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

Betti numbers for connected sums of graded Gorenstein Artinian algebras

Nasrin Altafi Nasrin Altafi: Department of Mathematics, KTH Royal Institute of Technology, S-100 44 Stockholm, Sweden and Department of Mathematics, Queen’s University, 505 Jeffery Hall, University Avenue, Queen’s University, Kingston, Ontario, Canada K7L 3N6 [email protected] Roberta Di Gennaro Roberta Di Gennaro: Dipartimento di Matematica e Applicazioni “Renato Caccioppoli”, Complesso Universitario Monte Sant’Angelo, Università degli Studi di Napoli Federico II, Via Cinthia 80126 Napoli, Italy [email protected] Federico Galetto Federico Galetto: Department of Mathematics and Statistics, Cleveland State University, 2121 Euclid Avenue, RT 1515 Cleveland, OH 44115-2215, USA [email protected] Sean Grate Sean Grate: Department of Mathematics and Statistics, Auburn University, 221 Parker Hall, Auburn, AL 36849, USA [email protected] Rosa M. Miró-Roig Rosa Maria Miró-Roig: Facultat de Matemàtiques i Informàtica, Universitat de Barcelona, Gran Via des les Corts Catalanes 585, 08007 Barcelona, Spain [email protected], ORCID 0000-0003-1375-6547 Uwe Nagel Uwe Nagel: Department of Mathematics, University of Kentucky, 715 Patterson Office Tower, Lexington, KY 40506-0027, USA [email protected] Alexandra Seceleanu Alexandra Seceleanu: Department of Mathematics, University of Nebraska-Lincoln, 203 Avery Hall, Lincoln, NE 68588, USA [email protected]  and  Junzo Watanabe Department of Mathematics Tokai University, Hiratsuka, Kanagawa 259–1292, Japan [email protected]
Abstract.

The connected sum construction, which takes as input Gorenstein rings and produces new Gorenstein rings, can be considered as an algebraic analogue for the topological construction having the same name. We determine the graded Betti numbers for connected sums of graded Artinian Gorenstein algebras. Along the way, we find the graded Betti numbers for fiber products of graded rings; an analogous result was obtained in the local case by Geller [Gel22]. We relate the connected sum construction to the doubling construction, which also produces Gorenstein rings. Specifically, we show that a connected sum of doublings is the doubling of a fiber product ring.

Altafi was supported by Swedish Research Council grant VR2021-00472, Galetto was supported by NSF DMS–2200844, Miró-Roig was partially supported by the grant PID2020-113674GB-I00, Nagel was partially supported by Simons Foundation grant #636513, Seceleanu was supported by NSF DMS–2101225.

1. Introduction

The connected sum is a topological construction that takes two manifolds to produce a new manifold [Mas91, p. 7]. An algebraic analog of this surgery construction was introduced by H. Ananthnarayan, L. Avramov, and W.F. Moore in their paper [AAM12] in the local case. In this paper, we elucidate some properties of this construction in the graded case.

Let AA and BB be two graded Artinian Gorenstein (AG) KK-algebras with the same socle degree dd, let TT be an AG KK-algebra of socle degree k<dk<d, and suppose there are surjective maps πA:AT\pi_{A}\colon A\rightarrow T, and πB:BT\pi_{B}\colon B\rightarrow T. From this data, one forms the fiber product algebra A×TBA\times_{T}B as the categorical pullback of πA,πB\pi_{A},\pi_{B}; the connected sum algebra A#TBA\#_{T}B is the quotient of A×TBA\times_{T}B by a certain principal ideal (τA,τB)A×TB\langle(\tau_{A},\tau_{B})\rangle\subset A\times_{T}B. The connected sum is again an AG KK-algebra (see Definition 2.10). As mentioned, this algebraic connected sum operation for local Gorenstein algebras A,BA,B over a local Cohen-Macaulay algebra TT was introduced in [AAM12].

In [Gel22] and [CGS23], the authors determined the minimal free resolution of a two-factor fiber product A×TBA\times_{T}B of local rings. In this paper, we extend their work to the setting of fiber products of graded rings and generalize it to fiber products involving multiple factors. We also consider connected sums with multiple sumands and we answer the following question:

Question 1.1.

Fix A1,,ArA_{1},\ldots,A_{r} graded AG KK-algebras with the same socle degree. What are the graded Betti numbers of their fiber product over KK? What are the graded Betti numbers of their connected sum over KK?

Our first series of main results answers the above question. For specific formulas we refer the reader to Theorem 3.2, Theorem 3.7, Theorem 3.15, Theorem 3.15 and their corollaries.

Celikbas, Laxmi and Weyman solved a particular case of Question 1.1 in [CLW19, Corollary 6.3]. Specifically, they determined a minimal free resolution of the connected sum of KK-algebras Ai:=K[xi]/(xidi)A_{i}:=K[x_{i}]/(x_{i}^{d_{i}}) by using the doubling construction (see section 2.5). A second goal of this paper is to generalize their result and investigate conditions for a connected sum of AG KK-algebras A1,,ArA_{1},\dots,A_{r} with the same socle degree to be a doubling. More precisely we ask:

Question 1.2.

Assume that A1,,ArA_{1},\ldots,A_{r} are graded AG KK-algebras with the same socle degree. Is the connected sum A=A1#K#KArA=A_{1}\#_{K}\cdots\#_{K}A_{r} a doubling? More precisely: if AiA_{i} is a doubling of Ai~\tilde{A_{i}}, is AA a doubling of A1~×K×KAr~\tilde{A_{1}}\times_{K}\cdots\times_{K}\tilde{A_{r}}?

We answer the above question in the affirmative in Theorem 4.3.

Our paper is structured as follows: section 2 introduces the necessary background and develops the basic properties of multi-factor fiber products and connected sums, section 3 computes the graded Betti numbers for multi-factor fiber products and connected sums, and section 4 analyzes connected sums that arise as doublings of certain fiber products.

Acknowledgement. The project got started at the meeting “Workshop on Lefschetz Properties in Algebra, Geometry, Topology and Combinatorics”, held at the Fields Institute in Toronto, Canada, May 15–19, 2023. The authors would like to thank the Fields Institute and the organizers for the invitation and financial support. Additionally, we thank Graham Denham for asking a question which motivated our work, and Mats Boij for useful discussions.

2. background

In this section, we fix some notation and recall some basic facts on Artinian Gorenstein (AG) algebras, fiber products, connected sums of graded Artinian algebras, as well as on Macaulay dual generators needed in the sequel.

2.1. Oriented AG algebras

Throughout this paper, KK is an arbitrary field. Given a graded KK-algebra AA, its homogeneous maximal ideal is mA=i1Aim_{A}=\oplus_{i\geq 1}A_{i}. A KK-algebra AA is called Artinian if it is a finite dimensional vector space over KK. The socle of an Artinian KK-algebra AA is the ideal (0:mA)(0:m_{A}); its socle degree is the largest integer dd such that Ad0A_{d}\neq 0. The socle degree of an Artinian KK-algebra agrees with its Castelnuovo-Mumford regularity, which is denoted by reg(A)\operatorname{reg}(A). The type of AA is the vector space dimension of its socle.

The Hilbert series of a graded KK-algebra AA is the generating function HA(t)=i0dim(Ai)tiH_{A}(t)=\sum_{i\geq 0}\dim(A_{i})t^{i}. The Hilbert function HFAHF_{A} of a KK-algebra AA is the sequence of coefficients of its Hilbert series.

Suppose that AA has a presentation A=R/IA=R/I as a quotient of a graded KK-algebra RR. The graded Betti numbers of AA over RR are the integers βijR(A)=dimKToriR(A,K)j\beta_{ij}^{R}(A)=\dim_{K}\operatorname{Tor}^{R}_{i}(A,K)_{j}. These homological invariants are our main focus. The graded Poincaré series of AA over RR is the generating function PAR(t,s)=i,jβijR(A)tisjP^{R}_{A}(t,s)=\sum_{i,j}\beta_{ij}^{R}(A)t^{i}s^{j}. If RR is regular, then the Poincaré series is in fact a polynomial.

A graded Artinian KK-algebra AA with socle degree dd is said to be Gorenstein if its socle (0:mA)(0:m_{A}) is a one dimensional KK-vector space. For any Artinian Gorenstein graded KK-algebra AA with socle degree dd and for any non-zero morphism of graded vector spaces fA:AK(d)f_{A}:A\to K(-d), known as an orientation of AA, there is a pairing

Ai×AdiK defined by (ai,adi)fA(aiadi)A_{i}\times A_{d-i}\to K\text{ defined by }(a_{i},a_{d-i})\mapsto f_{A}(a_{i}a_{d-i}) (2.1)

which is non-degenerate. We call the pair (A,fA)(A,f_{A}) an oriented AG KK-algebra.

Definition 2.1 ([IMS22, Lemma 2.1]).

Let (A,fA)(A,f_{A}) and (T,fT)(T,f_{T}) be two oriented AG KK-algebras with reg(A)=d\operatorname{reg}(A)=d and reg(T)=k\operatorname{reg}(T)=k, and let π:AT\pi:A\to T be a graded map. There exists a unique homogeneous element τAAdk\tau_{A}\in A_{d-k} such that fA(τa)=fT(π(a))f_{A}(\tau a)=f_{T}(\pi(a)) for all aAa\in A; we call it the Thom class for π:AT\pi:A\to T.

Remark 2.2.

Restating [IMS22, Remark 2.8], we have that τA\tau_{A} is the image of 1T1\in T under the composite map T(k)Extn(T,Q)Extn(A,Q)A(d)T(-k)\cong\operatorname{Ext}^{n}(T,Q)\to\operatorname{Ext}^{n}(A,Q)\cong A(-d), where the middle map is Extn(π,Q)\operatorname{Ext}^{n}(\pi,Q).

Example 2.3.

Let (A,fA)(A,f_{A}) be an oriented AG KK-algebra with socle degree reg(A)=d\operatorname{reg}(A)=d. Consider (K,fK)(K,f_{K}) where fK:KKf_{K}:K\to K is the identity map. Then the Thom class for the canonical projection π:AK\pi:A\to K is the unique element sAds\in A_{d} such that fA(s)=1f_{A}(s)=1.

Note that the Thom class for π:AT\pi:A\to T depends not only on the map π\pi, but also on the orientations chosen for AA and TT.

2.2. Macaulay dual generators

Let Q=K[x1,,xn]Q=K[x_{1},\ldots,x_{n}] be a polynomial ring and let Q=K[X1,,Xn]Q^{\prime}=K[X_{1},\ldots,X_{n}] be a divided power algebra, regarded as a QQ-module with the contraction action

xiXjk={Xjk1δijifk>00otherwisex_{i}\circ X_{j}^{k}=\begin{cases}X_{j}^{k-1}\delta_{ij}&\text{if}\ k>0\\ 0&\text{otherwise}\\ \end{cases}

where δij\delta_{ij} is the Kronecker delta. We regard QQ as a graded KK-algebra with degXi=degxi\deg X_{i}=\deg x_{i}.

For each degree i0i\geq 0, the action of QQ on QQ^{\prime} defines a non-degenerate KK-bilinear pairing

Qi×QiK with (f,F)fF.Q_{i}\times Q^{\prime}_{i}\longrightarrow K\text{ with }(f,F)\longmapsto f\circ F. (2.2)

This implies that for each i0i\geq 0 we have an isomorphism of KK-vector spaces QiHomK(Qi,K)Q^{\prime}_{i}\cong\operatorname{Hom}_{K}(Q_{i},K) given by F{ffF}F\mapsto\left\{f\mapsto f\circ F\right\}.

It is a classical result of Macaulay [Mac94] (cf. [IK99, Lemma 2.14]) that an Artinian KK-algebra A=Q/IA=Q/I is Gorenstein with socle degree dd if and only if I=AnnQ(F)={fQfF=0}I=\operatorname{Ann}_{Q}(F)=\{f\in Q\mid f\circ F=0\} for some homogeneous polynomial FQdF\in Q^{\prime}_{d}. Moreover, this polynomial, termed a Macaulay dual generator for AA, is unique up to a scalar multiple.

A choice of orientation on AA corresponds to a choice of Macaulay dual generator. Every orientation on AA can be written as the function fA:AKf_{A}:A\to K defined by fA(g)(gF)(0)f_{A}(g)\mapsto(g\circ F)(0) for some Macaulay dual generator FF of AA (the notation (gF)(0)(g\circ F)(0) refers to evaluating the element gFg\circ F of QQ^{\prime} at Xi=0X_{i}=0).

2.3. Fiber product

We start by recalling the definition of the fiber product.

Definition 2.4.

Let AA, BB and TT be graded KK-algebras and πA:AT\pi_{A}:A\to T and πB:BT\pi_{B}:B\to T morphisms of graded KK-algebras. We define the fiber product of AA and BB over TT as the graded KK-subalgebra of ABA\oplus B:

A×TB={(a,b)ABπA(a)=πB(b)}.A\times_{T}B=\{(a,b)\in A\oplus B\mid\pi_{A}(a)=\pi_{B}(b)\}.

If πA\pi_{A} and πB\pi_{B} are surjective, then there is a degree-preserving exact sequence

0A×TBABT00\to A\times_{T}B\to A\oplus B\to T\to 0 (2.3)

which allows to compute the Hilbert series of the fiber product as

HFA×TB(t)=HFA(t)+HFB(t)HFT(t).HF_{A\times_{T}B}(t)=HF_{A}(t)+HF_{B}(t)-HF_{T}(t). (2.4)

While presentations of arbitrary fiber products can be unruly, the case T=KT=K is best-behaved.

Lemma 2.5.

Let R=K[x1,,xm]R=K[x_{1},\ldots,x_{m}] and S=K[y1,,yn]S=K[y_{1},\ldots,y_{n}] be polynomial rings over KK with homogeneous maximal ideals 𝐱=(x1,,xm)\mathbf{x}=(x_{1},\ldots,x_{m}) and 𝐲=(y1,,yn)\mathbf{y}=(y_{1},\ldots,y_{n}), respectively. Let Q=RKS=K[x1,,xm,y1,,yn]Q=R\otimes_{K}S=K[x_{1},\ldots,x_{m},y_{1},\ldots,y_{n}]. If A=R/𝔞A=R/\mathfrak{a} and B=S/𝔟B=S/\mathfrak{b} have canonical projections πA:AK\pi_{A}:A\to K and πB:BK\pi_{B}:B\to K, then the fiber product over KK has presentation

A×KB=Q𝐱𝐲+𝔞+𝔟,A\times_{K}B=\frac{Q}{\mathbf{x}\cap\mathbf{y}+\mathfrak{a}\ +\mathfrak{b}}, (2.5)

where in (2.5) 𝔞,𝔟,𝐱,𝐲\mathfrak{a},\mathfrak{b},\mathbf{x},\mathbf{y} denote extensions of the respective ideals to QQ. In particular, if AA and BB are graded, then A×KBA\times_{K}B is a bigraded algebra with

[A×KB](i,j)={K if (i,j)=(0,0),AiBj if (i,j)(0,0).[A\times_{K}B]_{(i,j)}=\begin{cases}K&\text{ if }(i,j)=(0,0),\\ A_{i}\oplus B_{j}&\text{ if }(i,j)\neq(0,0).\end{cases}
Proof.

