now assume s2s\geq 2, then by Lemma 4.9 (2) we know πd’(t)\pi_{\textbf{d'}}(t) has nonzero coefficients at degree dj,ji+1d_{j},j\neq i+1 and degree aa, and πd”(t)\pi_{\textbf{d''}}(t) has nonzero coefficients at degree dj,jid_{j},j\neq i and degree aa. So by cancelling the coefficients in degree aa, there is a linear combination c1πd’(t)+c2πd”(t)c_{1}\pi_{\textbf{d'}}(t)+c_{2}\pi_{\textbf{d''}}(t) which is a polynomial with possible nonzero coefficients at degree di,0isd_{i},0\leq i\leq s. Now sgn(πd’,a)=(1)i+1sgn(πd”,a)=(1)isgn(\pi_{\textbf{d'},a})=(-1)^{i+1}\neq sgn(\pi_{\textbf{d''},a})=(-1)^{i}. Hence we have sgn(c1)=sgn(c2)sgn(c_{1})=sgn(c_{2}). This polynomial is still divisible by (1t)s(1-t)^{s}, so by Lemma 4.9 (1), it is a multiple of πd(t)\pi_{\textbf{d}}(t), and after rescaling we may assume πd(t)=c1πd’(t)+c2πd”(t)\pi_{\textbf{d}}(t)=c_{1}\pi_{\textbf{d'}}(t)+c_{2}\pi_{\textbf{d''}}(t) and sgn(c1)=sgn(c2)sgn(c_{1})=sgn(c_{2}). Now since s2s\geq 2, we have i1i\geq 1 or i+1s1i+1\leq s-1. In the first case, sgn(πd,d0)=sgn(πd’,d0)=sgn(πd”,d0)=1sgn(\pi_{\textbf{d},d_{0}})=sgn(\pi_{\textbf{d'},d_{0}})=sgn(\pi_{\textbf{d''},d_{0}})=1 and in the second case sgn(πd,ds)=sgn(πd’,ds)=sgn(πd”,ds)=(1)ssgn(\pi_{\textbf{d},d_{s}})=sgn(\pi_{\textbf{d'},d_{s}})=sgn(\pi_{\textbf{d''},d_{s}})=(-1)^{s}, so c1,c2>0c_{1},c_{2}>0. If i1i\neq 1, then either i=0,i+1=1i=0,i+1=1 or i2i\geq 2. We can apply τd1\tau_{\leq d_{1}} to the equation πd(t)=c1πd’(t)+c2πd”(t)\pi_{\textbf{d}}(t)=c_{1}\pi_{\textbf{d'}}(t)+c_{2}\pi_{\textbf{d''}}(t) to get αd(t)=c1αd’(t)+c2αd”(t)\alpha_{\textbf{d}}(t)=c_{1}\alpha_{\textbf{d'}}(t)+c_{2}\alpha_{\textbf{d''}}(t) and apply τd2\tau_{\geq d_{2}} to get αd(t)=c1αd’(t)+c2αd”(t)\alpha_{\textbf{d}}^{\prime}(t)=c_{1}\alpha_{\textbf{d'}}^{\prime}(t)+c_{2}\alpha_{\textbf{d''}}^{\prime}(t). The last statement is true for c1=c1c^{\prime}_{1}=c_{1} and c2=c2c^{\prime}_{2}=c_{2} by Proposition 4.10. ∎

Corollary 4.12.

Let d0=(d0,0<d0,1<<d0,s)\textbf{d}_{0}=(d_{0,0}<d_{0,1}<\ldots<d_{0,s}) be a degree sequence. For a degree sequence di=(di,0<di,1<<di,s),1id1d0\textbf{d}_{i}=(d_{i,0}<d_{i,1}<\ldots<d_{i,s}),1\leq i\leq d_{1}-d_{0} we say it satisfies condition 𝒫\mathcal{P} if di,1di,0=1d_{i,1}-d_{i,0}=1 and di,j=di,2+j2d_{i,j}=d_{i,2}+j-2 for 2is2\leq i\leq s. Let NN be pure of type d0\textbf{d}_{0}, then there exist a collection of NiN_{i}’s which are pure of type di\textbf{d}_{i}’s satisfying 𝒫\mathcal{P} such that L2L1(β(N))L_{2}L_{1}(\beta^{\bullet}(N)) decompose into L2L1(β(Ni))L_{2}L_{1}(\beta^{\bullet}(N_{i})).

Proof.

We fix the degree sequence d0=(d0,0<d0,1<<d0,s)\textbf{d}_{0}=(d_{0,0}<d_{0,1}<\ldots<d_{0,s}). Let AA be the set of degree sequences {d=(d0<d1<<ds)|d0,0d0d1d0,1,d0,2d2dsd0,s}\{\textbf{d}=(d_{0}<d_{1}<\ldots<d_{s})|d_{0,0}\leq d_{0}\leq d_{1}\leq d_{0,1},d_{0,2}\leq d_{2}\leq d_{s}\leq d_{0,s}\}. Then AA is a finite set since d0,0,d0,1,d0,2,d0,sd_{0,0},d_{0,1},d_{0,2},d_{0,s} are fixed. If NN is pure of type d that does not satisfy 𝒫\mathcal{P}, then d satisfies the hypothesis of Proposition 4.11 so L2L1(β(N))L_{2}L_{1}(\beta^{\bullet}(N)) decomposes, and moreover, using the notation in Proposition 4.11, the two degree sequences d,d′′\textbf{d}^{\prime},\textbf{d}^{\prime\prime} are still in AA. Let CC^{\prime} be the cone generated by L2L1(β(N))L_{2}L_{1}(\beta^{\bullet}(N)), where NN is pure of type dA\textbf{d}\in A. Consider the set B={L2L1(β(N))|NB=\{L_{2}L_{1}(\beta^{\bullet}(N))|N pure of type d, dA\textbf{d}\in A satisfies 𝒫}\mathcal{P}\}. Then BB contains the vertex set because every element in C\BC^{\prime}\backslash B can be decomposed into elements in CC^{\prime}. We know CC^{\prime} is finitely generated as a cone, so a vertex set of CC^{\prime} also generates CC^{\prime} by Theorem 1.26 of [4], hence BB generates CC^{\prime}, and L2L1(β(N))CL_{2}L_{1}(\beta^{\bullet}(N))\in C. This means that L2L1(β(N))L_{2}L_{1}(\beta^{\bullet}(N)) decomposes into elements in BB, which proves the corollary. ∎

The next proposition shows that L2L1(β(N))L_{2}L_{1}(\beta^{\bullet}(N)) is decomposable if the length of the degree sequence s1,ns\neq 1,n. Note that we always assume 1sn1\leq s\leq n.

Proposition 4.13.

Let 2sn12\leq s\leq n-1. Let d=(d0<d1<<ds)\textbf{d}=(d_{0}<d_{1}<\ldots<d_{s}) be a degree sequence. Construct two degree sequences d’=(d0<d1<<ds1<ds+1)\textbf{d'}=(d_{0}<d_{1}<\ldots<d_{s-1}<d_{s}+1) and d”=(d0<d1<<ds1<ds<ds+1)\textbf{d''}=(d_{0}<d_{1}<\ldots<d_{s-1}<d_{s}<d_{s}+1). The first degree sequence has length s+1s+1 and the second degree sequence has length s+2s+2. Then c1πd(t)+c2πd’(t)=πd”(t)c_{1}\pi_{\textbf{d}}(t)+c_{2}\pi_{\textbf{d'}}(t)=\pi_{\textbf{d''}}(t) where c1>0c_{1}>0 and c2<0c_{2}<0. Moreover we also have αd”(t)=c1αd(t)+c2αd’(t)\alpha_{\textbf{d''}}(t)=c_{1}\alpha_{\textbf{d}}(t)+c_{2}\alpha_{\textbf{d'}}(t) and αd”(t)=c1αd(t)+c2αd’(t)\alpha_{\textbf{d''}}^{\prime}(t)=c_{1}\alpha_{\textbf{d}}^{\prime}(t)+c_{2}\alpha_{\textbf{d'}}^{\prime}(t). In particular, let NN be a pure module of type d, then L2L1(β(N))=c1L2L1(β(N))+c2L2L1(β(N′′))L_{2}L_{1}(\beta^{\bullet}(N))=c^{\prime}_{1}L_{2}L_{1}(\beta^{\bullet}(N^{\prime}))+c^{\prime}_{2}L_{2}L_{1}(\beta^{\bullet}(N^{\prime\prime})) where NN^{\prime} is pure of type d’ and N′′N^{\prime\prime} is pure of type d”, where c1,c2>0c^{\prime}_{1},c^{\prime}_{2}>0 are elements in \mathbb{Q}.

Proof.