The presentation of the fiber product is given in [IMS22, Proposition 3.12]. The fact that the fiber product is bigraded follows from noticing that the relations in (2.5) are homogeneous with respect to natural bigrading of QQ. Finally, the formula for the graded components of A×KBA\times_{K}B follows from (2.3), which can be interpreted as an exact sequence of bigraded vector spaces. ∎

Example 2.6.

Consider the standard graded complete intersection algebras

A=K[x,y,z](x3,y4,z4) and B=K[u,v](u5,v5).A=\frac{K[x,y,z]}{(x^{3},y^{4},z^{4})}\ \text{ and }B=\frac{K[u,v]}{(u^{5},v^{5})}.

Their Hilbert functions are given by

HFA\displaystyle HF_{A} =\displaystyle= (1,3,6,9,10,9,6,3,1) and\displaystyle(1,3,6,9,10,9,6,3,1)\ \text{ and }
HFB\displaystyle HF_{B} =\displaystyle= (1,2,3,4,5,4,3,2,1).\displaystyle(1,2,3,4,5,4,3,2,1).

Set R=K[x,y,z]R=K[x,y,z] and S=K[u,v]S=K[u,v]. The minimal free resolutions of AA and BB are the Koszul complexes

0R(11)R(7)2R(8)R(4)2R(3)RA0,0\to R(-11)\to R(-7)^{2}\oplus R(-8)\to R(-4)^{2}\oplus R(-3)\to R\to A\to 0,

and

0S(10)S(5)2SB0.0\to S(-10)\to S(-5)^{2}\to S\to B\to 0.

The fiber product C=A×KBC=A\times_{K}B of AA and BB and its Hilbert function are

C=K[x,y,z,u,v]/(xu,xv,yu,yv,zu,zv,x3,y4,z4,u5,v5),C=K[x,y,z,u,v]/(xu,xv,yu,yv,zu,zv,x^{3},y^{4},z^{4},u^{5},v^{5}),

and

HFC=(1,5,9,13,15,13,9,5,2).HF_{C}=(1,5,9,13,15,13,9,5,2).

The Betti table of CC as a K[x,y,z,u,v]K[x,y,z,u,v]-module is shown in Table 1.

0 1 2 3 4 5
total 1 11 25 24 11 2
0: 1 . . . . .
1: . 6 9 5 1 .
2: . 1 2 1 . .
3: . 2 4 2 . .
4: . 2 6 6 2 .
5: . . 2 4 2 .
6: . . 1 2 1 .
7: . . . . . .
8: . . 1 4 5 2
Table 1. Betti table of CC in Example 2.6

Note that CC is an Artinian level KK-algebra of type 2, i.e., all the elements of its socle have the same degree and the socle has dimension 2.

Recall that an Artinian KK-algebra AA has the strong Lefschetz property (SLP) if there exists a linear form \ell such that the multiplication map ×k:AiAi+k\times\ell^{k}:A_{i}\rightarrow A_{i+k} has maximal rank (i.e, it is injective or surjective) for all ii and kk. It is known that if AA and BB are two AG KK-algebras with the same socle degree, and both have the SLP, then A×KBA\times_{K}B also has the SLP [IMS22, Proposition 5.6].

We also consider multi-factor fiber products, which we now define.

Definition 2.7.

Let A1,,ArA_{1},\ldots,A_{r} and TT be graded KK-algebras and let πi:AiT\pi_{i}:A_{i}\to T morphisms of graded KK-algebras. We define the fiber product of A1,,ArA_{1},\dots,A_{r} over TT as

A1×T×TAr={(a1,,ar)A1Arπi(ai)=πj(aj),1i,jr}.A_{1}\times_{T}\cdots\times_{T}A_{r}=\\ \{(a_{1},\ldots,a_{r})\in A_{1}\oplus\cdots\oplus A_{r}\mid\pi_{i}(a_{i})=\pi_{j}(a_{j}),1\leq i,j\leq r\}. (2.6)
Remark 2.8.

The multi-factor fiber product construction coincides with iteratively applying the two-factor fiber product construction to the list A1,,ArA_{1},\ldots,A_{r} a total of r1r-1 times, that is

A1×T×TAr=((A1×TA2)×T)×TAr.A_{1}\times_{T}\cdots\times_{T}A_{r}=\left((A_{1}\times_{T}A_{2})\times_{T}\cdots\right)\times_{T}A_{r}.

We will need the following generalizations of equations (2.3) and (2.5), which describe a presentation for fiber products with arbitrary many summands over the residue field.

Lemma 2.9.

Let R1,,RrR_{1},\dots,R_{r} be polynomial rings over KK with maximal ideals 𝐱1,,𝐱r\mathbf{x}_{1},\dots,\mathbf{x}_{r}, and let Q=R1KKRrQ=R_{1}\otimes_{K}\dots\otimes_{K}R_{r}. Suppose Ai=Ri/𝔞iA_{i}=R_{i}/\mathfrak{a}_{i}, for some homogeneous ideal 𝔞i\mathfrak{a}_{i} of RiR_{i}. For r2r\geq 2, A1×K×KArQ/JA_{1}\times_{K}\cdots\times_{K}A_{r}\cong Q/J with

J=𝔞1++𝔞r+1ijr(𝐱i𝐱j)J=\mathfrak{a}_{1}+\cdots+\mathfrak{a}_{r}+\sum_{1\leq i\neq j\leq r}(\mathbf{x}_{i}\cap\mathbf{x}_{j})

and there is an exact sequence of graded QQ-modules

0A1×K×KArA1ArKr10.0\to A_{1}\times_{K}\cdots\times_{K}A_{r}\to A_{1}\oplus\cdots\oplus A_{r}\to K^{r-1}\to 0. (2.7)
Proof.

This follows by induction on rr with (2.5) settling the case r=2r=2.

Setting Q=R1KKRr1Q^{\prime}=R_{1}\otimes_{K}\cdots\otimes_{K}R_{r-1}, consider the ideal of QQ^{\prime}

J=𝔞1++𝔞r1+1i<jr1(𝐱i𝐱j).J^{\prime}=\mathfrak{a}_{1}+\cdots+\mathfrak{a}_{r-1}+\sum_{1\leq i<j\leq r-1}(\mathbf{x}_{i}\cap\mathbf{x}_{j}).

Applying (2.5) to A1×K×KAr1Q/JA_{1}\times_{K}\cdots\times_{K}A_{r-1}\cong Q^{\prime}/J^{\prime}, we get

A1×K×KArQ/J×KRi/𝔞rQ/J,A_{1}\times_{K}\cdots\times_{K}A_{r}\cong Q^{\prime}/J^{\prime}\times_{K}R_{i}/\mathfrak{a}_{r}\cong Q/J,

where the ideal JJ is given by

J\displaystyle J =\displaystyle= J+𝔞r+𝐱r(𝐱1++𝐱r1)\displaystyle J^{\prime}+\mathfrak{a}_{r}+\mathbf{x}_{r}\cap(\mathbf{x}_{1}+\cdots+\mathbf{x}_{r-1})
=\displaystyle= 𝔞1++𝔞r1+𝔞r+1i<jr1(𝐱i𝐱j)+1ir1(𝐱i𝐱r)\displaystyle\mathfrak{a}_{1}+\cdots+\mathfrak{a}_{r-1}+\mathfrak{a}_{r}+\sum_{1\leq i<j\leq r-1}(\mathbf{x}_{i}\cap\mathbf{x}_{j})+\sum_{1\leq i\leq r-1}(\mathbf{x}_{i}\cap\mathbf{x}_{r})
=\displaystyle= 𝔞1++𝔞r+1ijr(𝐱i𝐱j).\displaystyle\mathfrak{a}_{1}+\cdots+\mathfrak{a}_{r}+\sum_{1\leq i\neq j\leq r}(\mathbf{x}_{i}\cap\mathbf{x}_{j}).

This yields the claimed presentation.

By Definition 2.7, we have that A1×K×KArA_{1}\times_{K}\cdots\times_{K}A_{r} is a KK-subalgebra of A1ArA_{1}\oplus\cdots\oplus A_{r}. The inclusion A1×K×KArA1ArA_{1}\times_{K}\cdots\times_{K}A_{r}\subseteq A_{1}\oplus\cdots\oplus A_{r} gives the first nonzero map in (2.7), while the second map can be defined by

(a1,,ar)(π2(a2)π1(a1),π3(a3)π1(a1),,πr(ar)π1(a1)),(a_{1},\ldots,a_{r})\mapsto(\pi_{2}(a_{2})-\pi_{1}(a_{1}),\pi_{3}(a_{3})-\pi_{1}(a_{1}),\ldots,\pi_{r}(a_{r})-\pi_{1}(a_{1})),

where the maps πi:AiK\pi_{i}:A_{i}\to K are the canonical projections. The claim regarding exactness of (2.7) follows from Definition 2.7. ∎

2.4. Connected sum

Let (A,fA)(A,f_{A}), (B,fB)(B,f_{B}) and (T,fB)(T,f_{B}) be oriented AG KK-algebras with reg(A)=reg(B)=d\operatorname{reg}(A)=\operatorname{reg}(B)=d and reg(T)=k\operatorname{reg}(T)=k and let πA:AT\pi_{A}:A\to T and πB:BT\pi_{B}:B\to T be surjective graded KK-algebra morphisms with Thom classes τAAdk\tau_{A}\in A_{d-k} and τBBdk\tau_{B}\in B_{d-k}, respectively. We assume that πA(τA)=πB(τB)\pi_{A}(\tau_{A})=\pi_{B}(\tau_{B}), so that (τA,τB)A×TB(\tau_{A},\tau_{B})\in A\times_{T}B.

Definition 2.10.

The connected sum of the oriented AG KK-algebras AA and BB over TT is the quotient ring of the fiber product A×TBA\times_{T}B by the principal ideal generated by the pair of Thom classes (τA,τB)(\tau_{A},\tau_{B}), i.e.

A#TB=(A×TB)/(τA,τB).A\#_{T}B=(A\times_{T}B)/\langle(\tau_{A},\tau_{B})\rangle.

Note that this definition depends on πA\pi_{A}, πB\pi_{B} and the orientations on AA and BB.

By [IMS22, Lemma 3.7] the connected sum is characterized by the following exact sequence of vector spaces:

0T(kd)A×TBA#TB0.0\to T(k-d)\to A\times_{T}B\to A\#_{T}B\to 0. (2.8)

Therefore, the Hilbert series of the connected sum satisfies

HFA#TB(t)=HFA(t)+HFB(t)(1+tdk)HFT(t).HF_{A\#_{T}B}(t)=HF_{A}(t)+HF_{B}(t)-(1+t^{d-k})HF_{T}(t). (2.9)

We recall the following characterization of connected sums.

Theorem 2.11 ([IMS22, Theorem 4.6]).

Let Q=K[x1,,xn]Q=K[x_{1},\ldots,x_{n}] be a polynomial ring, and let Q=K[X1,,Xn]Q^{\prime}=K[X_{1},\ldots,X_{n}] be its dual ring (a divided power algebra). Let F,GQdF,G\in Q^{\prime}_{d} be two linearly independent homogeneous forms of degree dd, and suppose that there exists τQdk\tau\in Q_{d-k} (for some k<dk<d) satisfying

  1. (a)

    τF=τG0\tau\circ F=\tau\circ G\neq 0, and

  2. (b)

    Ann(τF=τG)=Ann(F)+Ann(G)\operatorname{Ann}(\tau\circ F=\tau\circ G)=\operatorname{Ann}(F)+\operatorname{Ann}(G).

Define the oriented AG KK- algebras

A=QAnn(F),B=QAnn(G),T=QAnn(τF=τG),A=\frac{Q}{\operatorname{Ann}(F)},\ B=\frac{Q}{\operatorname{Ann}(G)},\ T=\frac{Q}{\operatorname{Ann}(\tau\circ F=\tau\circ G)},

and let πA:AT\pi_{A}\colon A\rightarrow T and πB:BT\pi_{B}\colon B\rightarrow T be the natural projection maps. Then there are KK-algebra isomorphisms

A×TBQAnn(F)Ann(G),A#TBQAnn(FG).A\times_{T}B\cong\frac{Q}{\operatorname{Ann}(F)\cap\operatorname{Ann}(G)},\ \ A\#_{T}B\cong\frac{Q}{\operatorname{Ann}(F-G)}. (2.10)

Conversely, every connected sum A#TBA\#_{T}B of graded AG KK- algebras with the same socle degree over a graded AG KK-algebra TT arises in this way.

In particular, when T=KT=K the polynomials FF and GG in the above theorem are polynomials expressed in disjoint sets of variables. The connected sum A#KBA\#_{K}B is a graded KK-algebra, but it is not bigraded. Moreover, it is shown in [IMS22, Proposition 5.7] that if AA and BB satisfy the SLP and they have the same socle degree, then A#KBA\#_{K}B also satisfies the SLP.

Example 2.12.

We will now build the connected sum of the standard graded complete intersection KK-algebras A=K[x,y,z]/(x3,y4,z4)A=K[x,y,z]/(x^{3},y^{4},z^{4}) and B=K[u,v]/(u5,v5)B=K[u,v]/(u^{5},v^{5}) described in Example 2.6. The connected sum D=A#KBD=A\#_{K}B is isomorphic to

K[x,y,z,u,v]/(xu,xv,yu,yv,zu,zv,x3,y4,z4,u5,v5,x2y3z3+u4v4).K[x,y,z,u,v]/(xu,xv,yu,yv,zu,zv,x^{3},y^{4},z^{4},u^{5},v^{5},x^{2}y^{3}z^{3}+u^{4}v^{4}).

Its Hilbert function is

HFD=(1,5,9,13,15,13,9,5,1)HF_{D}=(1,5,9,13,15,13,9,5,1)

and its Betti table is given in Table 2.

0 1 2 3 4 5
total 1 12 29 29 12 1
0: 1 . . . . .
1: . 6 9 5 1 .
2: . 1 2 1 . .
3: . 2 4 2 . .
4: . 2 6 6 2 .
5: . . 2 4 2 .
6: . . 1 2 1 .
7: . 1 5 9 6 .
8: . . . . . 1
Table 2. Betti table of DD in Example 2.12

So, DD is an AG KK-algebra with socle degree 8.

An important feature of the connected sum of AG KK-algebras is that it is also an AG KK-algebra with the same socle degree as AA and BB (see [IMS22, Lemma 3.8] or [AAM12, Theorem 1]), in contrast to the fiber product which is an algebra of type two, hence not Gorenstein.

As before, we consider multi-factor connected sums. The multi-factor connected sum construction defined below coincides with iteratively applying the two-factor construction to the list A1,,ArA_{1},\ldots,A_{r} a total of r1r-1 times. In order to define this, we need to define an appropriate orientation and find the Thom class of a connected sum.

Lemma 2.13.

Consider the setup of §2.4 and denote by τA\tau_{A} and τB\tau_{B} the Thom classes of πA\pi_{A} and πB\pi_{B} respectively. Then A#TBA\#_{T}B is an oriented AG KK-algebra with orientation f:A#TBKf:A\#_{T}B\to K defined by f(a,b)=fA(a)fB(b)f(a,b)=f_{A}(a)-f_{B}(b).

Moreover, provided that πA(τA)=0\pi_{A}(\tau_{A})=0, the surjective morphism

π:A#TBT with π(a,b)=πA(a)=πB(b)\pi:A\#_{T}B\to T\text{ with }\pi(a,b)=\pi_{A}(a)=\pi_{B}(b)

has Thom class τ=(τA,0)\tau=(\tau_{A},0).

Proof.

Recall from Theorem 2.11 that if the Macaulay dual generators of AA and BB are FF and GG respectively (chosen to correspond to the given orientations fAf_{A} and fBf_{B}), then the Macaulay dual generator of A#TBA\#_{T}B is FGF-G. This defines an orientation by

g(g(FG))(0)=(gF)(0)(gG)(0).g\mapsto\left(g\circ(F-G)\right)(0)=(g\circ F)(0)-(g\circ G)(0).

If g=(a,b)A#TBg=(a,b)\in A\#_{T}B, then (gF)(0)=fA(a)(g\circ F)(0)=f_{A}(a) and (gG)(0)=fB(b)(g\circ G)(0)=f_{B}(b).

To establish the claim regarding the Thom class we verify that

f(τg)=fT(π(g)) for all gA#TB.f(\tau g)=f_{T}(\pi(g))\text{ for all }g\in A\#_{T}B.

With g=(a,b)g=(a,b), we have τg=(τA,0)(a,b)=(τAa,0)\tau g=(\tau_{A},0)(a,b)=(\tau_{A}a,0) since

f(τg)=fA(τAa)=fA(πA(a))=f(π(g)).f(\tau g)=f_{A}(\tau_{A}a)=f_{A}(\pi_{A}(a))=f(\pi(g)).

We establish the convention that every connected sum in this paper will be oriented according to the orientation ff in Lemma 2.13.

Definition 2.14.

Let A1,,ArA_{1},\ldots,A_{r} and TT be graded AG KK-algebras with socle degrees reg(Ai)=d\operatorname{reg}(A_{i})=d and reg(T)=k\operatorname{reg}(T)=k and let πi:AiT\pi_{i}:A_{i}\to T be morphisms of graded KK-algebras with Thom classes τ1,,τr\tau_{1},\ldots,\tau_{r}, respectively such that πi(τi)=0\pi_{i}(\tau_{i})=0. We define the multi-factor connected sum A1#T#TArA_{1}\#_{T}\cdots\#_{T}A_{r} by

A1#T#TAr=A1×T×TAr/(τ1,0,,0,τi,0,,0)2ir.A_{1}\#_{T}\cdots\#_{T}A_{r}=\\ A_{1}\times_{T}\cdots\times_{T}A_{r}/\langle(\tau_{1},0,\ldots,0,\tau_{i},0,\ldots,0)\mid 2\leq i\leq r\rangle. (2.11)
Remark 2.15.

Note that the above definition coincides with iterating the two-factor connected sum construction. Indeed, repeatedly applying Lemma 2.13 shows that the Thom class of the iterated connected sum

((A1#TA2)#TA3)#TAi1((A_{1}\#_{T}A_{2})\#_{T}A_{3})\cdots\#_{T}A_{i-1}

is (τ1,0,,0)(\tau_{1},0,\ldots,0). Thus, to obtain

(((A1#TA2)#TA3)#TAi1)#TAi,(((A_{1}\#_{T}A_{2})\#_{T}A_{3})\cdots\#_{T}A_{i-1})\#_{T}A_{i},

one goes modulo (τ1,0,,0,τi)\langle(\tau_{1},0,\ldots,0,\tau_{i})\rangle.

Lemma 2.16.

In the setup of Definition 2.14, there is a short exact sequence

0T(dk)r1𝛼A1×T×TArA1#T#TAr0,0\to T(d-k)^{r-1}\xrightarrow{\alpha}A_{1}\times_{T}\cdots\times_{T}A_{r}\to A_{1}\#_{T}\cdots\#_{T}A_{r}\to 0, (2.12)

where the map α\alpha sends the generator of the ii-th summand TT to the image of (τ1,0,,0,τi+1,0,,0)(\tau_{1},0,\ldots,0,\tau_{i+1},0,\ldots,0) in A1×T×TArA_{1}\times_{T}\cdots\times_{T}A_{r}.

Proof.

It is clear from Definition 2.14 that the connected sum is the cokernel of α\alpha. It remains to justify that this map is injective. This follows from [IMS22, Lemma 2.6], where it is shown that the Gysin map ι:T(dk)Ai\iota:T(d-k)\to A_{i} that satisfies ι(1)=τi\iota(1)=\tau_{i} is injective. ∎

2.5. Doubling

Let us start by recalling the doubling construction and some basic facts needed later on.

Definition 2.17.

Set R=K[x1,,xn]R=K[x_{1},\ldots,x_{n}]. The canonical module of a graded RR-module MM is ωM=ExtRndimM(M,R)\omega_{M}=\operatorname{Ext}^{n-\dim M}_{R}(M,R).

For example, one has ωKK\omega_{K}\cong K.

Definition 2.18.

[KKR+21, Section 2.5] Let JRJ\subset R be a homogeneous ideal of codimension cc, such that R/JR/J is Cohen-Macaulay and ωR/J\omega_{R/J} is its canonical module. Furthermore, assume that R/JR/J satisfies the condition G0G_{0} (i.e., it is Gorenstein at all minimal primes). Let II be an ideal of codimension c+1c+1. II is called a doubling of JJ via ψ\psi if there exists a short exact sequence of R/JR/J modules

0ωR/J(d)ψR/JR/I0.0\rightarrow\omega_{R/J}(-d)\stackrel{{\scriptstyle\psi}}{{\rightarrow}}R/J\rightarrow R/I\rightarrow 0. (2.13)

By [BH93, Proposition 3.3.18], if II is a doubling, then R/IR/I is a Gorenstein ring.

Doubling plays an important role in the theory of Gorenstein liaison. Indeed, in [KMMR+01], doubling is used to produce suitable Gorenstein divisors on arithmetically Cohen-Macaulay subschemes in several foundational constructions. It is not true that every Artinian Gorenstein ideal of codimension c+1c+1 is a doubling of some codimension cc ideal (see, for instance, [KKR+21, Example 2.19]).

Moreover, the mapping cone of ψ\psi in (2.13) gives a resolution of R/IR/I. If it is minimal, then one can read off the Betti table of R/IR/I from the Betti table of R/JR/J. This mapping cone is the direct sum of the minimal free resolution FF_{\bullet} of R/JR/J with its dual (reversed) complex Hom(F,R)\operatorname{Hom}(F_{\bullet},R) which justifies the terminology of "doubling".

Lemma 2.19.

Let C=R/JC=R/J be a Cohen-Macaulay KK-algebra. Then:

  • (a)

    regωC=dimC\operatorname{reg}\omega_{C}=\dim C; and

  • (b)

    if CC is a doubling of a Cohen-Macaulay KK-algebra C~\tilde{C}, i.e., dimC~=dimC+1\dim\tilde{C}=\dim C+1 and there is an exact sequence of graded RR-modules

    0ωC~(t)C~C0,0\to\omega_{\tilde{C}}(-t)\to\tilde{C}\to C\to 0,

    then t=regCdimCt=\operatorname{reg}C-\dim C.

Proof.

(a) Consider a graded minimal free resolution of CC as an RR-module

0FcF1RC0,0\to F_{c}\to\cdots\to F_{1}\to R\to C\to 0,

where c=dimRdimCc=\dim R-\dim C denotes the codimension. Dualizing and then shifting it, we obtain a graded minimal free resolution of ωC\omega_{C} of the form

0R(dimR)F1(dimR)Fc(dimR)ωC0.0\to R(-\dim R)\to F_{1}^{*}(-\dim R)\to\cdots\to F_{c}^{*}(-\dim R)\to\omega_{C}\to 0.

Claim (a) follows using the characterization of regularity by graded Betti numbers.

(b) The Tor\operatorname{Tor} sequence of the given short exact sequence begins

0TorcR(C,K)Torc1R(ωC~,K)(t).0\to\operatorname{Tor}_{c}^{R}(C,K)\to\operatorname{Tor}_{c-1}^{R}(\omega_{\tilde{C}},K)(-t)\to\cdots.

Since CC is a Gorenstein algebra, TorcR(C,K)\operatorname{Tor}_{c}^{R}(C,K) is concentrated in degree c+regCc+\operatorname{reg}C. The above resolution of ωC\omega_{C} shows that Torc1R(ωC~,K)(t)\operatorname{Tor}_{c-1}^{R}(\omega_{\tilde{C}},K)(-t) is concentrated in degree t+dimRt+\dim R. It follows that c+regC=t+dimRc+\operatorname{reg}C=t+\dim R, which proves Claim (b). ∎

3. Graded Betti numbers of the Connected Sum

3.1. Two Summands

Let R=K[x1,,xm]R=K[x_{1},\ldots,x_{m}] and S=K[y1,,yn]S=K[y_{1},\ldots,y_{n}] be polynomial rings over KK with the standard grading, and let 𝐱=(x1,,xm)\mathbf{x}=(x_{1},\ldots,x_{m}) and 𝐲=(y1,,yn)\mathbf{y}=(y_{1},\ldots,y_{n}) denote the homogeneous maximal ideals of RR and SS, respectively. Set

Q=RKSK[x1,,xm,y1,,yn].Q=R\otimes_{K}S\cong K[x_{1},\ldots,x_{m},y_{1},\ldots,y_{n}].

Let A=R/𝔞A=R/\mathfrak{a} and B=S/𝔟B=S/\mathfrak{b} be standard graded KK-algebras. We will assume 𝔞1=𝔟1=0\mathfrak{a}_{1}=\mathfrak{b}_{1}=0, that is, the ideals 𝔞\mathfrak{a} and 𝔟\mathfrak{b} do not contain any non-zero linear forms. Note that R×KS=Q/(𝐱𝐲)R\times_{K}S=Q/(\mathbf{x}\cap\mathbf{y}) by Lemma 2.5. We start by determining the Betti numbers of this ring. Note that the ideal 𝐱𝐲\mathbf{x}\cap\mathbf{y} of QQ is a so-called Ferrers ideal and admits a minimal graded free cellular resolution that is supported on the join of two simplices (see [CN09]). We show other useful fact about this resolution below.

Henceforth, we set (ab)=0\binom{a}{b}=0 if b>ab>a.

Lemma 3.1.

The ideal 𝐱𝐲\mathbf{x}\cap\mathbf{y} of QQ has a 2-linear minimal free graded resolution over QQ and, for i1i\geq 1, we have

[ToriQ(Q/𝐱𝐲,K)]i+1coker([Tori+1Q(Q/𝐱,K)Tori+1Q(Q/𝐲,K)]i+1[Tori+1Q(K,K)]i+1).[\operatorname{Tor}_{i}^{Q}(Q/\mathbf{x}\cap\mathbf{y},K)]_{i+1}\cong\\ \operatorname{coker}\big{(}[\operatorname{Tor}^{Q}_{i+1}(Q/\mathbf{x},K)\oplus\operatorname{Tor}^{Q}_{i+1}(Q/\mathbf{y},K)]_{i+1}\to[\operatorname{Tor}^{Q}_{i+1}(K,K)]_{i+1}\big{)}.

In particular, one has

dimK[ToriQ(Q/𝐱𝐲,K)]i+1=(m+ni+1)(mi+1)(ni+1)\dim_{K}[\operatorname{Tor}_{i}^{Q}(Q/\mathbf{x}\cap\mathbf{y},K)]_{i+1}=\binom{m+n}{i+1}-\binom{m}{i+1}-\binom{n}{i+1}

and PQ/𝐱𝐲Q(t,s)=t1[(1+st)m1][(1+st)n1]+1P^{Q}_{Q/\mathbf{x}\cap\mathbf{y}}(t,s)=t^{-1}[(1+st)^{m}-1][(1+st)^{n}-1]+1.

Proof.

For brevity, let us write ToriQ(M)\operatorname{Tor}_{i}^{Q}(M) instead of ToriQ(M,K)\operatorname{Tor}_{i}^{Q}(M,K) for a QQ-module MM. The first part is well-known (see, e.g., [CN09, Theorem 2.1]). It also follows from the Mayer-Vietoris sequence

0Q/𝐱𝐲Q/𝐱Q/𝐲Q/𝐱+𝐲0.0\to Q/\mathbf{x}\cap\mathbf{y}\to Q/\mathbf{x}\oplus Q/\mathbf{y}\to Q/\mathbf{x}+\mathbf{y}\to 0.

Observe that Q/𝐱+𝐲KQ/\mathbf{x}+\mathbf{y}\cong K as a QQ-module. Thus, for any integer i0i\geq 0, the induced long exact sequence in Tor\operatorname{Tor} gives

[Tori+1Q(Q/𝐱𝐲)]i+1[Tori+1Q(Q/𝐱)Tori+1Q(Q/𝐲)]i+1\displaystyle[\operatorname{Tor}_{i+1}^{Q}(Q/\mathbf{x}\cap\mathbf{y})]_{i+1}\to[\operatorname{Tor}_{i+1}^{Q}(Q/\mathbf{x})\oplus\operatorname{Tor}_{i+1}^{Q}(Q/\mathbf{y})]_{i+1}\to
[Tori+1Q(K)]i+1[ToriQ(Q/𝐱𝐲)]i+1[ToriQ(Q/𝐱)ToriQ(Q/𝐲)]i+1.\displaystyle[\operatorname{Tor}_{i+1}^{Q}(K)]_{i+1}\to[\operatorname{Tor}_{i}^{Q}(Q/\mathbf{x}\cap\mathbf{y})]_{i+1}\to[\operatorname{Tor}_{i}^{Q}(Q/\mathbf{x})\oplus\operatorname{Tor}_{i}^{Q}(Q/\mathbf{y})]_{i+1}.

The left-most module in this sequence is zero because 𝐱𝐲\mathbf{x}\cap\mathbf{y} has a 2-linear resolution. Similarly, the right-most module is zero because 𝐱\mathbf{x} and 𝐲\mathbf{y} have linear resolutions. Thus, the second claim follows. The last claim can be verified directly based on the second. ∎

The following result gives a formula to compute the graded Betti numbers of the fiber product A×KBA\times_{K}B.

Notation.

Given a power series P(t,s)=cijsjti[s][[t]]P(t,s)=\sum c_{ij}s^{j}t^{i}\in\mathbb{Z}[s][[t]], we set P~(t,s)=j>icijsjti\widetilde{P}(t,s)=\sum\limits_{j>i}c_{ij}s^{j}t^{i} to be the sum of the terms of P(t,s)P(t,s) having j>ij>i.