Consider the \mathbb{Q}-vector space spanned by πd(t)\pi_{\textbf{d}}(t) and πd’(t)\pi_{\textbf{d'}}(t). The two polynomials are linearly independent because πd,ds0,πd’,ds=0,πd,ds+1=0,πd,ds0\pi_{\textbf{d},d_{s}}\neq 0,\pi_{\textbf{d'},d_{s}}=0,\pi_{\textbf{d},d_{s}+1}=0,\pi_{\textbf{d},d_{s}}\neq 0. So they span Vd”,sV_{\textbf{d''},s}. Also (1t)s+1|πd”(t)(1-t)^{s+1}|\pi_{\textbf{d''}}(t), hence there exist c1,c2c_{1},c_{2}\in\mathbb{Q} such that c1πd(t)+c2πd’(t)=πd”(t)c_{1}\pi_{\textbf{d}}(t)+c_{2}\pi_{\textbf{d'}}(t)=\pi_{\textbf{d''}}(t). Now sgn(c1)=sgn(πd,ds)/sgn(πd”,ds)=(1)s/(1)s=1sgn(c_{1})=sgn(\pi_{\textbf{d},d_{s}})/sgn(\pi_{\textbf{d''},d_{s}})=(-1)^{s}/(-1)^{s}=1 and sgn(c2)=sgn(πd’,ds+1)/sgn(πd”,ds+1)=(1)s/(1)s+1=1sgn(c_{2})=sgn(\pi_{\textbf{d'},d_{s}+1})/sgn(\pi_{\textbf{d''},d_{s}+1})=(-1)^{s}/(-1)^{s+1}=-1. Since s1s\neq 1, s2s\geq 2, by applying τd1\tau_{\leq d_{1}} and τd2\tau_{d_{2}} to this equation we get αd”(t)=c1αd(t)+c2αd’(t)\alpha_{\textbf{d''}}(t)=c_{1}\alpha_{\textbf{d}}(t)+c_{2}\alpha_{\textbf{d'}}(t) and αd”(t)=c1αd(t)+c2αd’(t)\alpha_{\textbf{d''}}^{\prime}(t)=c_{1}\alpha_{\textbf{d}}^{\prime}(t)+c_{2}\alpha_{\textbf{d'}}^{\prime}(t). By Proposition 4.10 this just means L2L1(β(N′′))=c1L2L1(β(N))+c2L2L1(β(N′′))L_{2}L_{1}(\beta^{\bullet}(N^{\prime\prime}))=c_{1}L_{2}L_{1}(\beta^{\bullet}(N))+c_{2}L_{2}L_{1}(\beta^{\bullet}(N^{\prime\prime})), therefore

L2L1(β(N))=c2c1L2L1(β(N))+1c1L2L1(β(N′′)).L_{2}L_{1}(\beta^{\bullet}(N))=-\frac{c_{2}}{c_{1}}L_{2}L_{1}(\beta^{\bullet}(N^{\prime}))+\frac{1}{c_{1}}L_{2}L_{1}(\beta^{\bullet}(N^{\prime\prime})).

Let c1=c2c1c^{\prime}_{1}=-\frac{c_{2}}{c_{1}} and c2=1c1c^{\prime}_{2}=\frac{1}{c_{1}}. Since c2<0<c1c_{2}<0<c_{1}, c1,c2c_{1},c_{2}\in\mathbb{Q}, we know c1,c2>0c^{\prime}_{1},c^{\prime}_{2}>0 and c1,c2c^{\prime}_{1},c^{\prime}_{2}\in\mathbb{Q}. ∎

Let CtC_{t} be the cone generated by the Ext-tables of modules of projective dimension at most 1; let CwfC_{wf} be the same as Corollary 4.7, that is, the cone generated by the Ext-tables of modules of projective dimension 1 without a free summand; let CfC_{f} be the cone generated by the Ext-tables of free modules. Then Ct=Cwf+CfC_{t}=C_{wf}+C_{f}. We want to know a generating set and the vertex set of CtC_{t}. To simplify the expressions, we introduce some more notations.

For a module NN of type d=d0<d1<<ds\textbf{d}=d_{0}<d_{1}<\ldots<d_{s}, denote L1(β(N))=adL_{1}(\beta^{\bullet}(N))=a_{\textbf{d}} and L2(ad)=bdL_{2}(a_{\textbf{d}})=b_{\textbf{d}}. If NN is pure of type d=d0<d1<<ds\textbf{d}=d_{0}<d_{1}<\ldots<d_{s} that does not satisfy the assumption in Proposition 4.11 or Proposition 4.13, then d must satisfy proposition 𝒫\mathcal{P} in Corollary 4.12 and is of length 2 or n+1n+1. So either s=1,d=d0<d0+1s=1,\textbf{d}=d_{0}<d_{0}+1, or s=n,d=d0,d0+1,d2,d2+1,,d2+n2s=n,\textbf{d}=d_{0},d_{0}+1,d_{2},d_{2}+1,\ldots,d_{2}+n-2. Consider the following 4 kinds of tables that are Ext-tables of some modules:

  1. (1)

    A1={L1(β(N))=ad=(1t)n(0,td0td0+1,0,,0)A_{1}=\{L_{1}(\beta^{\bullet}(N))=a_{\textbf{d}}=(1-t)^{-n}(0,t^{d_{0}}-t^{d_{0}+1},0,\ldots,0), where NN is pure of type d,\textbf{d}, d=(d0,d0+1)}\textbf{d}=(d_{0},d_{0}+1)\}.

  2. (2)

    A2={L1(β(N))=adA_{2}=\{L_{1}(\beta^{\bullet}(N))=a_{\textbf{d}}, where NN is pure of type d,\textbf{d}, d=(d0,d0+1,d2,d2+1,,d2+n2)}\textbf{d}=(d_{0},d_{0}+1,d_{2},d_{2}+1,\ldots,d_{2}+n-2)\}. In this case we have

    ad=(1t)n(i=2nπd,ditdi,i=0nπd,ditdi,0,,0).a_{\textbf{d}}=(1-t)^{-n}(\sum_{i=2}^{n}\pi_{\textbf{d},d_{i}}t^{d_{i}},\sum_{i=0}^{n}\pi_{\textbf{d},d_{i}}t^{d_{i}},0,\ldots,0).
  3. (3)

    A3={L1(β(N))=adA_{3}=\{L_{1}(\beta^{\bullet}(N))=a_{\textbf{d}}, where NN is pure of type d,\textbf{d}, d=(d0,d0+1,d2,d2+1,,d2+s2),2sn1}\textbf{d}=(d_{0},d_{0}+1,d_{2},d_{2}+1,\ldots,d_{2}+s-2),2\leq s\leq n-1\}. In this case we have

    ad=(1t)n(i=2sπd,ditdi,i=0sπd,ditdi,0,,0).a_{\textbf{d}}=(1-t)^{-n}(\sum_{i=2}^{s}\pi_{\textbf{d},d_{i}}t^{d_{i}},\sum_{i=0}^{s}\pi_{\textbf{d},d_{i}}t^{d_{i}},0,\ldots,0).
  4. (4)

    A4={E(R(d))=(1t)n(td,0,,0),d}A_{4}=\{E^{\bullet}(R(d))=(1-t)^{-n}(t^{d},0,\ldots,0),d\in\mathbb{Z}\}.

As a summary of the propositions above, we know:

Proposition 4.14.

(1) CwfC_{wf} is generated by A1A2A3A_{1}\cup A_{2}\cup A_{3}.

(2) CfC_{f} is generated by A4A_{4}.

(3) CtC_{t} is generated by A1A2A3A4A_{1}\cup A_{2}\cup A_{3}\cup A_{4}.

(4) An element in A3A_{3} decomposes into elements in A2A3A_{2}\cup A_{3}, so it cannot be a vertex.

Proof.

(1) This is proved by Corollary 4.7 and Corollary 4.12.

(2) This is true because every free module is a direct sum of free modules of rank 1.

(3) It is trivial by (1) and (2).

(4) This is proved by Proposition 4.13. ∎

Therefore, to find the vertex set of CtC_{t} it suffices to determine whether elements in A1A2A4A_{1}\cup A_{2}\cup A_{4} decompose into elements in A1A2A3A4A_{1}\cup A_{2}\cup A_{3}\cup A_{4} nontrivially.

Proposition 4.15.

The vertex set of CtC_{t} is A1A2A4A_{1}\cup A_{2}\cup A_{4}.

Proof.

Observe that only elements in A1A_{1} have a 0 entry in the first component and the elements in A2,A3A_{2},A_{3} and A4A_{4} have positive entries. So if an element in A1A_{1} decomposes, it can only decompose into elements in A1A_{1}, but elements in A1A_{1} are linearly independent, therefore, the decomposition is trivial. Similarly, checking the second component we know elements in A4A_{4} only have trivial decompositions. So it remains to check elements in A2A_{2}. We apply L2L_{2} again to the elements in A1,A2,A3A_{1},A_{2},A_{3} and A4A_{4}. We have:

L2((1t)n(0,td0td0+1,0,,0))=(td0td0+1,0,0,,0),L_{2}((1-t)^{-n}(0,t^{d_{0}}-t^{d_{0}+1},0,\ldots,0))=(t^{d_{0}}-t^{d_{0}+1},0,0,\ldots,0),
L2((1t)n(i=2nπd,ditdi,i=0nπd,ditdi,0,,0))=L_{2}((1-t)^{-n}(\sum_{i=2}^{n}\pi_{\textbf{d},d_{i}}t^{d_{i}},\sum_{i=0}^{n}\pi_{\textbf{d},d_{i}}t^{d_{i}},0,\ldots,0))=
(i=01πd,ditdi,i=2nπd,ditdi,0,,0),(\sum_{i=0}^{1}\pi_{\textbf{d},d_{i}}t^{d_{i}},\sum_{i=2}^{n}\pi_{\textbf{d},d_{i}}t^{d_{i}},0,\ldots,0),
L2((1t)n(i=2sπd,ditdi,i=0nπd,ditdi,0,,0))=L_{2}((1-t)^{-n}(\sum_{i=2}^{s}\pi_{\textbf{d},d_{i}}t^{d_{i}},\sum_{i=0}^{n}\pi_{\textbf{d},d_{i}}t^{d_{i}},0,\ldots,0))=
(i=01πd,ditdi,i=2sπd,ditdi,0,,0),(\sum_{i=0}^{1}\pi_{\textbf{d},d_{i}}t^{d_{i}},\sum_{i=2}^{s}\pi_{\textbf{d},d_{i}}t^{d_{i}},0,\ldots,0),
L2((1t)n(td,0,,0))=(td,td,0,,0).L_{2}((1-t)^{-n}(t^{d},0,\ldots,0))=(-t^{d},t^{d},0,\ldots,0).

The first 3 kinds of elements are also equal to L2(ad)=bdL_{2}(a_{\textbf{d}})=b_{\textbf{d}}. Now assume we have an equation

bd0=jJqjbdj+kKqkbdk+lLqlbdl+mMqm(tm,tm,0,,0).b_{\textbf{d}_{0}}=\sum_{j\in J}q_{j}b_{\textbf{d}_{j}}+\sum_{k\in K}q_{k}b_{\textbf{d}_{k}}+\sum_{l\in L}q_{l}b_{\textbf{d}_{l}}+\sum_{m\in M}q_{m}(-t^{m},t^{m},0,\ldots,0).

with bdj,bdk,bdlb_{\textbf{d}_{j}},b_{\textbf{d}_{k}},b_{\textbf{d}_{l}} belonging to L2(A1),L2(A2),L2(A3)L_{2}(A_{1}),L_{2}(A_{2}),L_{2}(A_{3}) respectively and qj,qk,ql,qmq_{j},q_{k},q_{l},q_{m} being positive rational numbers. We prove that this decomposition is trivial in the following steps.

(1) Observe the following fact: for each bdj,bdk,bdl,(tm,tm,0,,0)b_{\textbf{d}_{j}},b_{\textbf{d}_{k}},b_{\textbf{d}_{l}},(-t^{m},t^{m},0,\ldots,0), the lowest term of the second component has a positive coefficient. So let bmin2=min{dk,2,dl,2,m}b_{\min 2}=\min\{\textbf{d}_{k,2},\textbf{d}_{l,2},m\}, then bmin2=d0,2b_{\min 2}=\textbf{d}_{0,2}. In fact, if bmin2<d0,2b_{\min 2}<\textbf{d}_{0,2} then on the right side of (*) the coefficient of tbmin2t^{b_{\min 2}} in the second component is positive while on the left side it is 0. If bmin2>d0,2b_{\min 2}>\textbf{d}_{0,2} then on the left side of (*) the coefficient of td0,2t^{\textbf{d}_{0,2}} in the second component is positive while on the right side it is 0.

(2) Observe another fact: for each bdj,bdk,bdl,(tm,tm,0,,0)b_{\textbf{d}_{j}},b_{\textbf{d}_{k}},b_{\textbf{d}_{l}},(-t^{m},t^{m},0,\ldots,0), the highest term of the first component has a negative coefficient. Thus we can use the same method as in (1) to prove that if bmax1=max{dj,1,dk,1,dl,1,m}b_{\max 1}=\max\{\textbf{d}_{j,1},\textbf{d}_{k,1},\textbf{d}_{l,1},m\}, then bmax1=d0,1b_{\max 1}=\textbf{d}_{0,1}.

(3) For an integer mm, md0,1m\leq\textbf{d}_{0,1} and md0,2m\geq\textbf{d}_{0,2} are contradictory to each other because d0,2>d0,1\textbf{d}_{0,2}>\textbf{d}_{0,1}. So in (*) the term (tm,tm,0,,0)(-t^{m},t^{m},0,\ldots,0) cannot appear.

(4) For each bdj,bdk,bdlb_{\textbf{d}_{j}},b_{\textbf{d}_{k}},b_{\textbf{d}_{l}}, the lowest term of the first component has a positive coefficient. So let bmin1=min{dj,0,dk,0,dl,0}b_{\min 1}=\min\{\textbf{d}_{j,0},\textbf{d}_{k,0},\textbf{d}_{l,0}\}, then bmin1=d0,0b_{\min 1}=\textbf{d}_{0,0}.

(5) Observe the fact that d0,1=d0,0+1\textbf{d}_{0,1}=\textbf{d}_{0,0}+1, dj,1=dj,0+1\textbf{d}_{j,1}=\textbf{d}_{j,0}+1, dk,1=dk,0+1\textbf{d}_{k,1}=\textbf{d}_{k,0}+1, dl,1=dl,0+1\textbf{d}_{l,1}=\textbf{d}_{l,0}+1 for any j,k,lj,k,l. This, together with (2) and (4) implies that d0,0=dj,0=dk,0=dl,0\textbf{d}_{0,0}=\textbf{d}_{j,0}=\textbf{d}_{k,0}=\textbf{d}_{l,0} for any j,k,lj,k,l.

(6) Apply L21L_{2}^{-1} to (*) to get

ad0=jJqjadj+kKqkadk+lLqladl.a_{\textbf{d}_{0}}=\sum_{j\in J}q_{j}a_{\textbf{d}_{j}}+\sum_{k\in K}q_{k}a_{\textbf{d}_{k}}+\sum_{l\in L}q_{l}a_{\textbf{d}_{l}}.

The second entry of ada_{\textbf{d}} is (1t)nπd(t)(1-t)^{-n}\pi_{\textbf{d}}(t). If the length of d is ss, then the order of zero at t=1t=1 of πd(t)\pi_{\textbf{d}}(t) is ss, so the order of pole at t=1t=1 of (1t)nπd(t)(1-t)^{-n}\pi_{\textbf{d}}(t) is nsn-s; as 1sn1\leq s\leq n, (1t)nπd(t)(1-t)^{-n}\pi_{\textbf{d}}(t) is a Laurent polynomial if and only if n=sn=s, and if nsn\neq s, (1t)nπd(t)(1-t)^{-n}\pi_{\textbf{d}}(t) is the second entry of a multiple of an Ext-table of a module, hence all the coefficients are positive. So in (**) the term qkadkq_{k}a_{\textbf{d}_{k}} and qladlq_{l}a_{\textbf{d}_{l}} does not appear, otherwise the second exponent of the right side is a power series with infinitely many terms with positive coefficients, while the second exponent of the left side is a Laurent polynomial, which is a contradiction.

(7) We get that in (*),

bd0=jJqjbdj.b_{\textbf{d}_{0}}=\sum_{j\in J}q_{j}b_{\textbf{d}_{j}}.

All the elements are in A2A_{2}, so they are of the form bdj=dj,0<dj,0+1<d2,j<<d2,j+n2b_{\textbf{d}_{j}}=d_{j,0}<d_{j,0}+1<d_{2,j}<\ldots<d_{2,j}+n-2. Also by (5) all the dj,0d_{j,0} are equal to d0,0d_{0,0}. But in this case the second entry of bdjb_{\textbf{d}_{j}}, which is αdj(t)\alpha_{\textbf{d}_{j}}^{\prime}(t), only has nonzero entries in dj,2,,dj,2+n2d_{j,2},\ldots,d_{j,2}+n-2, so all these αdj(t)\alpha_{\textbf{d}_{j}}^{\prime}(t)’s are linearly independent, which implies that all the bdjb_{\textbf{d}_{j}}’s are linearly independent. Therefore, the decomposition is trivial. ∎

Smirnov and De Stefani, in [5], express each local cohomology table of finitely generated graded modules of dimension at most 2 as a finite positive linear combination of the vertices. However, this is not the case in projective dimension 1. When n>2n>2, A3A_{3}\neq\emptyset, and we have:

Proposition 4.16.

Any element in A3A_{3} is not a positive linear combination of elements in A1A2A4A_{1}\cup A_{2}\cup A_{4}.

Proof.