Theorem 3.2.

For any integer i1i\geq 1, there is an isomorphism of graded KK-vector spaces

[ToriQ(A×KB,K)]j{0 if ji,[ToriQ(A,K)]i+1[ToriQ(B,K)]i+1[ToriQ(Q/𝐱𝐲,K)]i+1 if j=i+1,[ToriQ(A,K)]j[ToriQ(B,K)]j if ji+2.[\operatorname{Tor}_{i}^{Q}(A\times_{K}B,K)]_{j}\\ \cong\begin{cases}0&\text{ if $j\leq i$},\\[2.0pt] [\operatorname{Tor}_{i}^{Q}(A,K)]_{i+1}\oplus[\operatorname{Tor}_{i}^{Q}(B,K)]_{i+1}&\\ \oplus[\operatorname{Tor}_{i}^{Q}(Q/\mathbf{x}\cap\mathbf{y},K)]_{i+1}&\text{ if $j=i+1$},\\[2.0pt] [\operatorname{Tor}_{i}^{Q}(A,K)]_{j}\oplus[\operatorname{Tor}_{i}^{Q}(B,K)]_{j}&\text{ if $j\geq i+2$}.\end{cases}

In particular, one has

PA×KBQ(t,s)=P~AQ(t,s)+P~BQ(t,s)+PQ/𝐱𝐲Q(t,s)P^{Q}_{A\times_{K}B}(t,s)=\widetilde{P}^{Q}_{A}(t,s)+\widetilde{P}^{Q}_{B}(t,s)+P^{Q}_{Q/\mathbf{x}\cap\mathbf{y}}(t,s)

and

reg(A×KB)=max{regA,regB}.\operatorname{reg}(A\times_{K}B)=\max\{\operatorname{reg}A,\operatorname{reg}B\}.
Proof.

Again, we write ToriQ(M)\operatorname{Tor}_{i}^{Q}(M) instead of ToriQ(M,K)\operatorname{Tor}_{i}^{Q}(M,K) for any QQ-module MM. Consider the exact sequence (2.3) of graded QQ-modules

0A×KBABK0.0\to A\times_{K}B\to A\oplus B\to K\to 0.

Its long exact sequence in Tor\operatorname{Tor} gives exact sequences

Tori+1Q(K)ToriQ(A×KB)ToriQ(A)ToriQ(B)ToriQ(K).\operatorname{Tor}_{i+1}^{Q}(K)\to\operatorname{Tor}_{i}^{Q}(A\times_{K}B)\to\operatorname{Tor}_{i}^{Q}(A)\oplus\operatorname{Tor}_{i}^{Q}(B)\to\operatorname{Tor}_{i}^{Q}(K). (3.1)

As a QQ-module, KK is resolved by a Koszul complex, which shows in particular [ToriQ(K)]j0[\operatorname{Tor}_{i}^{Q}(K)]_{j}\neq 0 if and only if 0i=jm+n0\leq i=j\leq m+n. Considering the sequence (3.1) in degree jj, we conclude that

[ToriQ(A×KB,K)]j[ToriQ(A,K)]j[ToriQ(B,K)]j[\operatorname{Tor}_{i}^{Q}(A\times_{K}B,K)]_{j}\cong[\operatorname{Tor}_{i}^{Q}(A,K)]_{j}\oplus[\operatorname{Tor}_{i}^{Q}(B,K)]_{j}

if ji+2j\geq i+2.

Using the fact that A×KBQ/(𝐱𝐲,𝔞,𝔟)A\times_{K}B\cong Q/(\mathbf{x}\cap\mathbf{y},\mathfrak{a},\mathfrak{b})–as in (2.5)–and that the initial degree of (𝐱𝐲,𝔞,𝔟)(\mathbf{x}\cap\mathbf{y},\mathfrak{a},\mathfrak{b}) is two since 𝔞\mathfrak{a} and 𝔟\mathfrak{b} do not contain any linear forms by assumption, we see that for jij\leq i,

[ToriQ(A×KB,K)]j=0.[\operatorname{Tor}_{i}^{Q}(A\times_{K}B,K)]_{j}=0.

It remains to consider [ToriQ(A×KB,K)]i+1[\operatorname{Tor}_{i}^{Q}(A\times_{K}B,K)]_{i+1}. To this end, we use a longer part of the exact Tor\operatorname{Tor} sequence and consider it in degree i+1i+1:

[Tori+1Q(A)Tori+1Q(B)]i+1γ[Tori+1Q(K)]i+1\displaystyle[\operatorname{Tor}_{i+1}^{Q}(A)\oplus\operatorname{Tor}_{i+1}^{Q}(B)]_{i+1}\stackrel{{\scriptstyle\gamma}}{{\longrightarrow}}[\operatorname{Tor}_{i+1}^{Q}(K)]_{i+1}\to
[ToriQ(A×KB)]i+1[ToriQ(A)ToriQ(B)]i+1[ToriQ(K)]i+1.\displaystyle\hskip 28.45274pt[\operatorname{Tor}_{i}^{Q}(A\times_{K}B)]_{i+1}\to[\operatorname{Tor}_{i}^{Q}(A)\oplus\operatorname{Tor}_{i}^{Q}(B)]_{i+1}\to[\operatorname{Tor}_{i}^{Q}(K)]_{i+1}.

The right-most module in this sequence is zero because KK has a linear resolution as a QQ-module. Note that there are isomorphisms of graded QQ-modules AQ/(𝔞,𝐲)A\cong Q/(\mathfrak{a},\mathbf{y}) and B=Q/(𝐱,𝔟)B=Q/(\mathbf{x},\mathfrak{b}). Since 𝔞1=𝔟1=0\mathfrak{a}_{1}=\mathfrak{b}_{1}=0, it follows that

[Tori+1Q(A)Tori+1Q(B)]i+1[Tori+1Q(Q/𝐲)Tori+1Q(Q/𝐱)]i+1.[\operatorname{Tor}_{i+1}^{Q}(A)\oplus\operatorname{Tor}_{i+1}^{Q}(B)]_{i+1}\cong[\operatorname{Tor}_{i+1}^{Q}(Q/\mathbf{y})\oplus\operatorname{Tor}_{i+1}^{Q}(Q/\mathbf{x})]_{i+1}.

Therefore, the cokernel of the map γ\gamma is equal to the cokernel of the map [Tori+1Q(Q/𝐱)Tori+1Q(Q/𝐲)]i+1[Tori+1Q(K)]i+1[\operatorname{Tor}_{i+1}^{Q}(Q/\mathbf{x})\oplus\operatorname{Tor}_{i+1}^{Q}(Q/\mathbf{y})]_{i+1}\to[\operatorname{Tor}_{i+1}^{Q}(K)]_{i+1}. By Lemma 3.1, the latter is isomorphic to [ToriQ(Q/𝐱𝐲,K)]i+1[\operatorname{Tor}_{i}^{Q}(Q/\mathbf{x}\cap\mathbf{y},K)]_{i+1}. Hence, the above sequence proves the claim for [ToriQ(A×KB,K)]i+1[\operatorname{Tor}_{i}^{Q}(A\times_{K}B,K)]_{i+1}.

The claim regarding the Poincaré series follows by taking the dimensions of the Tor\operatorname{Tor} modules and using the isomorphism in the first part of the theorem. ∎

Remark 3.3.

The previous result allows us to write the Betti numbers of the fiber product in terms of those of the summands:

βi,jQ(A×KB)={0 if jiβi,i+1Q(A)+βi,i+1Q(B)+βi,i+1Q(Q/𝐱𝐲) if j=i+1βi,jQ(A)+βi,jQ(B) if ji+2.\beta^{Q}_{i,j}(A\times_{K}B)=\begin{cases}0&\text{ if }j\leq i\\[2.0pt] \beta^{Q}_{i,i+1}(A)+\beta^{Q}_{i,i+1}(B)+\beta^{Q}_{i,i+1}(Q/\mathbf{x}\cap\mathbf{y})&\text{ if }j=i+1\\[2.0pt] \beta^{Q}_{i,j}(A)+\beta^{Q}_{i,j}(B)&\text{ if }j\geq i+2.\end{cases}

In Theorem 3.2, we have computed the graded Betti numbers of the fiber product A×KBA\times_{K}B in terms of the graded Betti numbers of AA and BB as QQ-modules and we will now convert these formulas into formulas depending only on [ToriR(A,K)]j[\operatorname{Tor}_{i}^{R}(A,K)]_{j} and [ToriS(B,K)]j[\operatorname{Tor}_{i}^{S}(B,K)]_{j}.

Notation.

We set [x]+=max{0,x}[x]_{+}=\max\{0,x\}.

Proposition 3.4.

The identities

PAQ(t,s)\displaystyle P^{Q}_{A}(t,s) =\displaystyle= PAR(t,s)(1+st)dimS, and\displaystyle P^{R}_{A}(t,s)\cdot(1+st)^{\dim S},\text{ and}
PBQ(t,s)\displaystyle P^{Q}_{B}(t,s) =\displaystyle= PBS(t,s)(1+st)dimR\displaystyle P^{S}_{B}(t,s)\cdot(1+st)^{\dim R}

yield

P~AQ(t,s)\displaystyle\widetilde{P}^{Q}_{A}(t,s) =\displaystyle= (PAR(t,s)1)(1+st)dimS, and\displaystyle(P^{R}_{A}(t,s)-1)\cdot(1+st)^{\dim S},\text{ and}
P~BQ(t,s)\displaystyle\widetilde{P}^{Q}_{B}(t,s) =\displaystyle= (PBS(t,s)1)(1+st)dimR.\displaystyle(P^{S}_{B}(t,s)-1)\cdot(1+st)^{\dim R}.

Thus, with m=dimRm=\dim R and n=dimSn=\dim S, we have

βi,jQ(A)\displaystyle\beta_{i,j}^{Q}(A) =\displaystyle= =0min(i,m)1(n[im]++)βmin(i,m),j[im]+R(A), and\displaystyle\sum_{\ell=0}^{\min(i,m)-1}\binom{n}{[i-m]_{+}+\ell}\beta_{\min(i,m)-\ell,j-\ell-[i-m]_{+}}^{R}(A),\text{ and}
βi,jQ(B)\displaystyle\beta_{i,j}^{Q}(B) =\displaystyle= =0min(i,n)1(m[in]++)βmin(i,n),j[in]+S(B).\displaystyle\sum_{\ell=0}^{\min(i,n)-1}\binom{m}{[i-n]_{+}+\ell}\beta_{\min(i,n)-\ell,j-\ell-[i-n]_{+}}^{S}(B).
Proof.

Since QQ is a free RR-module, we have PQ/𝔞Q(t,s)=PR/𝔞R(t,s)=PAR(t,s)P^{Q}_{Q/\mathfrak{a}}(t,s)=P^{R}_{R/\mathfrak{a}}(t,s)=P^{R}_{A}(t,s). We know A=Q/(𝔞+𝐲)A=Q/(\mathfrak{a}+\mathbf{y}), and ToriK(Q/𝔞,Q/𝐲)=0\operatorname{Tor}_{i}^{K}(Q/\mathfrak{a},Q/\mathbf{y})=0 for i1i\geq 1, so the minimal free resolution of AA as a QQ-module is obtained by tensoring the minimal free resolutions of Q/𝔞Q/\mathfrak{a} and Q/𝐲Q/\mathbf{y}. This justifies the identity

PAQ(t,s)=PQ/𝔞Q(t,s)PQ/𝐲Q(t,s)=PAR(t,s)(1+st)dimS.P^{Q}_{A}(t,s)=P^{Q}_{Q/\mathfrak{a}}(t,s)\cdot P^{Q}_{Q/\mathbf{y}}(t,s)=P^{R}_{A}(t,s)\cdot(1+st)^{\dim S}.

Since the terms with equal exponents for ss and tt in PAQ(t,s)P^{Q}_{A}(t,s) arise from the constant term of the first factor multiplied with the second factor, removing this term yields P~AQ(t,s)\widetilde{P}^{Q}_{A}(t,s). Therefore, using the convention (na)=0\binom{n}{a}=0 if a>na>n, we conclude that

βi,jQ(A)={=0i1(n)βi,jR(A) if 1im, and =0m1(nim+)βm,ji+mR(A) if m<im+n.\beta_{i,j}^{Q}(A)=\begin{cases}\sum_{\ell=0}^{i-1}\binom{n}{\ell}\beta_{i-\ell,j-\ell}^{R}(A)\text{ if }1\leq i\leq m,\text{ and }\\ \\ \sum_{\ell=0}^{m-1}\binom{n}{i-m+\ell}\beta_{m-\ell,j-i+m-\ell}^{R}(A)\text{ if }m<i\leq m+n.\end{cases}

The claims regarding BB are justified similarly. ∎

Combining Lemma 3.1, Theorem 3.2 and Proposition 3.4, we get:

Corollary 3.5.

With the above notation, we have

PA×KBQ(t,s)=(PAR(t,s)1)(1+st)n+(PBS(t,s)1)(1+st)m++t1[(1+st)m1][(1+st)n1]+1,\begin{split}P^{Q}_{A\times_{K}B}(t,s)={}&(P^{R}_{A}(t,s)-1)\cdot(1+st)^{n}+(P^{S}_{B}(t,s)-1)\cdot(1+st)^{m}+\\ &+t^{-1}[(1+st)^{m}-1][(1+st)^{n}-1]+1,\end{split} (3.2)

that is,

βi,jQ(A×KB)={0 if ji,(i,j)(0,0),1 if i=j=0,=0min(i,m)1(n[im]++)βmin(i,m),i+1[im]+R(A)+=0min(i,n)1(m[in]++)βmin(i,n),i+1[in]+S(B)+(m+ni+1)(mi+1)(ni+1) if j=i+1,=0min(i,m)1(n[im]++)βmin(i,m),j[im]+R(A)+=0min(i,n)1(m[in]++)βmin(i,n),j[in]+S(B) if ji+2.\beta^{Q}_{i,j}(A\times_{K}B)=\\ \begin{cases}0&\text{ if }j\leq i,(i,j)\neq(0,0),\\ 1&\text{ if }i=j=0,\\[7.11317pt] \begin{array}[]{l}\sum_{\ell=0}^{\min(i,m)-1}\binom{n}{[i-m]_{+}+\ell}\beta_{\min(i,m)-\ell,i+1-\ell-[i-m]_{+}}^{R}(A)+\\ \sum_{\ell=0}^{\min(i,n)-1}\binom{m}{[i-n]_{+}+\ell}\beta_{\min(i,n)-\ell,i+1-\ell-[i-n]_{+}}^{S}(B)\\ +\binom{m+n}{i+1}-\binom{m}{i+1}-\binom{n}{i+1}\end{array}&\text{ if }j=i+1,\\[14.22636pt] \begin{array}[]{l}\sum_{\ell=0}^{\min(i,m)-1}\binom{n}{[i-m]_{+}+\ell}\beta_{\min(i,m)-\ell,j-\ell-[i-m]_{+}}^{R}(A)\\ +\sum_{\ell=0}^{\min(i,n)-1}\binom{m}{[i-n]_{+}+\ell}\beta_{\min(i,n)-\ell,j-\ell-[i-n]_{+}}^{S}(B)\end{array}&\text{ if }j\geq i+2.\end{cases}
Remark 3.6.