For elements in A4A_{4} the second component is 0. For elements in A1A_{1} the second component is (tdtd+1)/(1t)n(t^{d}-t^{d+1})/(1-t)^{n}. It has a pole at t=1t=1 of order n1n-1, and limt1(1t)n1(tdtd+1)/(1t)n=1>0\lim_{t\to 1}(1-t)^{n-1}(t^{d}-t^{d+1})/(1-t)^{n}=1>0. For elements in A4A_{4} the second component is πd(t)/(1t)n\pi_{\textbf{d}}(t)/(1-t)^{n} which is regular at t=1t=1. So for every linear combination of elements in A1A4A_{1}\cup A_{4} the second component is regular at t=1t=1; for every linear combination of elements in A1A2A4A_{1}\cup A_{2}\cup A_{4} where an element in A2A_{2} appears, the second component of this sum is a series f(t)f(t) which has a pole of order n1n-1 such that limt1(1t)n1f(t)>0\lim_{t\to 1}(1-t)^{n-1}f(t)>0. But for an element in A3A_{3} the second component has a pole at t=1t=1 of order nsn-s where 2sn12\leq s\leq n-1, so it cannot be a positive linear combination of elements in A1A2A4A_{1}\cup A_{2}\cup A_{4}. ∎

Proposition 4.14, 4.15 and 4.16 describe the cone of Ext-tables. By the local duality, they also give a description of the cone of local cohomology tables. Let Se=k[x1,,xe]S_{e}=k[x_{1},\ldots,x_{e}], Bd,e={(d0,d0+1),d0}{(d0,d0+1,d2,d2+1,,d2+e2),d0,d2,d2d0+2}B_{d,e}=\{(d_{0},d_{0}+1),d_{0}\in\mathbb{Z}\}\cup\{(d_{0},d_{0}+1,d_{2},d_{2}+1,\ldots,d_{2}+e-2),d_{0},d_{2}\in\mathbb{Z},d_{2}\geq d_{0}+2\}. Bd,e={(d0,d0+1,d2,d2+1,,d2+s2),d0,d2,d2d0+2,2se}B^{\prime}_{d,e}=\{(d_{0},d_{0}+1,d_{2},d_{2}+1,\ldots,d_{2}+s-2),d_{0},d_{2}\in\mathbb{Z},d_{2}\geq d_{0}+2,2\leq s\leq e\}, let BM,eB_{M,e} be the following set of SeS_{e}-modules and view them as RR-modules via projection π:RSR/(xe+1,,xn)\pi:R\to S\cong R/(x_{e+1},\ldots,x_{n}): MM is in BM,eB_{M,e} if and only if either it is free over SeS_{e} of rank 1, or when viewing MM as an SeS_{e}-module, its projective dimension is 1, does not have an SeS_{e}-summand, and Tr(M)\textup{Tr}(M) is pure of type d for dBd,e\textbf{d}\in B_{d,e}. Define BM,eB^{\prime}_{M,e} similarly as BM,eB_{M,e} where we replace Bd,eB_{d,e} by Bd,eB^{\prime}_{d,e}.

Theorem 4.17.

Let RR be a polynomial ring of dimension n2n\geq 2. Let CC be the cone of local cohomology tables of modules of projective dimension at most 11. Then:

(1) {H(M),MBM,n}\{H^{\bullet}(M),M\in B^{\prime}_{M,n}\} generates CC.

(2) {H(M),MBM,n}\{H^{\bullet}(M),M\in B_{M,n}\} is the vertex set of CC.

(3) If n>2n>2, not every element in CC is a positive linear combination of the extremal rays.

Theorem 4.18.

Let RR be a polynomial ring of dimension n3n\geq 3. Let CHC_{H} be the cone of local cohomology tables of all finitely generated graded RR-modules. Then CHC_{H} is not generated by its vertices.

Proof.

Let CC be the cone of local cohomology tables of modules of projective dimension at most 11. Then CCHC\subset C_{H}, and xCHx\in C_{H} is in CC if and only if all columns of xx vanish except for the last two columns which represent the (n1)(n-1)-th and nn-th local cohomology. All elements in CHC_{H} have nonnegative entries. Thus, if an element in CC decomposes in CHC_{H}, then this decomposition must lie in CC, so a vertex of CHC_{H} in CC is also a vertex of CC. Since n3n\geq 3, we can choose xCx\in C that does not decompose into vertices of CC. But any decomposition of xx in CHC_{H} also lies in CC, so xx does not decompose into vertices of CHC_{H}. ∎

Corollary 4.19.

Let ee be an integer with 2en2\leq e\leq n. Let CeC_{e} be the cone generated by local cohomology tables of modules MM with dimMe\dim M\leq e, depthMe1M\geq e-1. View SeS_{e}-modules as RR-modules via π\pi. Then:

(1) {H(M),MBM,e}\{H^{\bullet}(M),M\in B^{\prime}_{M,e}\} generates CeC_{e}.

(2) {H(M),MBM,e}\{H^{\bullet}(M),M\in B_{M,e}\} is the vertex set of CeC_{e}.

(3) If n>2n>2, not every element in CeC_{e} is a positive linear combination of the extremal rays.

Proof.

Use Theorem 4.17 and Lemma 3.7. ∎

Finally, we can describe the cone CsC_{s} of local cohomology tables of saCM modules; by Proposition 3.6 every local cohomology table of an saCM module decomposes into that of its dimension factors, which are almost Cohen-Macaulay, so Cs=i0nCiC_{s}=\sum^{n}_{i\geq 0}C_{i}. Let AM=2enBM,e{k(a),k[x](a),a}A^{\prime}_{M}=\cup_{2\leq e\leq n}B^{\prime}_{M,e}\cup\{k(a),k[x](a),a\in\mathbb{Z}\}, AM=2enBM,e{k(a),k[x](a),a}A_{M}=\cup_{2\leq e\leq n}B_{M,e}\cup\{k(a),k[x](a),a\in\mathbb{Z}\}. Note that the local cohomology table of SeS_{e} lies in both BM,eB_{M,e} and BM,e+1B_{M,e+1}; the reason is that viewing SeS_{e} as Se+1S_{e+1}-module it is pure with degree sequence d0=0<d1=1d_{0}=0<d_{1}=1 and its projective dimension is 1. We have the following description of CsC_{s}:

Corollary 4.20.

(1) {H(M),MAM}\{H^{\bullet}(M),M\in A^{\prime}_{M}\} generates CsC_{s}.

(2) {H(M),MAM}\{H^{\bullet}(M),M\in A_{M}\} is the vertex set of CsC_{s}.

(3) If n>2n>2, not every element in the cone CsC_{s} is a positive linear combination of the extremal rays.

Proof.

(1) is trivial; the union of the generating sets of cones generates the sum of the cones. For (2) and (3), we pick an element in the generating set {H(Mi),MiAM}\{H^{\bullet}(M_{i}),M_{i}\in A^{\prime}_{M}\}. If it is of the form k(a)k(a) or k[x](a),ak[x](a),a\in\mathbb{Z} then it is already extremal. If it is of the form H(Mi),MiBM,eH^{\bullet}(M_{i}),M_{i}\in B^{\prime}_{M,e}, then this local cohomology table only has two nonvanishing terms, that is, H𝔪e(Mi)H^{e}_{\mathfrak{m}}(M_{i}) and H𝔪e1(Mi)H^{e-1}_{\mathfrak{m}}(M_{i}). So if it decomposes into some tables of the form H(Mi),MiAMH^{\bullet}(M_{i}),M_{i}\in A^{\prime}_{M} then these tables also have zero local cohomologies except for the ee-th and the (e1)(e-1)-th local cohomology, which implies that MiBM,eM_{i}\in B^{\prime}_{M,e}. So a decomposition of a generator in CeC_{e} also lies in CeC_{e}, so a vertex of CsC_{s} which lies in CeC_{e} is also a vertex in CeC_{e}, which implies (2) and (3). ∎

5. The Γ\Gamma functor

Let MM be a finitely generated graded RR-module. Recall that the module of global sections of MM is

Γ(M)=tH0(Proj(R),M~(t)).\Gamma(M)=\oplus_{t\in\mathbb{Z}}H^{0}(\textup{Proj}(R),\tilde{M}(t)).

We can view Γ\Gamma as a functor from the category of graded RR-modules to itself. One might hope that it maps the category of finitely generated graded RR-modules to itself, but this is not true in general. However, if we focus on the problem of the decomposition of local cohomology tables and apply Lemma 3.3 and Lemma 3.4, we may always assume that 0ptM10ptM\geq 1, and the maximal submodule of dimension at most 1 is M1=0M_{1}=0. For such modules we have:

Proposition 5.1.

Let MM be a module, 0ptM10ptM\geq 1 and M1=0M_{1}=0. Then H𝔪1(M)H^{1}_{\mathfrak{m}}(M) has finite length.

Proof.

We may assume 0ptM=10ptM=1, otherwise H𝔪1(M)=0H^{1}_{\mathfrak{m}}(M)=0. The condition l(H𝔪1(M))<l(H^{1}_{\mathfrak{m}}(M))<\infty is equivalent to l(l(Ext(M,R)Rn1)<{}^{n-1}_{R}(M,R))<\infty by local duality. Since the module ExtRn1(M,R)\textup{Ext}^{n-1}_{R}(M,R) is finitely generated, this module has finite length if and only if Ext(M,R)𝔭Rn1=0{}^{n-1}_{R}(M,R)_{\mathfrak{p}}=0, \forall ht 𝔭=n1{\mathfrak{p}}=n-1, which just means Ext(M𝔭,R𝔭)R𝔭n1=0{}^{n-1}_{R_{\mathfrak{p}}}(M_{\mathfrak{p}},R_{\mathfrak{p}})=0, \forall ht 𝔭=n1{\mathfrak{p}}=n-1. Now the ring R𝔭R_{\mathfrak{p}} is a regular ring of dimension n1n-1, so apply the local duality on R𝔭R_{\mathfrak{p}} to get the equivalent condition H𝔭R𝔭0(M𝔭)=0H^{0}_{{\mathfrak{p}}R_{\mathfrak{p}}}(M_{\mathfrak{p}})=0, \forall ht 𝔭=n1\mathfrak{p}=n-1. Equivalently, 𝔭\mathfrak{p}\notinAss(M𝔭)(M_{\mathfrak{p}}), \forall ht 𝔭=n1{\mathfrak{p}}=n-1, or 𝔭\mathfrak{p}\notin AssMM, \forall ht 𝔭=n1\mathfrak{p}=n-1. This is true if and only if MM has no submodule of dimension 1. ∎

Below are some characterizations of the functor Γ\Gamma.