The identity (3.2) extends [Gel22, Theorem 1.1] to the graded case. It can be checked that the two results agree upon substituting s=1s=1 in (3.2).

We now turn to the graded Betti numbers of the connected sum. AG KK-algebras with socle degree two have been classified by [Sal79]. Explicitly, if A=R/𝔞A=R/\mathfrak{a} has hh-vector (1,n,1)(1,n,1) then, up to isomorphism, one has

𝔞=(xixj 1ijn)+(x12x22,,x12xn2).\mathfrak{a}=(x_{i}x_{j}\;\mid\;1\leq i\neq j\leq n)+(x_{1}^{2}-x_{2}^{2},\ldots,x_{1}^{2}-x_{n}^{2}).

The graded minimal free resolution of AA as RR-module has the form

0R(n2)R(n)βn1R(2)β1RA0,0\to R(-n-2)\to R(-n)^{\beta_{n-1}}\to\cdots\to R(-2)^{\beta_{1}}\to R\to A\to 0,

and a straightforward computation gives us

βi=βn1i=i(ni+1)+(ni)(nni+1)\beta_{i}=\beta_{n-1-i}=i\binom{n}{i+1}+(n-i)\binom{n}{n-i+1}

for 1in11\leq i\leq n-1. Thus, it is harmless to consider AG KK-algebras whose socle degree is at least three.

Theorem 3.7.

Assume that AA and BB are AG KK-algebras such that reg(A)=reg(B)=e3\operatorname{reg}(A)=\operatorname{reg}(B)=e\geq 3. For any integer i1i\geq 1, there is an isomorphism of graded KK-vector spaces

[ToriQ(A#KB,K)]j{0 if ji and (i,j)(0,0),K if (i,j)=(0,0),[ToriQ(A,K)]i+1[ToriQ(B,K)]i+1[ToriQ(Q/𝐱𝐲,K)]i+1 if j=i+1,[ToriQ(A,K)]j[ToriQ(B,K)]j if i+2ji+e2,[ToriQ(A,K)]j[ToriQ(B,K)]j[Torm+niQ(Q/𝐱𝐲,K)]m+ni+1 if j=i+e1,K if (i,j)=(m+n,e+m+n),0 if je+i and(i,j)(m+n,e+m+n);[\operatorname{Tor}_{i}^{Q}(A\#_{K}B,K)]_{j}\cong\\ \begin{cases}0&\text{ if }j\leq i\text{ and }(i,j)\neq(0,0),\\[2.0pt] K&\text{ if }(i,j)=(0,0),\\[2.0pt] [\operatorname{Tor}_{i}^{Q}(A,K)]_{i+1}\oplus[\operatorname{Tor}_{i}^{Q}(B,K)]_{i+1}&\\ \oplus[\operatorname{Tor}_{i}^{Q}(Q/\mathbf{x}\cap\mathbf{y},K)]_{i+1}&\text{ if }j=i+1,\\[2.0pt] [\operatorname{Tor}_{i}^{Q}(A,K)]_{j}\oplus[\operatorname{Tor}_{i}^{Q}(B,K)]_{j}&\text{ if }i+2\leq j\leq i+e-2,\\[2.0pt] [\operatorname{Tor}_{i}^{Q}(A,K)]_{j}\oplus[\operatorname{Tor}_{i}^{Q}(B,K)]_{j}&\\ \oplus[\operatorname{Tor}_{m+n-i}^{Q}(Q/\mathbf{x}\cap\mathbf{y},K)]_{m+n-i+1}&\text{ if }j=i+e-1,\\[2.0pt] K&\text{ if }(i,j)=(m+n,e+m+n),\\[2.0pt] 0&\text{ if }j\geq e+i\text{ and}\\ &(i,j)\neq(m+n,e+m+n);\end{cases}

equivalently,

PA#KBQ(t,s)=P~AQ(t,s)+P~BQ(t,s)+PQ/𝐱𝐲Q(t,s)+sm+n+etm+nPQ/𝐱𝐲Q(t1,s1).P^{Q}_{A\#_{K}B}(t,s)=\widetilde{P}_{A}^{Q}(t,s)+\widetilde{P}_{B}^{Q}(t,s)+P^{Q}_{Q/\mathbf{x}\cap\mathbf{y}}(t,s)\\ +s^{m+n+e}t^{m+n}P^{Q}_{Q/\mathbf{x}\cap\mathbf{y}}(t^{-1},s^{-1}).
Proof.

For a QQ-module MM, we write ToriQ(M)\operatorname{Tor}_{i}^{Q}(M) instead of ToriQ(M,K)\operatorname{Tor}_{i}^{Q}(M,K). Consider the exact sequence (2.8) of graded QQ-modules

0K(e)A×KBA#KB0.0\to K(-e)\to A\times_{K}B\to A\#_{K}B\to 0.

Its long exact Tor\operatorname{Tor} sequence gives exact sequences

[ToriQ(K)]je[ToriQ(A×KB)]j[ToriQ(A#KB)]j[Tori1Q(K)]je.[\operatorname{Tor}_{i}^{Q}(K)]_{j-e}\to[\operatorname{Tor}_{i}^{Q}(A\times_{K}B)]_{j}\to\\ [\operatorname{Tor}_{i}^{Q}(A\#_{K}B)]_{j}\to[\operatorname{Tor}_{i-1}^{Q}(K)]_{j-e}.

Since ToriQ(K)\operatorname{Tor}_{i}^{Q}(K) is concentrated in degree ii we conclude that

[ToriQ(A#KB)]j[ToriQ(A×KB)]j[\operatorname{Tor}_{i}^{Q}(A\#_{K}B)]_{j}\cong[\operatorname{Tor}_{i}^{Q}(A\times_{K}B)]_{j}

if j{e+i1,e+i}j\notin\{e+i-1,e+i\}. Combined with Theorem 3.2, this determines [ToriQ(A#KB)]j[\operatorname{Tor}_{i}^{Q}(A\#_{K}B)]_{j} if je+i2j\leq e+i-2.

Using the fact that reg(A#KB)=regA=regB=e\operatorname{reg}(A\#_{K}B)=\operatorname{reg}A=\operatorname{reg}B=e which can be deduced from (2.9), we know that [ToriQ(A#KB)]j=0[\operatorname{Tor}_{i}^{Q}(A\#_{K}B)]_{j}=0 if je+i+1j\geq e+i+1. It remains to determine [ToriQ(A#KB)]j[\operatorname{Tor}_{i}^{Q}(A\#_{K}B)]_{j} if j{e+i1,e+i}j\in\{e+i-1,e+i\}. To this end we utilize the fact that A#KBA\#_{K}B is Gorenstein. Thus, its graded minimal free resolution is symmetric. In particular, since dimQ=m+n\dim Q=m+n, one has

[ToriQ(A#KB)]j[Torm+niQ(A#KB)]e+m+nj.[\operatorname{Tor}_{i}^{Q}(A\#_{K}B)]_{j}\cong[\operatorname{Tor}_{m+n-i}^{Q}(A\#_{K}B)]_{e+m+n-j}.

Similarly, for AA and BB we have ToriQ(A)j[Torm+niQ(A)]e+m+nj\operatorname{Tor}_{i}^{Q}(A)_{j}\cong[\operatorname{Tor}_{m+n-i}^{Q}(A)]_{e+m+n-j} and ToriQ(B)j[Torm+niQ(B)]e+m+nj\operatorname{Tor}_{i}^{Q}(B)_{j}\cong[\operatorname{Tor}_{m+n-i}^{Q}(B)]_{e+m+n-j}.

Combined with Theorem 3.2 and using e3e\geq 3, which implies that the degrees e+i1,e+ie+i-1,e+i are not self-dual under the isomorphisms given above, the claim regarding Tor\operatorname{Tor} modules follows.

The Poincaré series formula follows from the above considerations and the identities

i=0m+nβm+ni,m+ni+1(Q/𝐱𝐲)tisi+e1\displaystyle\sum_{i=0}^{m+n}\beta_{m+n-i,m+n-i+1}(Q/\mathbf{x}\cap\mathbf{y})t^{i}s^{i+e-1}
=\displaystyle= j=0m+nβj,j+1(Q/𝐱𝐲)tm+njsm+nj+e1\displaystyle\sum_{j=0}^{m+n}\beta_{j,j+1}(Q/\mathbf{x}\cap\mathbf{y})t^{m+n-j}s^{m+n-j+e-1}
=\displaystyle= tm+nsm+n+ej=0m+nβj,j+1(Q/𝐱𝐲)tjsj1\displaystyle t^{m+n}s^{m+n+e}\sum_{j=0}^{m+n}\beta_{j,j+1}(Q/\mathbf{x}\cap\mathbf{y})t^{-j}s^{-j-1}
=\displaystyle= tm+nsm+n+ePQ/𝐱𝐲Q(t1,s1).\displaystyle t^{m+n}s^{m+n+e}P^{Q}_{Q/\mathbf{x}\cap\mathbf{y}}(t^{-1},s^{-1}).

Using again Lemma 3.1 and Proposition 3.4, we can convert Theorem 3.7 into explicit formulas depending only on the Betti numbers of AA as an RR-module and of BB as an SS-module.

Corollary 3.8.

With the above notation we have:

βi,jQ(A#KB)={0 if ji,(i,j)(0,0),1 if (i,j)=(0,0),=0min(i,m)1(n[im]++)βmin(i,m),i+1[im]+R(A)+=0min(i,n)1(m[in]++)βmin(i,n),i+1[in]+S(B)+(m+ni+1)(mi+1)(ni+1) if j=i+1,=0min(i,m)1(n[im]++)βmin(i,m),j[im]+R(A)+=0min(i,m)1(m[in]++)βmin(i,n),j[in]+S(B) if i+2ji+e2,=0min(i,m)1(n[im]++)βmin(i,m),i+1[im]+R(A)+=0min(i,n)1(m[in]++)βmin(i,n),i+1[in]+S(B)+(m+nm+ni1)(mmi1)(nni1) if im+n,j=i+e1,1 if (i,j)=(m+n,e+m+n),0 if je+i and(i,j)(m+n,e+m+n).\beta_{i,j}^{Q}(A\#_{K}B)=\\ \begin{cases}0&\text{ if }j\leq i,(i,j)\neq(0,0),\\[11.38092pt] 1&\text{ if }(i,j)=(0,0),\\[2.0pt] \sum_{\ell=0}^{min(i,m)-1}\binom{n}{[i-m]_{+}+\ell}\beta_{min(i,m)-\ell,i+1-\ell-[i-m]_{+}}^{R}(A)\\ +\sum_{\ell=0}^{min(i,n)-1}\binom{m}{[i-n]_{+}+\ell}\beta_{min(i,n)-\ell,i+1-\ell-[i-n]_{+}}^{S}(B)\\ +\binom{m+n}{i+1}-\binom{m}{i+1}-\binom{n}{i+1}&\text{ if }j=i+1,\\[11.38092pt] \sum_{\ell=0}^{min(i,m)-1}\binom{n}{[i-m]_{+}+\ell}\beta_{min(i,m)-\ell,j-\ell-[i-m]_{+}}^{R}(A)\\ +\sum_{\ell=0}^{min(i,m)-1}\binom{m}{[i-n]_{+}+\ell}\beta_{min(i,n)-\ell,j-\ell-[i-n]_{+}}^{S}(B)&\text{ if }i+2\leq j\leq i+e-2,\\[14.22636pt] \sum_{\ell=0}^{min(i,m)-1}\binom{n}{[i-m]_{+}+\ell}\beta_{min(i,m)-\ell,i+1-\ell-[i-m]_{+}}^{R}(A)\\ +\sum_{\ell=0}^{min(i,n)-1}\binom{m}{[i-n]_{+}+\ell}\beta_{min(i,n)-\ell,i+1-\ell-[i-n]_{+}}^{S}(B)\\ +\binom{m+n}{m+n-i-1}-\binom{m}{m-i-1}-\binom{n}{n-i-1}&\text{ if }i\leq m+n,j=i+e-1,\\[2.0pt] 1&\text{ if }(i,j)=(m+n,e+m+n),\\[11.38092pt] 0&\text{ if }j\geq e+i\text{ and}\\ &(i,j)\neq(m+n,e+m+n).\end{cases}

3.2. Arbitrary many summands

For i=1,,ri=1,\ldots,r, consider standard graded polynomial rings Ri=K[xi,1,,xi,ni]R_{i}=K[x_{i,1},\ldots,x_{i,n_{i}}] with irrelevant maximal ideals 𝐱i=(xi,1,,xi,ni)\mathbf{x}_{i}=(x_{i,1},\ldots,x_{i,n_{i}}). Also, let Q=R1KKRrQ=R_{1}\otimes_{K}\cdots\otimes_{K}R_{r}. In the following, we abuse notation to write 𝐱i\mathbf{x}_{i} to also denote the extensions of these ideals to ideals of QQ.

Lemma 3.9.

Consider ideals I1,,IrI_{1},\ldots,I_{r} of a commutative ring PP and the map φ:P/j=1rIjφj=1rP/Ij\varphi\colon P/\bigcap_{j=1}^{r}I_{j}\stackrel{{\scriptstyle\varphi}}{{\longrightarrow}}\bigoplus_{j=1}^{r}P/I_{j}, where φ\varphi maps the image of pPp\in P in P/j=1rIjP/\bigcap_{j=1}^{r}I_{j} onto the image of (p,,p)Pr(p,\ldots,p)\in P^{r} in j=1rP/Ij\bigoplus_{j=1}^{r}P/I_{j}. The annihilator of cokerφ\operatorname{coker}\varphi as a PP-module is i=1r(Ii+jiIj)\bigcap_{i=1}^{r}\big{(}I_{i}+\bigcap_{j\neq i}I_{j}\big{)}.

Proof.

First, we prove the inclusion Ann(cokerφ)i=1r(Ii+jiIj)\operatorname{Ann}(\operatorname{coker}\varphi)\subseteq\bigcap_{i=1}^{r}\big{(}I_{i}+\bigcap_{j\neq i}I_{j}\big{)}. Consider aAnn(cokerφ)a\in\operatorname{Ann}(\operatorname{coker}\varphi) and let mm be the image of (1,0,,0)Pr(1,0,\dots,0)\in P^{r} in j=1rP/Ij\bigoplus_{j=1}^{r}P/I_{j}, so that amimφam\in\operatorname{im}\varphi. Then there exists pj1Ijp\in\bigcap_{j\neq 1}I_{j} such that apI1a-p\in I_{1}, which shows that aI1+j1Ija\in I_{1}+\bigcap_{j\neq 1}I_{j}. Repeating the argument with 2ir2\leq i\leq r proves the desired inclusion.

For the reverse inclusion, observe that every aI1+j1Ija\in I_{1}+\bigcap_{j\neq 1}I_{j} annihilates the image of (1,0,,0)Pr(1,0,\dots,0)\in P^{r} in cokerφ\operatorname{coker}\varphi, and similarly for 2ir2\leq i\leq r. ∎

Lemma 3.9 will be applied to the following family of ideals.

Lemma 3.10.

Define the ideals IjI_{j} of QQ by

Ij=𝐱1++𝐱^j++𝐱rI_{j}=\mathbf{x}_{1}+\cdots+\widehat{\mathbf{x}}_{j}+\cdots+\mathbf{x}_{r}

so that Q/IjRjQ/I_{j}\cong R_{j}. Then one has

1ijr𝐱i𝐱j=j=1rIj.\sum_{1\leq i\neq j\leq r}\mathbf{x}_{i}\cap\mathbf{x}_{j}=\bigcap_{j=1}^{r}I_{j}.
Proof.

We proceed by induction on r2r\geq 2, the base case follows immediately from the definitions.

Assume the statement holds for r1r-1, i.e.,

1ijr1𝐱i𝐱j=j=1r1I~j,\sum_{1\leq i\neq j\leq r-1}\mathbf{x}_{i}\cap\mathbf{x}_{j}=\bigcap_{j=1}^{r-1}\tilde{I}_{j},

where I~j=𝐱1++𝐱^j++𝐱r1\tilde{I}_{j}=\mathbf{x}_{1}+\cdots+\widehat{\mathbf{x}}_{j}+\cdots+\mathbf{x}_{r-1}. Then we have

j=1rIj=j=1r1IjIr=j=1r1(I~j+𝐱r)Ir=(j=1r1I~j+𝐱r)Ir=(1ijr1𝐱i𝐱j+𝐱r)Ir=1ijr1𝐱i𝐱j+1ir1𝐱i𝐱r\begin{split}&\bigcap_{j=1}^{r}I_{j}=\bigcap_{j=1}^{r-1}I_{j}\cap I_{r}=\bigcap_{j=1}^{r-1}(\tilde{I}_{j}+\mathbf{x}_{r})\cap I_{r}=\left(\bigcap_{j=1}^{r-1}\tilde{I}_{j}+\mathbf{x}_{r}\right)\cap I_{r}\\ =&\left(\sum_{1\leq i\neq j\leq r-1}\mathbf{x}_{i}\cap\mathbf{x}_{j}+\mathbf{x}_{r}\right)\cap I_{r}=\sum_{1\leq i\neq j\leq r-1}\mathbf{x}_{i}\cap\mathbf{x}_{j}+\sum_{1\leq i\leq r-1}\mathbf{x}_{i}\cap\mathbf{x}_{r}\end{split}

because 𝐱i𝐱jIr\mathbf{x}_{i}\cap\mathbf{x}_{j}\subset I_{r} whenever 1ijr11\leq i\neq j\leq r-1. ∎

Combining Lemma 3.9 with Lemma 3.10, we obtain:

Corollary 3.11.

There is an exact sequence of graded QQ-modules

0Q/(1ijr𝐱i𝐱j)j=1rRjKr10.0\to Q/\left(\sum_{1\leq i\neq j\leq r}\mathbf{x}_{i}\cap\mathbf{x}_{j}\right)\to\bigoplus_{j=1}^{r}R_{j}\to K^{r-1}\to 0. (3.3)
Proof.

Consider the ideal I1,,IrI_{1},\ldots,I_{r} of QQ as defined in Lemma 3.10. Observe that Q/IjRjQ/I_{j}\cong R_{j} and j=1rIj=1ijr𝐱i𝐱j\bigcap_{j=1}^{r}I_{j}=\sum_{1\leq i\neq j\leq r}\mathbf{x}_{i}\cap\mathbf{x}_{j}. Thus, the map in Lemma 3.9 becomes φ:Q/(1ijr𝐱i𝐱j)j=1rRj.\varphi\colon Q/\left(\sum_{1\leq i\neq j\leq r}\mathbf{x}_{i}\cap\mathbf{x}_{j}\right)\to\bigoplus_{j=1}^{r}R_{j}.

The definition of the ideals IjI_{j} implies that for each ii, Ii+jiIjI_{i}+\bigcap_{j\neq i}I_{j} is the maximal ideal 𝔪\mathfrak{m} generated by all the variables of QQ. Hence, Lemma 3.9 shows that Ann(cokerφ)=𝔪\operatorname{Ann}(\operatorname{coker}\varphi)=\mathfrak{m}.

Notice that imφ\operatorname{im}\varphi is a cyclic QQ-module whose minimal generator can also be taken as a minimal generator of j=1rRj\bigoplus_{j=1}^{r}R_{j}. It follows that cokerφ\operatorname{coker}\varphi is minimally generated by r1r-1 elements of degree zero. Since Ann(cokerφ)=𝔪\operatorname{Ann}(\operatorname{coker}\varphi)=\mathfrak{m}, we conclude that cokerφKr1\operatorname{coker}\varphi\cong K^{r-1}, which completes the argument. ∎

We compute the Betti numbers for the leftmost term in the short exact sequence (3.3).

Lemma 3.12.

The ideal ij𝐱i𝐱j\sum_{i\neq j}\mathbf{x}_{i}\cap\mathbf{x}_{j} has a 2-linear minimal free resolution, and for t1t\geq 1 we have that

[TortQ(Q/ij𝐱i𝐱j,K)]t+1\displaystyle\left[\operatorname{Tor}_{t}^{Q}(Q/\sum_{i\neq j}\mathbf{x}_{i}\cap\mathbf{x}_{j},K)\right]_{t+1}\cong
coker([j=1rTort+1Q(Q/Ij,K)]t+1[j=1r1Tort+1Q(K,K)]t+1).\displaystyle\operatorname{coker}\left(\left[\bigoplus_{j=1}^{r}\operatorname{Tor}^{Q}_{t+1}(Q/I_{j},K)\right]_{t+1}\to\left[\bigoplus_{j=1}^{r-1}\operatorname{Tor}^{Q}_{t+1}(K,K)\right]_{t+1}\right).

Thus with N=n1++nrN=n_{1}+\cdots+n_{r},

βt,t+1Q(Q/ij𝐱i𝐱j)=(r1)(Nt+1)k=1r(Nnkt+1).\beta_{t,t+1}^{Q}\left(Q/\sum_{i\neq j}\mathbf{x}_{i}\cap\mathbf{x}_{j}\right)=(r-1)\binom{N}{t+1}-\sum_{k=1}^{r}\binom{N-n_{k}}{t+1}.
Proof.

Since reg(Q/Ij)=0\operatorname{reg}(Q/I_{j})=0 for each 1jr1\leq j\leq r and reg(K)=0\operatorname{reg}(K)=0, the short exact sequence (3.3) implies, by means of the formula

reg(Q/ij𝐱i𝐱j)max{reg(j=1rQ/Ij),reg(Kr1)+1}=1,\operatorname{reg}\left(Q/\sum_{i\neq j}\mathbf{x}_{i}\cap\mathbf{x}_{j}\right)\leq\max\left\{\operatorname{reg}(\bigoplus_{j=1}^{r}Q/I_{j}),\operatorname{reg}(K^{r-1})+1\right\}=1,

that ij𝐱i𝐱j\sum_{i\neq j}\mathbf{x}_{i}\cap\mathbf{x}_{j} has a 2-linear minimal free resolution. Moreover, for every t0t\geq 0, it induces the following long exact sequence. Note that we write TortQ(M)\operatorname{Tor}^{Q}_{t}(M) instead of TortQ(M,K)\operatorname{Tor}^{Q}_{t}(M,K) for a QQ-module MM.

[Tort+1Q(Q/ij𝐱i𝐱j)]t+1[j=1rTort+1Q(Q/Ij)]t+1[j=1r1Tort+1Q(K)]t+1[TortQ(Q/ij𝐱i𝐱j)]t+1[j=1rTortQ(Q/Ij)]t+1.\left[\operatorname{Tor}_{t+1}^{Q}\left(Q/\sum_{i\neq j}\mathbf{x}_{i}\cap\mathbf{x}_{j}\right)\right]_{t+1}\to\left[\bigoplus_{j=1}^{r}\operatorname{Tor}^{Q}_{t+1}(Q/I_{j})\right]_{t+1}\\ \to\left[\bigoplus_{j=1}^{r-1}\operatorname{Tor}^{Q}_{t+1}(K)\right]_{t+1}\to\left[\operatorname{Tor}_{t}^{Q}\left(Q/\sum_{i\neq j}\mathbf{x}_{i}\cap\mathbf{x}_{j}\right)\right]_{t+1}\\ \to\left[\bigoplus_{j=1}^{r}\operatorname{Tor}^{Q}_{t}(Q/I_{j})\right]_{t+1}.

Since reg(ijQ/𝐱i𝐱j)=1\operatorname{reg}(\sum_{i\neq j}Q/\mathbf{x}_{i}\cap\mathbf{x}_{j})=1 and reg(Q/Ij)=0\operatorname{reg}(Q/I_{j})=0, the left-most and right-most modules in the above long exact sequence are zero. Since the QQ-modules Q/IjQ/I_{j} and KK are minimally resolved by Koszul complexes, the dimensions of the second and third terms of the sequence are given by sums of the appropriate binomial coefficients. Therefore, taking the difference of these dimensions yields the desired formula for the Betti numbers. ∎

For every i=1,,ri=1,\dots,r we consider a standard graded ring Ai=Ri/𝔞iA_{i}=R_{i}/\mathfrak{a}_{i} with 𝔞i(𝐱i)2\mathfrak{a}_{i}\subseteq(\mathbf{x}_{i})^{2}. We abuse notation to write 𝔞i\mathfrak{a}_{i} to also denote the extensions of these ideals to ideals of QQ.

Theorem 3.13.

For every t1t\geq 1, we have that

[TortQ(A1×K×KAr,K)]s={0 if st,i=1r[TortQ(Ai,K)]t+1[TortQ(Q/ij𝐱i𝐱j,K)]t+1 if s=t+1,i=1r[TortQ(Ai,K)]s if st+2.\begin{split}&[\operatorname{Tor}_{t}^{Q}(A_{1}\times_{K}\cdots\times_{K}A_{r},K)]_{s}=\\ &\begin{cases}0&\text{ if }s\leq t,\\[2.0pt] \displaystyle\bigoplus_{i=1}^{r}[\operatorname{Tor}_{t}^{Q}(A_{i},K)]_{t+1}\oplus\left[\operatorname{Tor}_{t}^{Q}\left(Q/\sum_{i\neq j}\mathbf{x}_{i}\cap\mathbf{x}_{j},K\right)\right]_{t+1}&\text{ if }s=t+1,\\[2.0pt] \displaystyle\bigoplus_{i=1}^{r}[\operatorname{Tor}_{t}^{Q}(A_{i},K)]_{s}&\text{ if }s\geq t+2.\end{cases}\end{split}
Proof.

Consider the short exact sequence of graded QQ-modules (2.7)

0A1×K×KArA1ArKr10.0\to A_{1}\times_{K}\cdots\times_{K}A_{r}\to A_{1}\oplus\cdots\oplus A_{r}\to K^{r-1}\to 0.

For every t0t\geq 0, it induces the following long exact sequence, where we write TortQ(M)\operatorname{Tor}^{Q}_{t}(M) instead of TortQ(M,K)\operatorname{Tor}^{Q}_{t}(M,K) for a QQ-module MM.

Tort+1Q(Kr1)TortQ(A1×K×KAr)TortQ(A1Ar)TortQ(Kr1).\operatorname{Tor}^{Q}_{t+1}(K^{r-1})\to\operatorname{Tor}^{Q}_{t}(A_{1}\times_{K}\cdots\times_{K}A_{r})\\ \to\operatorname{Tor}^{Q}_{t}(A_{1}\oplus\cdots\oplus A_{r})\to\operatorname{Tor}^{Q}_{t}(K^{r-1}).

We have that [TortQ(Kr1)]s0[\operatorname{Tor}^{Q}_{t}(K^{r-1})]_{s}\neq 0 if and only if 0t=sn1++nr0\leq t=s\leq n_{1}+\cdots+n_{r}. Thus, for every st+2s\geq t+2, we get

[TortQ(A1×K×KAr)]si=1r[TortQ(Ai)]s.[\operatorname{Tor}_{t}^{Q}(A_{1}\times_{K}\cdots\times_{K}A_{r})]_{s}\cong\bigoplus_{i=1}^{r}[\operatorname{Tor}_{t}^{Q}(A_{i})]_{s}.

Restricting to degree s=t+1s=t+1, we get the exact sequence:

i=1r[Tort+1Q(Ai)]t+1[Tort+1Q(Kr1)]t+1[TortQ(A1×K×KAr)]t+1i=1r[TortQ(Ai)]t+10.\bigoplus_{i=1}^{r}[\operatorname{Tor}^{Q}_{t+1}(A_{i})]_{t+1}\to[\operatorname{Tor}^{Q}_{t+1}(K^{r-1})]_{t+1}\\ \to[\operatorname{Tor}_{t}^{Q}(A_{1}\times_{K}\cdots\times_{K}A_{r})]_{t+1}\to\bigoplus_{i=1}^{r}[\operatorname{Tor}^{Q}_{t}(A_{i})]_{t+1}\to 0.

For every 1ir1\leq i\leq r, we have Ai=Q/(Ii,𝔞i)A_{i}=Q/(I_{i},\mathfrak{a}_{i}), and since we assume that each 𝔞i\mathfrak{a}_{i} is generated in degrees at least two, we also have

i=1r[Tort+1Q(Ai)]t+1i=1r[Tort+1Q(Q/Ii)]t+1.\bigoplus_{i=1}^{r}[\operatorname{Tor}^{Q}_{t+1}(A_{i})]_{t+1}\cong\bigoplus_{i=1}^{r}[\operatorname{Tor}^{Q}_{t+1}(Q/I_{i})]_{t+1}.

This implies that

[TortQ(A1×K×KAr)]t+1i=1r[TortQ(Ai)]t+1coker(i=1r[Tort+1Q(Q/Ii)]t+1i=1r1[Tort+1Q(K)]t+1).[\operatorname{Tor}^{Q}_{t}(A_{1}\times_{K}\cdots\times_{K}A_{r})]_{t+1}\cong\bigoplus_{i=1}^{r}[\operatorname{Tor}^{Q}_{t}(A_{i})]_{t+1}\\ \oplus\operatorname{coker}\left(\bigoplus_{i=1}^{r}[\operatorname{Tor}^{Q}_{t+1}(Q/I_{i})]_{t+1}\to\bigoplus_{i=1}^{r-1}[\operatorname{Tor}^{Q}_{t+1}(K)]_{t+1}\right).

Using Lemma 3.12, we get the desired formula. ∎

From the previous result, we can compute the graded Betti numbers of the fiber product A1×K×KArA_{1}\times_{K}\cdots\times_{K}A_{r} in terms of the graded Betti numbers of the AiA_{i} as QQ-modules. A straightforward computation allows us to translate into a formula depending only on the Betti numbers of the AiA_{i} as RiR_{i}-modules.