Proposition 5.2 (Universal property).

Denote the natural map MΓ(M)M\to\Gamma(M) by ii. Let MM and NN be two finitely generated graded RR-modules, where 0pt(M)10pt(M)\geq 1, 0pt(N)20pt(N)\geq 2. Suppose f:MNf:M\to N is an embedding, then there exists a unique embedding f:Γ(M)Nf^{\prime}:\Gamma(M)\to N such that f=fif=f^{\prime}i.

Proof.

Γ\Gamma is left exact because sheafification, tensoring with 𝒪n1(t)\mathcal{O}_{\mathbb{P}^{n-1}}(t) and H0H^{0} are all left exact. Also, when depthN2N\geq 2, Γ(N)=N\Gamma(N)=N. So an embedding of modules f:MNf:M\to N induces another embedding Γ(f):Γ(M)N\Gamma(f):\Gamma(M)\to N. Let Γ(f)=f\Gamma(f)=f^{\prime}. Suppose conversely we have f=fif=f^{\prime}i for some embedding f:Γ(M)Nf^{\prime}:\Gamma(M)\to N, then Γ(f)=Γ(f)Γ(i)\Gamma(f)=\Gamma(f^{\prime})\Gamma(i), but Γ(f)=f\Gamma(f^{\prime})=f^{\prime} and Γ(i)=idΓ(M)\Gamma(i)=\textup{id}_{\Gamma(M)}, hence f=Γ(f)f^{\prime}=\Gamma(f) is unique. ∎

Proposition 5.3.

Let MM and NN be two finitely generated graded RR-modules such that MM embeds into NN, 0pt(M)10pt(M)\geq 1 and 0pt(N)20pt(N)\geq 2. If l(N/M)<l(N/M)<\infty, then N=Γ(M)N=\Gamma(M).

Proof.

By the universal property Γ(M)\Gamma(M) embeds into NN. If it is not equal to NN, then by the depth lemma, N/Γ(M)N/\Gamma(M) has depth at least 1, hence l(N/Γ(M))=l(N/\Gamma(M))=\infty, hence l(N/M)=l(N/M)=\infty, which is a contradiction. ∎

Corollary 5.4.

Let MM and NN be two finitely generated graded RR-modules such that MM embeds into NN, 0pt(M)10pt(M)\geq 1 and 0pt(N)20pt(N)\geq 2. Let Msat=M:N𝔪M^{sat}=M:_{N}\mathfrak{m}^{\infty}. Then Γ(M)=Msat\Gamma(M)=M^{sat}.

Proof.

By construction, H𝔪0(N/M)=Msat/MH^{0}_{\mathfrak{m}}(N/M)=M^{sat}/M, (N/M)/H𝔪0(N/M)=N/Msat(N/M)/H^{0}_{\mathfrak{m}}(N/M)=N/M^{sat}, so 0pt(N/Msat)10pt(N/M^{sat})\geq 1. And 0pt(N)20pt(N)\geq 2, hence we can apply the depth lemma to get 0pt(Msat)20pt(M^{sat})\geq 2, and Msat/MM^{sat}/M is of finite length. By Proposition 5.3, Γ(M)=Msat\Gamma(M)=M^{sat}. ∎

Corollary 5.5.

Let MM be a finitely generated RR-module of depth 1 with no dimension 1 submodule. Then Γ(M)\Gamma(M) is also finitely generated.

Proof.

If 0pt(M)>00pt(M)>0, then H𝔪0(M)=0H^{0}_{\mathfrak{m}}(M)=0, so 0MΓ(M)H𝔪1(M)00\to M\to\Gamma(M)\to H^{1}_{\mathfrak{m}}(M)\to 0 is exact. Now MM is finitely generated, and H𝔪1(M)H^{1}_{\mathfrak{m}}(M) is of finite length by Proposition 5.1, hence H𝔪1(M)H^{1}_{\mathfrak{m}}(M) is also finitely generated, so Γ(M)\Gamma(M) is also finitely generated. ∎

In summary, to find a decomposition of H(M)H^{\bullet}(M) for a general module MM, we can consider two exact sequences 0H𝔪0(M)MM/H𝔪0(M)00\to H^{0}_{\mathfrak{m}}(M)\to M\to M/H^{0}_{\mathfrak{m}}(M)\to 0 and 0M1MM/M100\to M_{1}\to M\to M/M_{1}\to 0. The long exact sequences of local cohomology both have 0 connecting map, so they induce decompositions of the local cohomology table of MM. Hence, we can reduce to the case 0pt(M)=10pt(M)=1 where MM has no submodule of dimension 1. In this case by Corollary 5.5, Γ(M)\Gamma(M) is finitely generated, of depth at least 2 which contains MM such that H(M)H^{\bullet}(M) is equal to H(Γ(M))H^{\bullet}(\Gamma(M)) at position i2i\geq 2 and equal to the Hilbert function of Γ(M)/M\Gamma(M)/M at position 1 which has finite length. So we need to study the local cohomology tables of modules of depth at least 2, and their quotients of finite length.

6. Decomposition in dimension 3

In this section, we analyze whether the decomposition of H(Γ(M))H^{\bullet}(\Gamma(M)) induces that of H(M)H^{\bullet}(M). In dimension 2, this is the case, but things get complicated in dimension 3. From now on, we assume nn = 3, that is, RR is a polynomial ring over 3 variables. Let M,ΓM,\Gamma be two finitely generated graded RR-modules such that MΓM\subset\Gamma. Take another submodule FF of Γ\Gamma. Then we have a 3×33\times 3 exact diagram

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}MF\textstyle{M\cap F\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M/MF\textstyle{M/M\cap F\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F\textstyle{F\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Γ\textstyle{\Gamma\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Γ/F\textstyle{\Gamma/F\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F/MF\textstyle{F/M\cap F\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Γ/M\textstyle{\Gamma/M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Γ/(M+F)\textstyle{\Gamma/(M+F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}0\textstyle{0}0\textstyle{0}0\textstyle{0}

Diagram 2

This diagram induces 3 horizontal long exact sequences, 3 vertical long exact sequences, and 4 morphisms between these 6 exact sequences. Denote the 3 horizontal long exact sequences from top to bottom by C1,C2,C3C_{1},C_{2},C_{3} and 3 vertical ones from left to right D1,D2,D3D_{1},D_{2},D_{3}. The four morphism are f:C1C2f:C_{1}\to C_{2}, f:C2C3f^{\prime}:C_{2}\to C_{3}, g:D1D2g:D_{1}\to D_{2} and g:D2D3g^{\prime}:D_{2}\to D_{3}. These morphisms of complexes induces morphisms between long exact sequences which consist of 12 RR-linear maps, denoted by fi,fi,gi,gi,1i12f_{i},f^{\prime}_{i},g_{i},g^{\prime}_{i},1\leq i\leq 12. For example, fif_{i} is given by:

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪0(MF)\textstyle{H^{0}_{\mathfrak{m}}(M\cap F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f1\scriptstyle{f_{1}}H𝔪0(M)\textstyle{H^{0}_{\mathfrak{m}}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f2\scriptstyle{f_{2}}H𝔪0(M/MF)\textstyle{H^{0}_{\mathfrak{m}}(M/M\cap F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f3\scriptstyle{f_{3}}H𝔪1(MF)\textstyle{H^{1}_{\mathfrak{m}}(M\cap F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f4\scriptstyle{f_{4}}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪0(F)\textstyle{H^{0}_{\mathfrak{m}}(F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪0(Γ)\textstyle{H^{0}_{\mathfrak{m}}(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪0(Γ/F)\textstyle{H^{0}_{\mathfrak{m}}(\Gamma/F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪1(F)\textstyle{H^{1}_{\mathfrak{m}}(F)}H𝔪1(M)\textstyle{H^{1}_{\mathfrak{m}}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f5\scriptstyle{f_{5}}H𝔪1(M/MF)\textstyle{H^{1}_{\mathfrak{m}}(M/M\cap F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f6\scriptstyle{f_{6}}H𝔪2(MF)\textstyle{H^{2}_{\mathfrak{m}}(M\cap F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f7\scriptstyle{f_{7}}H𝔪2(M)\textstyle{H^{2}_{\mathfrak{m}}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f8\scriptstyle{f_{8}}H𝔪1(Γ)\textstyle{H^{1}_{\mathfrak{m}}(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪1(Γ/F)\textstyle{H^{1}_{\mathfrak{m}}(\Gamma/F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪2(F)\textstyle{H^{2}_{\mathfrak{m}}(F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪2(Γ)\textstyle{H^{2}_{\mathfrak{m}}(\Gamma)}
H𝔪2(M/MF)\textstyle{H^{2}_{\mathfrak{m}}(M/M\cap F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f9\scriptstyle{f_{9}}H𝔪3(MF)\textstyle{H^{3}_{\mathfrak{m}}(M\cap F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f10\scriptstyle{f_{10}}H𝔪3(M)\textstyle{H^{3}_{\mathfrak{m}}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f11\scriptstyle{f_{11}}H𝔪3(M/MF)\textstyle{H^{3}_{\mathfrak{m}}(M/M\cap F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f12\scriptstyle{f_{12}}0\textstyle{0}H𝔪2(Γ/F)\textstyle{H^{2}_{\mathfrak{m}}(\Gamma/F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪3(F)\textstyle{H^{3}_{\mathfrak{m}}(F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪3(Γ)\textstyle{H^{3}_{\mathfrak{m}}(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪3(Γ/F)\textstyle{H^{3}_{\mathfrak{m}}(\Gamma/F)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

We prove a general decomposition principle, which allows us to decompose H(M)H^{\bullet}(M) as the sum of local cohomology tables of a submodule and a quotient module of MM.

Proposition 6.1 (General decomposition principle).

Given a 3×33\times 3 diagram as above, and suppose it satisfies the following three conditions:

(1) depthF2F\geq 2, depthΓ/F2\Gamma/F\geq 2, depthΓ2\Gamma\geq 2.

(2) l(Γ/M)<l(\Gamma/M)<\infty, or equivalently, Γ=Γ(M)\Gamma=\Gamma(M).

(3) The connecting homomorphism H𝔪2(Γ/F)H𝔪3(F)H^{2}_{\mathfrak{m}}(\Gamma/F)\to H^{3}_{\mathfrak{m}}(F) is 0.

Then the following three properties hold.

(4) H(Γ)=H(F)+H(Γ/F)H^{\bullet}(\Gamma)=H^{\bullet}(F)+H^{\bullet}(\Gamma/F).

(5) F=Γ(MF)F=\Gamma(M\cap F) and Γ/F=Γ(M/MF)\Gamma/F=\Gamma(M/M\cap F).

(6) H(M)=H(MF)+H(M/MF)H^{\bullet}(M)=H^{\bullet}(M\cap F)+H^{\bullet}(M/M\cap F).

Proof.

(4) Consider the long exact sequence induced by C2C_{2}. By condition (3), it decomposes into 2 short exact sequences, hence the equality holds.

(5) By the short exact sequence at the bottom, Γ/M\Gamma/M is of finite length implies that F/MFF/M\cap F and Γ/M+F\Gamma/M+F are of finite length. So by condition (1), (5) holds.

(6) By (4) it suffices to decompose H𝔪1H^{1}_{\mathfrak{m}}. But by (5), H𝔪1(M)=Γ/MH^{1}_{\mathfrak{m}}(M)=\Gamma/M, H𝔪1(MF)=F/MFH^{1}_{\mathfrak{m}}(M\cap F)=F/M\cap F, H𝔪1(M/MF)=Γ/(M+F)H^{1}_{\mathfrak{m}}(M/M\cap F)=\Gamma/(M+F). So by the short exact sequence C3C_{3} at the bottom, H𝔪1H^{1}_{\mathfrak{m}} also decomposes. ∎

The general decomposition principle tells us that if we can find a submodule FF satisfying (1)-(3), then H(M)H^{\bullet}(M) decomposes.

Corollary 6.2.

We assume (2) of 6.1 holds in diagram 22. If Γ\Gamma is decomposable, then H(M)H^{\bullet}(M) decomposes.

Proof.

Suppose FF is a direct summand, then Γ=FΓ/F\Gamma=F\oplus\Gamma/F. In this case the condition (1) and (3) is satisfied. ∎

Corollary 6.3.

We assume (2) of 6.1 holds in diagram 22. If H𝔪2(M)=0H^{2}_{\mathfrak{m}}(M)=0 or H𝔪3(M)=0H^{3}_{\mathfrak{m}}(M)=0, then H(M)H^{\bullet}(M) decomposes as a sum of H(Mi)H^{\bullet}(M_{i}) such that Γ(Mi)\Gamma(M_{i}) is cyclic.

Proof.

If H𝔪3(M)=0H^{3}_{\mathfrak{m}}(M)=0, then Γ(M)\Gamma(M) is Cohen-Macaulay of dimension 2, then it reduces to the case in [5]. If H𝔪2(M)=0H^{2}_{\mathfrak{m}}(M)=0, then H𝔪i(Γ)H^{i}_{\mathfrak{m}}(\Gamma) is zero except for i=3i=3, so Γ\Gamma is free, hence it is a direct sum of free cyclic modules, so (1) and (3) hold. And (2) holds by assumption, so we get (4)-(6). Using (5) and (6) repeatedly, we get H(M)H^{\bullet}(M) decomposes into H(Mi)H^{\bullet}(M_{i}) such that Γ(Mi)\Gamma(M_{i}) is an indecomposable direct summand of Γ\Gamma, hence must be free cyclic. ∎

Remark 6.4.

The above two lemmas explain why things are different in dimension 2 and dimension 3. In dimension 2, Γ\Gamma is free because depthΓ2\Gamma\geq 2, so it reduces to the case where Γ\Gamma is free of rank 1. In dimension 3 this is not always true.

Pick a module MM of depth 1 without a dimension 1 submodule and let Γ=Γ(M)\Gamma=\Gamma(M). In general it is hard to find a submodule that satisfies both (1) and (3). There are two ways to approach this. The first way is to go modulo a submodule of dimension 2; and the second way is to go modulo a free submodule. In the first way, (3) is satisfied but the quotient Γ/F\Gamma/F may violate (1) because we may have depthΓ/F=1\Gamma/F=1.

Lemma 6.5.

Let Γ\Gamma be a module of depth 22. Then the maximal submodule of dimension at most 22 is the torsion submodule Tor(Γ)\textup{Tor}(\Gamma). If depthΓ2\Gamma\geq 2, then Tor(Γ)\textup{Tor}(\Gamma) is Cohen-Macaulay of dimension 22.

Proof.

The maximal submodule of dimension at most 2 is generated by all mΓm\in\Gamma such that ann(m)R0{}_{R}(m)\neq 0, so it is Tor(Γ)\textup{Tor}(\Gamma). We have an exact sequence 0Tor(Γ)ΓQ00\to\textup{Tor}(\Gamma)\to\Gamma\to Q\to 0 where QQ is the quotient module. By assumption depth(Γ\Gamma)\geq 2. Also, H𝔪0(Q)=0H^{0}_{\mathfrak{m}}(Q)=0 because QQ does not have submodule of dimension less than 2, so depth(QQ)\geq 1. Therefore we have depth(TorΓ)2\textup{Tor}\Gamma)\geq 2 by the depth lemma, but dim(TorΓ)2\textup{Tor}\Gamma)\leq 2, so it is Cohen-Macaulay of dimension 2. ∎

The measurement of Γ/F\Gamma/F violating (1) is given by the module H𝔪1(Q)H^{1}_{\mathfrak{m}}(Q). It suffices to give a description of this module.

Lemma 6.6.

Let Γ\Gamma be a module of depth 22, then Γ\Gamma^{*} and Γ\Gamma^{**} are modules of depth at least 22.

Proof.

If Γ\Gamma has a free summand FF, then FF^{*} is a free summand of Γ\Gamma^{*}, so it suffices to prove in the case where Γ\Gamma has no free summand. In this case Γ\Gamma^{*} is the second syzygy of Tr(Γ)\textup{Tr}(\Gamma). Hence projdim(Γ)32=1(\Gamma^{*})\leq 3-2=1. So depth(Γ)2(\Gamma^{*})\geq 2. Replace Γ\Gamma by Γ\Gamma^{*}, we get depth(Γ)2(\Gamma^{**})\geq 2. ∎

Proposition 6.7.

There is an exact sequence 0Tor(Γ)ΓΓL00\to\textup{Tor}(\Gamma)\to\Gamma\to\Gamma^{**}\to L\to 0. Let N=Tr(Γ)N=\textup{Tr}(\Gamma), then L=ExtR2(N,R)L=\textup{Ext}^{2}_{R}(N,R). Let Q=Γ/TorΓQ=\Gamma/\textup{Tor}\Gamma, then QQ has depth at least 11 and H𝔪1(Q)=H𝔪0(L)H^{1}_{\mathfrak{m}}(Q)=H^{0}_{\mathfrak{m}}(L) is of finite length.

Proof.