Corollary 3.14.

With N=n1++nrN=n_{1}+\cdots+n_{r}, we have

PA1×K×KArQ(t,s)\displaystyle P^{Q}_{A_{1}\times_{K}\cdots\times_{K}A_{r}}(t,s) =\displaystyle= i=1r(PAiRi(t,s)1)(1+st)Nni\displaystyle\sum_{i=1}^{r}(P^{R_{i}}_{A_{i}}(t,s)-1)(1+st)^{N-n_{i}}
+(r1)(1+st)NNst1t\displaystyle+(r-1)\frac{(1+st)^{N}-Nst-1}{t}
i=1r(1+st)Nni(Nni)st1t+1.\displaystyle-\sum_{i=1}^{r}\frac{(1+st)^{N-n_{i}}-(N-n_{i})st-1}{t}+1.
Proof.

The generating series is derived from Theorem 3.13, the numerical formula in Lemma 3.12, and an analogue of Proposition 3.4. ∎

Recall that it is harmless to assume that an AG KK-algebra has socle degree at least three because AG algebras with smaller socle degrees are well understood (see the description above Theorem 3.7).

Theorem 3.15.

Assume that A1,,ArA_{1},\ldots,A_{r} are AG KK-algebras with reg(A)=e3\operatorname{reg}(A_{\ell})=e\geq 3 for all 1r1\leq\ell\leq r. Then, for any integer i0i\geq 0, the graded Betti numbers of the connected sum A1#K#KArA_{1}\#_{K}\cdots\#_{K}A_{r} over the polynomial ring QQ with dim(Q)=N\dim(Q)=N are given by

[TortQ(A1#K#KAr,K)]s\displaystyle[\operatorname{Tor}_{t}^{Q}(A_{1}\#_{K}\cdots\#_{K}A_{r},K)]_{s}\cong
{0 if st and (t,s)(0,0),K if (t,s)=(0,0),i=1r[TortQ(Ai,K)]t+1[TortQ(Q/ij𝐱i𝐱j,K)]t+1 if j=i+1,i=1r[TortQ(Ai,K)]s if t+2st+e2,i=1r[TortQ(Ai,K)]s[TorNtQ(Q/ij𝐱i𝐱j,K)]Nt+1 if s=t+e1,K if (t,s)=(N,e+N),0 if se+t and(t,s)(N,e+N).\displaystyle\begin{cases}0&\text{ if }s\leq t\text{ and }(t,s)\neq(0,0),\\[2.0pt] K&\text{ if }(t,s)=(0,0),\\[2.0pt] \bigoplus_{i=1}^{r}[\operatorname{Tor}_{t}^{Q}(A_{i},K)]_{t+1}\\ \oplus[\operatorname{Tor}_{t}^{Q}(Q/\sum_{i\neq j}\mathbf{x}_{i}\cap\mathbf{x}_{j},K)]_{t+1}&\text{ if }j=i+1,\\[2.0pt] \bigoplus_{i=1}^{r}[\operatorname{Tor}_{t}^{Q}(A_{i},K)]_{s}&\text{ if }t+2\leq s\leq t+e-2,\\[2.0pt] \bigoplus_{i=1}^{r}[\operatorname{Tor}_{t}^{Q}(A_{i},K)]_{s}&\\ \oplus[\operatorname{Tor}_{N-t}^{Q}(Q/\sum_{i\neq j}\mathbf{x}_{i}\cap\mathbf{x}_{j},K)]_{N-t+1}&\text{ if }s=t+e-1,\\[2.0pt] K&\text{ if }(t,s)=(N,e+N),\\[2.0pt] 0&\text{ if }s\geq e+t\text{ and}\\ &(t,s)\neq(N,e+N).\end{cases}
Proof.

As before, we write TortQ(M)\operatorname{Tor}_{t}^{Q}(M) instead of TortQ(M,K)\operatorname{Tor}_{t}^{Q}(M,K) for any QQ-module MM. We denote C=A1#K#KArC=A_{1}\#_{K}\cdots\#_{K}A_{r} and D=A1×K×KArD=A_{1}\times_{K}\cdots\times_{K}A_{r}, and consider the exact sequence of graded QQ-modules (2.12)

0Kr1(e)DC0.0\to K^{r-1}(-e)\to D\to C\to 0.

Its long exact Tor\operatorname{Tor} sequence gives exact sequences

[TortQ(Kr1)]se[TortQ(D)]s[TortQ(C)]s[Tort1Q(Kr1)]se.[\operatorname{Tor}_{t}^{Q}(K^{r-1})]_{s-e}\to[\operatorname{Tor}_{t}^{Q}(D)]_{s}\to[\operatorname{Tor}_{t}^{Q}(C)]_{s}\to[\operatorname{Tor}_{t-1}^{Q}(K^{r-1})]_{s-e}.

Since TortQ(K)\operatorname{Tor}_{t}^{Q}(K) is concentrated in degree tt we conclude that

[TortQ(C)]s[TortQ(D)]s[\operatorname{Tor}_{t}^{Q}(C)]_{s}\cong[\operatorname{Tor}_{t}^{Q}(D)]_{s}

if s{e+t1,e+t}s\notin\{e+t-1,e+t\}. Combined with Theorem 3.13,this determines [TortQ(C)]s[\operatorname{Tor}_{t}^{Q}(C)]_{s} if se+t2s\leq e+t-2.

Using that reg(C)=regAi=e\operatorname{reg}(C)=\operatorname{reg}A_{i}=e, we know [TortQ(C)]s=0[\operatorname{Tor}_{t}^{Q}(C)]_{s}=0 if se+t+1s\geq e+t+1. It remains to determine [TortQ(C)]s[\operatorname{Tor}_{t}^{Q}(C)]_{s} if s{e+t1,e+i}s\in\{e+t-1,e+i\}. To this end, we utilize the fact that CC is Gorenstein. Thus, its graded minimal free resolution is symmetric. In particular, one has

[TortQ(C)]s[TorNtQ(C)]e+Ns[\operatorname{Tor}_{t}^{Q}(C)]_{s}\cong[\operatorname{Tor}_{N-t}^{Q}(C)]_{e+N-s}

and similarly, we have TortQ(Ai)s[TorNtQ(A)]e+Ns\operatorname{Tor}_{t}^{Q}(A_{i})_{s}\cong[\operatorname{Tor}_{N-t}^{Q}(A)]_{e+N-s} for each AiA_{i}.

Combined with Theorem 3.13 and using e3e\geq 3, which implies that the degrees e+i1,e+ie+i-1,e+i are not self-dual under the isomorphisms given above, the claim regarding the Tor\operatorname{Tor} modules follows. ∎

Corollary 3.16.

With the notation of Theorem 3.15, we have

PA1#K#KArQ(t,s)=i=1r(PAiRi(t,s)1)(1+st)Nni+1+(r1)(1+st)NNst1ti=1r(1+st)Nni(Nni)st1t+(r1)sN+etN+1[(1+s1t1)NNst1]sN+etN+1i=1r[(1+s1t1)NniNnist1].\begin{split}P^{Q}_{A_{1}\#_{K}\cdots\#_{K}A_{r}}(t,s)&=\sum_{i=1}^{r}(P^{R_{i}}_{A_{i}}(t,s)-1)(1+st)^{N-n_{i}}+1\\ &+(r-1)\frac{(1+st)^{N}-Nst-1}{t}\\ &-\sum_{i=1}^{r}\frac{(1+st)^{N-n_{i}}-(N-n_{i})st-1}{t}\\ &+(r-1)s^{N+e}t^{N+1}[(1+s^{-1}t^{-1})^{N}-\frac{N}{st}-1]\\ &-s^{N+e}t^{N+1}\sum_{i=1}^{r}\left[(1+s^{-1}t^{-1})^{N-n_{i}}-\frac{N-n_{i}}{st}-1\right].\end{split}
Proof.

As a first step, we show

PA1#K#KArQ(t,s)=i=1rPAiQ(t,s)+PQ/ij𝐱i𝐱jQ(t,s)+sN+etNPQ/ij𝐱i𝐱jQ(t1,s1)2.\begin{split}P^{Q}_{A_{1}\#_{K}\cdots\#_{K}A_{r}}(t,s)&=\sum_{i=1}^{r}P^{Q}_{A_{i}}(t,s)+P^{Q}_{Q/\sum_{i\neq j}\mathbf{x}_{i}\cap\mathbf{x}_{j}}(t,s)\\ &+s^{N+e}t^{N}P^{Q}_{Q/\sum_{i\neq j}\mathbf{x}_{i}\cap\mathbf{x}_{j}}(t^{-1},s^{-1})-2.\end{split}

This formula follows from Theorem 3.15 and the identities

u=0NβNu,Nu+1(Q/ij𝐱i𝐱j)tusu+e1\displaystyle\sum_{u=0}^{N}\beta_{N-u,N-u+1}\left(Q/\sum_{i\neq j}\mathbf{x}_{i}\cap\mathbf{x}_{j}\right)t^{u}s^{u+e-1}
=\displaystyle= v=0Nβv,v+1(Q/ij𝐱i𝐱j)tNvsNv+e1\displaystyle\sum_{v=0}^{N}\beta_{v,v+1}\left(Q/\sum_{i\neq j}\mathbf{x}_{i}\cap\mathbf{x}_{j}\right)t^{N-v}s^{N-v+e-1}
=\displaystyle= tNsN+ev=0Nβv,v+1(Q/ij𝐱i𝐱j)tvsv1\displaystyle t^{N}s^{N+e}\sum_{v=0}^{N}\beta_{v,v+1}\left(Q/\sum_{i\neq j}\mathbf{x}_{i}\cap\mathbf{x}_{j}\right)t^{-v}s^{-v-1}
=\displaystyle= tNsN+ePQ/ij𝐱i𝐱jQ(t1,s1).\displaystyle t^{N}s^{N+e}P^{Q}_{Q/\sum_{i\neq j}\mathbf{x}_{i}\cap\mathbf{x}_{j}}(t^{-1},s^{-1}).

Substituting the formulas of Corollary 3.14 and Lemma 3.12 into the formula above yields the claim. ∎

4. Connected Sum as a Doubling

4.1. Motivating examples

We discuss examples of monomial complete intersections. Using the so-called doubling method, Celikbas, Laxmi and Weyman solved a particular case of Questions 1.1 and 1.2. Indeed, in [CLW19, Corollary 6.3], they determine a minimal free resolution of the connected sum of KK-algebras Ai=K[xi]/(xidi)A_{i}=K[x_{i}]/(x_{i}^{d_{i}}) by using the doubling construction. The goal of this section is to generalize their result to AG KK-algebras with the same socle degree. We start with a toy example.

Example 4.1.

The Betti table of the connected sum

C=K[x](x4)#KK[y](y4)#KK[z](z4)C=\frac{K[x]}{(x^{4})}\#_{K}\frac{K[y]}{(y^{4})}\#_{K}\frac{K[z]}{(z^{4})}

is described on the left in Table 3.

0 1 2 3
total 1 3 3 1
0: 1 . . .
1: . 3 2 .
2: . 2 3 .
3: . . . 1
      
0 1 2 3
total 1 3 3 1
0: 1 . . .
1: . 3 2 .
Table 3. Betti tables of R/IR/I and R/JR/J in Example 4.1

It should be understood as follows. The connected sum CC has the presentation

C=K[x,y,z](xy,xz,yz,x3+y3,x3+z3).C=\frac{K[x,y,z]}{(xy,xz,yz,x^{3}+y^{3},x^{3}+z^{3})}.

Let Q=K[x,y,z],I=(xy,xz,yz,x3+y3,x3+z3)Q=K[x,y,z],I=(xy,xz,yz,x^{3}+y^{3},x^{3}+z^{3}) and J=(xy,xz,yz)J=(xy,xz,yz). Then the Betti table of Q/JQ/J is given on the right in Table 3. It follows from this that ωQ/J\omega_{Q/J} has two generators and there is an exact sequence

0ωQ/J(3)Q/JC0,0\to\omega_{Q/J}(-3)\to Q/J\to C\to 0,

which maps the generators of ωQ/J\omega_{Q/J} to the elements x3+y3x^{3}+y^{3} and x3+z3x^{3}+z^{3} in Q/JQ/J. The resolution of CC is obtained as a mapping cone from the previous exact sequence.

Each of the summands in CC is obtained by doubling a polynomial ring. Indeed, the short exact sequence

0ωK[x](3)K[x]K[x](x4)00\to\omega_{K[x]}(-3)\to K[x]\to\frac{K[x]}{(x^{4})}\to 0

sending the generator of ωK[x]K[x](1)\omega_{K[x]}\cong K[x](-1) to x4x^{4}, shows that K[x]/(x4)K[x]/(x^{4}) is a doubling of K[x]K[x]. Similarly, the remaining summands are doublings of K[y]K[y] and K[z]K[z], respectively. Furthermore, the ring Q/JQ/J from above can be identified with the fiber product of the rings being doubled

Q/J=K[x]×KK[y]×KK[z].Q/J=K[x]\times_{K}K[y]\times_{K}K[z].

The following example is the first generalization of the [CLW19, Corollary 6.3] to every monomial complete intersections.

Example 4.2.

We focus on the connected sum A=A1#K#KArA=A_{1}\#_{K}\cdots\#_{K}A_{r} of complete intersection algebras

Ai:=K[xi,1,,xi,ni]/(xi,1di,1,,xi,nidi,ni)A_{i}:=K[x_{i,1},\dots,x_{i,n_{i}}]/(x_{i,1}^{d_{i,1}},\dots,x_{i,n_{i}}^{d_{i,n_{i}}})

for i=1,,ri=1,\dots,r, satisfying

j=1nidi,jni=j=1nidi,jni whenever 1i,ir.\sum_{j=1}^{n_{i}}d_{i,j}-n_{i}=\sum_{j=1}^{n_{i^{\prime}}}d_{i^{\prime},j}-n_{i^{\prime}}\;\text{ whenever }1\leq i,i^{\prime}\leq r. (4.1)

Let Ri=K[xi,1,,xi,ni]R_{i}=K[x_{i,1},\dots,x_{i,n_{i}}], Q=R1KKRrQ=R_{1}\otimes_{K}\cdots\otimes_{K}R_{r} and let cc be the quantity defined in (4.1). The connected sum of the KK-algebras AiA_{i} admits the presentation AQ/IA\cong Q/I where

I\displaystyle I =(xi,jixh,jh| 1i<hr,1jini,1jhnh)\displaystyle=\left(x_{i,j_{i}}x_{h,j_{h}}\,|\,1\leq i<h\leq r,1\leq j_{i}\leq n_{i},1\leq j_{h}\leq n_{h}\right)
+(xi,lidi,li| 1ir,1lini1)\displaystyle+\left(x_{i,l_{i}}^{d_{i,l_{i}}}\,\middle|\,1\leq i\leq r,1\leq l_{i}\leq n_{i}-1\right)
+(xi,1di,11xi,nidi,ni1+x1,1d1,11x1,n1d1,n11| 2ir).\displaystyle+\left(x_{i,1}^{d_{i,1}-1}\cdots x_{i,n_{i}}^{d_{i,n_{i}}-1}+x_{1,1}^{d_{1,1}-1}\cdots x_{1,n_{1}}^{d_{1,n_{1}}-1}\,\middle|\,2\leq i\leq r\right).