The exact sequence is well-known, so we omit the proof of the exactness. Now let Q=Γ/TorΓQ=\Gamma/\textup{Tor}\Gamma. The composition map Tor(Γ)ΓΓ\textup{Tor}(\Gamma)\to\Gamma\to\Gamma^{**} is 0, hence we have a map QΓQ\to\Gamma^{**}. This map is injective because it is a map between torsion-free modules over RR and it stays injective after tensoring with KK. So there are two short exact sequences 0Tor(Γ)ΓQ00\to\textup{Tor}(\Gamma)\to\Gamma\to Q\to 0 and 0QΓL00\to Q\to\Gamma^{**}\to L\to 0. This leads to two long exact sequences:

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪1Tor(Γ)=0\textstyle{H^{1}_{\mathfrak{m}}\textup{Tor}(\Gamma)=0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪1(Γ)=0\textstyle{H^{1}_{\mathfrak{m}}(\Gamma)=0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪1(Q)\textstyle{H^{1}_{\mathfrak{m}}(Q)}H𝔪2Tor(Γ)\textstyle{H^{2}_{\mathfrak{m}}\textup{Tor}(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪2(Γ)\textstyle{H^{2}_{\mathfrak{m}}(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪2(Q)\textstyle{H^{2}_{\mathfrak{m}}(Q)}H𝔪3Tor(Γ)\textstyle{H^{3}_{\mathfrak{m}}\textup{Tor}(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪3(Γ)\textstyle{H^{3}_{\mathfrak{m}}(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪3(Q)\textstyle{H^{3}_{\mathfrak{m}}(Q)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0,\textstyle{0,}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪0(Q)=0\textstyle{H^{0}_{\mathfrak{m}}(Q)=0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪0(Γ)=0\textstyle{H^{0}_{\mathfrak{m}}(\Gamma^{**})=0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪0(L)\textstyle{H^{0}_{\mathfrak{m}}(L)}H𝔪1(Q)\textstyle{H^{1}_{\mathfrak{m}}(Q)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪1(Γ)=0\textstyle{H^{1}_{\mathfrak{m}}(\Gamma^{**})=0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪1(L)\textstyle{H^{1}_{\mathfrak{m}}(L)}H𝔪2(Q)\textstyle{H^{2}_{\mathfrak{m}}(Q)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪2(Γ)\textstyle{H^{2}_{\mathfrak{m}}(\Gamma^{**})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪2(L)=0\textstyle{H^{2}_{\mathfrak{m}}(L)=0}H𝔪3(Q)\textstyle{H^{3}_{\mathfrak{m}}(Q)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪3(Γ)\textstyle{H^{3}_{\mathfrak{m}}(\Gamma^{**})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪3(L)=0\textstyle{H^{3}_{\mathfrak{m}}(L)=0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

As in the proof of Lemma 6.5, depth(Q)(Q)\geq 1. Also depth(Γ)(\Gamma^{**})\geq 2, so H𝔪0(Γ)=H𝔪1(Γ)=0H^{0}_{\mathfrak{m}}(\Gamma^{**})=H^{1}_{\mathfrak{m}}(\Gamma^{**})=0, hence H𝔪1(Q)H𝔪0(L)H^{1}_{\mathfrak{m}}(Q)\cong H^{0}_{\mathfrak{m}}(L). Note that LL is a finitely generated module over RR, so H𝔪0(L)H^{0}_{\mathfrak{m}}(L) is of finite length. ∎

Theorem 6.8.

Let MM be a module of depth 11 and assume MM has no dimension 11 submodule. Let Γ=Γ(M)\Gamma=\Gamma(M). Then H(M)H^{\bullet}(M)=H(TorM)+H(M/TorM)(0,HS(H𝔪1(Q)),HS(H𝔪1(Q)),0)H^{\bullet}(\textup{Tor}M)+H^{\bullet}(M/\textup{Tor}M)-(0,HS(H^{1}_{\mathfrak{m}}(Q)),HS(H^{1}_{\mathfrak{m}}(Q)),0). In particular, if H𝔪1(Q)=0H^{1}_{\mathfrak{m}}(Q)=0, then H(M)=H^{\bullet}(M)=

H(TorM)H^{\bullet}(\textup{Tor}M)+H(M/TorM)+H^{\bullet}(M/\textup{Tor}M).

Proof.

Take F=TorΓF=\textup{Tor}\Gamma in diagram 2, then MF=Tor(M)M\cap F=\textup{Tor}(M), Γ/F=Q\Gamma/F=Q. Let f:C1C2f:C_{1}\to C_{2} be the corresponding morphism. Then fi,7i12f_{i},7\leq i\leq 12 are isomorphisms, and f6f_{6} is surjective. Note that in this case H𝔪3(F)=H𝔪3(MF)=0H^{3}_{\mathfrak{m}}(F)=H^{3}_{\mathfrak{m}}(M\cap F)=0. Now eliminate all the 0’s and get a diagram:

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪1(TorM)\textstyle{H^{1}_{\mathfrak{m}}(\textup{Tor}M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f4\scriptstyle{f_{4}}H𝔪1(M)\textstyle{H^{1}_{\mathfrak{m}}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f5\scriptstyle{f_{5}}H𝔪1(M/TorM)\textstyle{H^{1}_{\mathfrak{m}}(M/\textup{Tor}M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f6\scriptstyle{f_{6}}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪1(Q)\textstyle{H^{1}_{\mathfrak{m}}(Q)}H𝔪2(TorM)\textstyle{H^{2}_{\mathfrak{m}}(\textup{Tor}M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f7\scriptstyle{f_{7}}H𝔪2(M)\textstyle{H^{2}_{\mathfrak{m}}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f8\scriptstyle{f_{8}}H𝔪2(M/TorM)\textstyle{H^{2}_{\mathfrak{m}}(M/\textup{Tor}M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f9\scriptstyle{f_{9}}0\textstyle{0}H𝔪2(TorΓ)\textstyle{H^{2}_{\mathfrak{m}}(\textup{Tor}\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪2(Γ)\textstyle{H^{2}_{\mathfrak{m}}(\Gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪2(Q)\textstyle{H^{2}_{\mathfrak{m}}(Q)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

plus isomorphisms H𝔪3(M)=H𝔪3(Γ)=H𝔪3(M/TorM)=H𝔪3(Q)H^{3}_{\mathfrak{m}}(M)=H^{3}_{\mathfrak{m}}(\Gamma)=H^{3}_{\mathfrak{m}}(M/\textup{Tor}M)=H^{3}_{\mathfrak{m}}(Q). So the kernel of the map H𝔪2(TorM)H𝔪2(M)H^{2}_{\mathfrak{m}}(\textup{Tor}M)\to H^{2}_{\mathfrak{m}}(M) is isomorphic to H𝔪1(Q)H^{1}_{\mathfrak{m}}(Q). This leads to 3 equations:

HS(H𝔪3(M))=HS(H𝔪3(TorM)),HS(H^{3}_{\mathfrak{m}}(M))=HS(H^{3}_{\mathfrak{m}}(\textup{Tor}M)),
HS(H𝔪2(M))=HS(H𝔪2(M/TorM))+HS(H𝔪2(TorM))HS(H𝔪1(Q)),HS(H^{2}_{\mathfrak{m}}(M))=HS(H^{2}_{\mathfrak{m}}(M/\textup{Tor}M))+HS(H^{2}_{\mathfrak{m}}(\textup{Tor}M))-HS(H^{1}_{\mathfrak{m}}(Q)),
HS(H𝔪1(M))=HS(H𝔪1(M/TorM))+HS(H𝔪1(TorM))HS(H𝔪1(Q)).HS(H^{1}_{\mathfrak{m}}(M))=HS(H^{1}_{\mathfrak{m}}(M/\textup{Tor}M))+HS(H^{1}_{\mathfrak{m}}(\textup{Tor}M))-HS(H^{1}_{\mathfrak{m}}(Q)).

Equivalently, we have

H(M)=H(TorM)+H(M/TorM)e,H^{\bullet}(M)=H^{\bullet}(\textup{Tor}M)+H^{\bullet}(M/\textup{Tor}M)-e,

where e=(0,HS(H𝔪1(Q)),HS(H𝔪1(Q)),0).e=(0,HS(H^{1}_{\mathfrak{m}}(Q)),HS(H^{1}_{\mathfrak{m}}(Q)),0).

The above theorem shows that H𝔪1(Q)H^{1}_{\mathfrak{m}}(Q) is the error term we want to get rid of in the sense that if it vanishes then H(M)H^{\bullet}(M) decomposes as a sum of two local cohomology tables H(TorM)H^{\bullet}(\textup{Tor}M) and H(M/TorM)H^{\bullet}(M/\textup{Tor}M). Here TorM\textup{Tor}M is a submodule of MM of dimension 2. In general H𝔪1(Q)H^{1}_{\mathfrak{m}}(Q) does not vanish. By Proposition 6.7 we see H𝔪1(Q)H𝔪0(L)H^{1}_{\mathfrak{m}}(Q)\cong H^{0}_{\mathfrak{m}}(L), and the next two proposition shows when it vanishes and how to calculate it in terms of MM.

Proposition 6.9.

Using the same notation as Theorem 6.8, let L=ExtR2(Tr(Γ),R)L=\textup{Ext}^{2}_{R}(\textup{Tr}(\Gamma),R). Then H𝔪1(Q)=0H^{1}_{\mathfrak{m}}(Q)=0 if and only L=0L=0 or LL is Cohen-Macaulay of dimension 11.

Proof.

For a finitely generated module Γ\Gamma, dimL32=1L\leq 3-2=1. So if L0L\neq 0, then H𝔪0(L)=0H^{0}_{\mathfrak{m}}(L)=0 if and only if depthLL\geq 1, if and only if LL is a Cohen-Macaulay module of dimension 1. ∎

Proposition 6.10.

Let Γ\Gamma be a finitely generated module over RR without a free summand. Let N=Tr(Γ)N=\textup{Tr}(\Gamma), N1N_{1} be the maximal submodule of dimension at most 11, Q=N/N1Q^{\prime}=N/N_{1}, and L=ExtR2(N,R)L=\textup{Ext}^{2}_{R}(N,R). Let Γ=Γ(Q)\Gamma^{\prime}=\Gamma(Q^{\prime}). Then H𝔪0(L)H^{0}_{\mathfrak{m}}(L) = Hom(Γ/Q,E)R{}_{R}(\Gamma^{\prime}/Q^{\prime},E).

Proof.

We may assume depthNN\geq 1 because Ext(N,R)R2={}^{2}_{R}(N,R)= Ext(N/H𝔪0(N),R)R2{}^{2}_{R}(N/H^{0}_{\mathfrak{m}}(N),R). In this case N1N_{1} is Cohen-Macaulay of dimension 1. The short exact sequence 0N1NQ00\to N_{1}\to N\to Q^{\prime}\to 0 induces a long exact sequence:

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪0(N1)\textstyle{H^{0}_{\mathfrak{m}}(N_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪0(N)\textstyle{H^{0}_{\mathfrak{m}}(N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪0(Q)=0\textstyle{H^{0}_{\mathfrak{m}}(Q^{\prime})=0}H𝔪1(N1)\textstyle{H^{1}_{\mathfrak{m}}(N_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪1(N)\textstyle{H^{1}_{\mathfrak{m}}(N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪1(Q)\textstyle{H^{1}_{\mathfrak{m}}(Q^{\prime})}H𝔪2(N1)=0\textstyle{H^{2}_{\mathfrak{m}}(N_{1})=0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪2(N)\textstyle{H^{2}_{\mathfrak{m}}(N)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H𝔪2(Q).\textstyle{H^{2}_{\mathfrak{m}}(Q^{\prime}).}

So we have an exact sequence 0H𝔪1(N1)H𝔪1(N)H𝔪1(Q)00\to H^{1}_{\mathfrak{m}}(N_{1})\to H^{1}_{\mathfrak{m}}(N)\to H^{1}_{\mathfrak{m}}(Q^{\prime})\to 0. By local duality, 00\to Ext(Q,R)R2L{}^{2}_{R}(Q^{\prime},R)\to L\to Ext(N1,R)R20{}^{2}_{R}(N_{1},R)\to 0 is exact. Now dimL1L\leq 1, Ext(N1,R)R2{}^{2}_{R}(N_{1},R) is Cohen-Macaulay of dimension 1, and H𝔪1(Q)H^{1}_{\mathfrak{m}}(Q^{\prime}) is of finite length, hence Ext(Q,R)R2{}^{2}_{R}(Q^{\prime},R) is of finite length. This means that Ext(Q,R)R2=H𝔪0(L){}^{2}_{R}(Q^{\prime},R)=H^{0}_{\mathfrak{m}}(L). Finally Ext(Q,R)R2={}^{2}_{R}(Q^{\prime},R)= Hom(H𝔪1(Q),E)R{}_{R}(H^{1}_{\mathfrak{m}}(Q^{\prime}),E) and H𝔪1(Q)=Γ/QH^{1}_{\mathfrak{m}}(Q^{\prime})=\Gamma^{\prime}/Q^{\prime}, so we are done. ∎

Theorem 6.8 describes a way to decompose a local cohomology table of a module MM of dimension 3 using a submodule TorM\textup{Tor}M of dimension 2. There is another way to decompose H(M)H^{\bullet}(M), which is induced by a free submodule FΓF\subset\Gamma. Note that if Γ\Gamma does not have a free submodule, then dimΓ2\dim\Gamma\leq 2, dimM2\dim M\leq 2, and the decomposition of the local cohomology table of MM is known by [5]. So we may always assume that Γ\Gamma has a free submodule. In this case (1) is automatically satisfied because the depth lemma implies that if depthΓ2\Gamma\leq 2 and depthF3F\leq 3, then depthΓ/F2\Gamma/F\geq 2. Hence FF may only violate (3) in the general decomposition principle. We observe that H𝔪2(Γ/F)H𝔪3(F)H^{2}_{\mathfrak{m}}(\Gamma/F)\to H^{3}_{\mathfrak{m}}(F) is 0 if and only if H𝔪3(F)H𝔪3(Γ)H^{3}_{\mathfrak{m}}(F)\to H^{3}_{\mathfrak{m}}(\Gamma) is injective, if and only if ΓF\Gamma^{*}\to F^{*} is surjective.

Lemma 6.11.

In the 3×33\times 3 diagram at the beginning of this section, let F=Re,eΓF=Re,e\in\Gamma be a free submodule. Then (3) of Proposition 5.1 is equivalent to the following condition: there exists hΓh\in\Gamma^{*} such that h(e)=1h(e)=1.

Proof.

Since FF is a free cyclic module generated by ee, FF^{*} is free cyclic and generated by ee^{*}. So the map ΓF\Gamma^{*}\to F^{*} is surjective if and only if ee^{*} is in the image. Moreover, h(e)=1h(e)=1 if and only if the image of hh is ee^{*}. ∎

Proposition 6.12.

Let Γ\Gamma be a module of dimension 33 and depth at least 22. Suppose Γ\Gamma^{*} has a free summand G=ReG=Re^{\prime} and L=ExtR2(Tr(Γ),R)=0L=\textup{Ext}^{2}_{R}(\textup{Tr}(\Gamma),R)=0. Then F=GΓF=G^{*}\subset\Gamma is a free summand, and the pair (F,Γ)(F,\Gamma) satisfies condition (33) of Proposition 6.1.

Proof.

Since FF is a free summand of Γ\Gamma^{**}, Γ\Gamma^{**} surjects onto FF. Now since L=0L=0, Γ\Gamma surjects onto Γ\Gamma^{**}, so Γ\Gamma surjects onto FF, but FF is projective, hence FF is a free summand. The map ΓF\Gamma^{*}\to F^{*} induced by the inclusion FΓF\to\Gamma is just the projection onto the summand GG which it is surjective. This means that the pair (F,Γ)(F,\Gamma) satisfies (3) of Proposition 6.1. ∎

Proposition 6.13.

Let Γ\Gamma be a module of dimension 33 and depth at least 22. Then Γ\Gamma^{*} does not have a free summand if and only if Γ=Tr(L)\Gamma^{*}=\textup{Tr}(L^{\prime}) for a module LL^{\prime} of finite length.

Proof.

By the previous proposition projdim(Γ)1(\Gamma^{*})\leq 1. So Γ\Gamma^{*} does not have a free summand if and only if Γ=Tr(L)\Gamma^{*}=\textup{Tr}(L^{\prime}) for L=Tr(Γ)L^{\prime}=\textup{Tr}(\Gamma^{*}). But Γ\Gamma^{*} is the second syzygy of N=Tr(Γ)N=\textup{Tr}(\Gamma), hence L=ExtR3(N,R)L^{\prime}=\textup{Ext}^{3}_{R}(N,R), and this module has finite length. ∎

By Proposition 6.12 and 6.13 we immediately have:

Corollary 6.14.

We keep the same notation from Proposition 6.12 and 6.13. Suppose Γ=Γ(M)\Gamma=\Gamma(M), L=0L=0, and ΓTr(L)\Gamma^{*}\neq Tr(L^{\prime}) for any module LL^{\prime} of finite length. Then there exists a free submodule FΓF\subset\Gamma such that H(M)=H(MF)+H(M/MF)H^{\bullet}(M)=H^{\bullet}(M\cap F)+H^{\bullet}(M/M\cap F).

Since l(F/MF)<l(F/M\cap F)<\infty, we know dim(MF)(M\cap F) = dimFF = 3, so H(M)H^{\bullet}(M) is the sum of H(MF)H^{\bullet}(M\cap F) and H(M/MF)H^{\bullet}(M/M\cap F) where MFM\cap F is a submodule of MM of dimension 3.

In conclusion, for a finitely generated graded module MM of dimension 3, H(M)H^{\bullet}(M) is decomposable in two cases; in Theorem 6.8 a submodule of MM of dimension 2 induces a decomposition and in Corollary 6.14 a submodule of dimension 3 induces a decomposition.

Acknowledgements

The author would like to thank Giulio Caviglia for introducing this problem and providing references. The author is supported by the Ross-Lynn Research Scholar Fund of Purdue University.

References

  • [1] Maurice Auslander and Mark Bridger. Stable module theory. Number 94. American Mathematical Soc., 1969.
  • [2] Mats Boij and Jonas Söderberg. Graded betti numbers of cohen–macaulay modules and the multiplicity conjecture. Journal of the London Mathematical Society, 78(1):85–106, 2008.
  • [3] Mats Boij and Jonas Söderberg. Betti numbers of graded modules and the multiplicity conjecture in the non-cohen–macaulay case. Algebra & Number Theory, 6(3):437–454, 2012.
  • [4] Winfried Bruns and Joseph Gubeladze. Polytopes, rings, and k-theory. 2009.
  • [5] Alessandro De Stefani and Ilya Smirnov. Decomposition of graded local cohomology tables. Mathematische Zeitschrift, 297(1):1–24, 2021.
  • [6] David Eisenbud, Gunnar Fløystad, and Jerzy Weyman. The existence of equivariant pure free resolutions. In Annales de l’Institut Fourier, volume 61, pages 905–926, 2011.
  • [7] David Eisenbud and Frank-Olaf Schreyer. Betti numbers of graded modules and cohomology of vector bundles. Journal of the American Mathematical Society, 22(3):859–888, 2009.
  • [8] David Eisenbud and Frank-Olaf Schreyer. Cohomology of coherent sheaves and series of supernatural bundles. Journal of the European Mathematical Society, 12(3):703–722, 2010.
  • [9] Peter Schenzel. On the dimension filtration and cohen-macaulay filtered modules. Lecture Notes in Pure and Applied Mathematics, pages 245–264, 1999.