It can be verified that AA is a doubling of A~=Q/J\tilde{A}=Q/J, where JJ is an ideal defining rr coordinate points in 𝔸Kn1××𝔸Knr\mathbb{A}_{K}^{n_{1}}\times\cdots\times\mathbb{A}_{K}^{n_{r}} with multiplicity; more precisely, J=i=1,,rJiJ=\bigcap_{i=1,\ldots,r}J_{i}, where

Ji=(xj,h,xi,lidi,li|ji,1hnj,1lini1).J_{i}=\left(x_{j,h},x_{i,l_{i}}^{d_{i,l_{i}}}\,\middle|\,j\neq i,1\leq h\leq n_{j},1\leq l_{i}\leq n_{i}-1\right).

More importantly, setting Ai~=Ri/(xi,lidi,li| 1lini1)\tilde{A_{i}}=R_{i}/(x_{i,l_{i}}^{d_{i,l_{i}}}\,|\,1\leq l_{i}\leq n_{i}-1), we see that each ring AiA_{i} is a doubling of Ai~\tilde{A_{i}} via the sequence

0ωAi~(c)Ai~Ai00\to\omega_{\tilde{A_{i}}}(-c)\to\tilde{A_{i}}\to A_{i}\to 0

sending the generator of ωAi~(c)Ai~(di,ni)\omega_{\tilde{A_{i}}}(-c)\cong\tilde{A_{i}}(-d_{i,n_{i}}) to xi,nidi,nix_{i,n_{i}}^{d_{i,n_{i}}}, and that A~=A1~×K×KAr~\tilde{A}=\tilde{A_{1}}\times_{K}\cdots\times_{K}\tilde{A_{r}}. The Betti numbers of A~\tilde{A} can thus be obtained via Corollary 3.14.

We shall explain this observation as part of a general phenomenon in the following result.

Theorem 4.3.

Let A1,,ArA_{1},\ldots,A_{r} be graded AG KK-algebras with reg(Ai)=d\operatorname{reg}(A_{i})=d for all 1i,jr1\leq i,j\leq r. Suppose that for each 1ir1\leq i\leq r, AiA_{i} is a doubling of some 1-dimensional Cohen-Macaulay algebra A~i\tilde{A}_{i}, then the connected sum A1#K#KArA_{1}\#_{K}\cdots\#_{K}A_{r} is a doubling of A~1×K×KA~r\tilde{A}_{1}\times_{K}\cdots\times_{K}\tilde{A}_{r}.

Proof.

We proceed by induction on rr. We first prove the base case where r=2r=2. Set A~1=R/𝔞~1\tilde{A}_{1}=R/\tilde{\mathfrak{a}}_{1} and A~2=S/𝔞~2\tilde{A}_{2}=S/\tilde{\mathfrak{a}}_{2} and let Q=R1KR2Q=R_{1}\otimes_{K}R_{2}. By [AAM12, Lemma 1.5] the ring A~1×KA~2\tilde{A}_{1}\times_{K}\tilde{A}_{2} is Cohen Macaulay of dimension one. By Lemma 2.19, our assumptions imply that for each ii we have exact sequences

0ωA~i(d)A~iAi0.0\to\omega_{{\tilde{A}}_{i}}(-d)\to\tilde{A}_{i}\to A_{i}\to 0. (4.2)

Considering these in degree zero we conclude that

[ωA~i]d=0.[\omega_{{\tilde{A}}_{i}}]_{-d}=0. (4.3)

Combining the exact sequences (4.2) for i{1,2}i\in\{1,2\} with the sequence in (2.3), we obtain the following commutative diagram of QQ-modules with exact rows and middle column.

00(ωA~1ωA~2)(d)=(ωA~1ωA~2)(d)0A~1×KA~2σA~1A~2μK0=0A1×KA2A1A2K000\displaystyle\begin{CD}00\\ @V{}V{}V@V{}V{}V\\ (\omega_{\tilde{A}_{1}}\oplus\omega_{\tilde{A}_{2}})(-d)@>{}>{=}>(\omega_{\tilde{A}_{1}}\oplus\omega_{\tilde{A}_{2}})(-d)\\ @V{}V{}V@V{}V{}V\\ 0@>{}>{}>\tilde{A}_{1}\times_{K}\tilde{A}_{2}@>{\sigma}>{}>\tilde{A}_{1}\oplus\tilde{A}_{2}@>{\mu}>{}>K@>{}>{}>0\\ @V{}V{}V@V{}V{}V@V{}V{=}V\\ 0@>{}>{}>A_{1}\times_{K}A_{2}@>{}>{}>A_{1}\oplus A_{2}@>{}>{}>K@>{}>{}>0\\ @V{}V{}V@V{}V{}V\\ 00\\ \end{CD} (4.4)

The vertical map A~1×KA~2A1×KA2\tilde{A}_{1}\times_{K}\tilde{A}_{2}\to A_{1}\times_{K}A_{2} in (4.4) is uniquely determined by viewing A1×KA2A_{1}\times_{K}A_{2} as a pullback in the category of KK-algebras and utilizing the universal property of this categorical construction. Moreover, by the snake lemma, the kernel of this map is the module (ωA~1ωA~2)(d)(\omega_{\tilde{A}_{1}}\oplus\omega_{\tilde{A}_{2}})(-d).

Applying the functor Hom(,Q)\operatorname{Hom}(-,Q) to the diagram (4.4) yields a new commutative diagram (4.5). The middle row in (4.5) comes from the top of (4.4), and the top row of (4.5) contains the non-vanishing Ext\operatorname{Ext} modules for the QQ-modules in the middle row of (4.4). According to Remark 2.2, the map marked ν\nu satisfies ν(1)=(τA1,τA2)\nu(1)=(\tau_{A_{1}},\tau_{A_{2}}) after identifying ωA1ωA2A1A2\omega_{A_{1}}\oplus\omega_{A_{2}}\cong A_{1}\oplus A_{2}.

000KωA~1×KA~2ωA~1ωA~20η(A~1A~2)(d)=(A~1A~2)(d)χ0ωA1×KA2ωA1ωA2K000\displaystyle\setcounter{MaxMatrixCols}{11}\begin{CD}00\\ @V{}V{}V@V{}V{}V\\ 0@<{}<{}<K@<{}<{}<\omega_{\tilde{A}_{1}\times_{K}\tilde{A}_{2}}@<{}<{}<\omega_{\tilde{A}_{1}}\oplus\omega_{\tilde{A}_{2}}@<{}<{}<0\\ @V{}V{\eta}V@V{}V{}V\\ (\tilde{A}_{1}\oplus\tilde{A}_{2})(d)@<{}<{=}<(\tilde{A}_{1}\oplus\tilde{A}_{2})(d)\\ @V{}V{}V@V{}V{\chi}V\\ 0@<{}<{}<\omega_{A_{1}\times_{K}A_{2}}@<{}<{}<\omega_{A_{1}}\oplus\omega_{A_{2}}@<{}<{}<K@<{}<{}<0\\ @V{}V{}V@V{}V{}V\\ 00\end{CD} (4.5)

The snake lemma applied to (4.5) yields a connecting isomorphism θ:KK\theta\colon K\to K. Let sωA~1×KA~2s\in\omega_{\tilde{A}_{1}\times_{K}\tilde{A}_{2}} be such that ξ(s)=θ(1)\xi(s)=\theta(1). Then χ(η(s))=ν(1)\chi(\eta(s))=\nu(1) can be identified with (τA1,τA2)A1A2(\tau_{A_{1}},\tau_{A_{2}})\in A_{1}\oplus A_{2}, that is, η(s)\eta(s) is equivalent to (τA1,τA2)(\tau_{A_{1}},\tau_{A_{2}}) modulo the image of ωA~1ωA~2\omega_{\tilde{A}_{1}}\oplus\omega_{\tilde{A}_{2}}.

We want to compare the image of

η[d]:ωA~1×KA~2(d)A~1A~2\eta[-d]\colon\omega_{\tilde{A}_{1}\times_{K}\tilde{A}_{2}}(-d)\to\tilde{A}_{1}\oplus\tilde{A}_{2}

and the kernel of the map μ\mu from Diagram (4.4). The image of η[d]\eta[-d] is trivial in degree zero by Equation (4.3). Since KK is concentrated in degree zero, the map μ\mu has zero image in every degree other than zero. It follows that the image of η[d]\eta[-d] is contained in kerμ=imσA~1×KA~2\ker\mu=\operatorname{im}\sigma\cong\tilde{A}_{1}\times_{K}\tilde{A}_{2}. Hence η[d]\eta[-d] induces an injective graded QQ-module homomorphism

δ:ωA~1×KA~2(d)A~1×KA~2.\delta\colon\omega_{\tilde{A}_{1}\times_{K}\tilde{A}_{2}}(-d)\to\tilde{A}_{1}\times_{K}\tilde{A}_{2}.

Its existence proves that ωA~1×KA~2(d)\omega_{\tilde{A}_{1}\times_{K}\tilde{A}_{2}}(-d) can be identified with an ideal of A~1×KA~2\tilde{A}_{1}\times_{K}\tilde{A}_{2}.

The following diagram combines the left column of Diagram (4.4) and the top row of (4.5). By previous considerations indicating that δ(s)=η(s)\delta(s)=\eta(s) is equivalent to (τA1,τA2)(\tau_{A_{1}},\tau_{A_{2}}) modulo the image of ωA~1ωA~2\omega_{\tilde{A}_{1}}\oplus\omega_{\tilde{A}_{2}}, the diagram commutes provided that ξ(s)\xi(s) is mapped by τ\tau to (τA1,τA2)A1×KA2(\tau_{A_{1}},\tau_{A_{2}})\in A_{1}\times_{K}A_{2}. With this choice, the cokernel of τ\tau is A1#KA2A_{1}\#_{K}A_{2} by Definition 2.10.

00(ωA~1ωA~1)(d)=(ωA~1ωA~2)(d)0ωA~1×KA~2(d)δA~1×KA~2C0ξ0K(d)τA1×KA2A1#KA2000\displaystyle\small{\begin{CD}00\\ @V{}V{}V@V{}V{}V\\ (\omega_{\tilde{A}_{1}}\oplus\omega_{\tilde{A}_{1}})(-d)@>{=}>{}>(\omega_{\tilde{A}_{1}}\oplus\omega_{\tilde{A}_{2}})(-d)\\ @V{}V{}V@V{}V{}V\\ 0@>{}>{}>\omega_{\tilde{A}_{1}\times_{K}\tilde{A}_{2}}(-d)@>{}>{\delta}>\tilde{A}_{1}\times_{K}\tilde{A}_{2}@>{}>{}>C@>{}>{}>0\\ @V{}V{\xi}V@V{}V{}V\\ 0@>{}>{}>K(-d)@>{}>{\tau}>A_{1}\times_{K}A_{2}@>{}>{}>A_{1}\#_{K}A_{2}@>{}>{}>0\\ @V{}V{}V@V{}V{}V\\ 00\\ \end{CD}}

Setting CC be the cokernel of δ\delta, the snake lemma yields an isomorphism CA1#KA2C\cong A_{1}\#_{K}A_{2}. This shows that A1#KA2A_{1}\#_{K}A_{2} is a doubling of A~1×KA~2\tilde{A}_{1}\times_{K}\tilde{A}_{2}, as desired for the base case of induction.

Now, we assume that the AG KK-algebra A1#K#KAr1A_{1}\#_{K}\cdots\#_{K}A_{r-1} is a doubling of A~1×K×KA~r1\tilde{A}_{1}\times_{K}\cdots\times_{K}\tilde{A}_{r-1}. The base case applied to AG KK-algebras A1#K#KAr1A_{1}\#_{K}\cdots\#_{K}A_{r-1} and ArA_{r} implies by way of Remarks 2.8 and 2.15 that A1#K#KArA_{1}\#_{K}\cdots\#_{K}A_{r} is a doubling of A~1×K×KA~r\tilde{A}_{1}\times_{K}\cdots\times_{K}\tilde{A}_{r} completing the proof. ∎

Theorem 4.3 generalizes [CLW19, Theorem 5.5], which considered the case of AG algebras A1,,ArA_{1},\ldots,A_{r} of embedding dimension one, establishing an analogous doubling result.


References

  • [AAM12] H. Ananthnarayan, L. Avramov, and W. F. Moore. Connected sums of Gorenstein local rings. J. Reine Angew. Math., 667:149–176, 2012.
  • [BH93] W. Bruns and J. Herzog. Cohen-Macaulay rings, volume 39 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1993.
  • [CGS23] E. Celikbas, H. Geller, and T. Se. Classifying betti numbers of fiber products, arXiv:2307.05715, 2023.
  • [CLW19] E. Celikbas, J. Laxmi, and J. Weyman. Embeddings of canonical modules and resolutions of connected sums. J. Pure Appl. Algebra, 223(1):175–187, 2019.
  • [CN09] A. Corso and U. Nagel. Monomial and toric ideals associated to Ferrers graphs. Trans. Amer. Math. Soc., 361(3):1371–1395, 2009.
  • [Gel22] H. Geller. Minimal free resolutions of fiber products. Proc. Amer. Math. Soc., 150(10):4159–4172, 2022.
  • [IK99] A. Iarrobino and V. Kanev. Power sums, Gorenstein algebras, and determinantal loci, volume 1721 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1999. Appendix C by Iarrobino and Steven L. Kleiman.
  • [IMS22] A. Iarrobino, C. McDaniel, and A. Seceleanu. Connected sums of graded Artinian Gorenstein algebras and Lefschetz properties. J. Pure Appl. Algebra, 226(1):Paper No. 106787, 52, 2022.
  • [KKR+21] G. Kapustka, M. Kapustka, K. Ranestad, H. Schenck, M. Stillman, and B. Yuan. Quaternary quartic forms and Gorenstein rings, arXiv:2111.05817, 2021.
  • [KMMR+01] J. O. Kleppe, J. C. Migliore, R. M. Miró-Roig, U. Nagel, and C. Peterson. Gorenstein liaison, complete intersection liaison invariants and unobstructedness. Mem. Amer. Math. Soc., 154(732):viii+116, 2001.
  • [Mac94] F. S. Macaulay. The algebraic theory of modular systems. Cambridge Mathematical Library. Cambridge University Press, Cambridge, 1994. Revised reprint of the 1916 original, With an introduction by Paul Roberts.
  • [Mas91] W. S. Massey. A basic course in algebraic topology, volume 127 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1991.
  • [Sal79] J. D. Sally. Stretched Gorenstein rings. J. London Math. Soc. (2), 20(1):19–26, 1979.