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

\newaliascnt

lemthm \aliascntresetthelem \newaliascntcorthm \aliascntresetthecor \newaliascntpropthm \aliascntresettheprop \newaliascntdfnthm \aliascntresetthedfn \newaliascntremthm \aliascntresettherem

Syzygies of the residue field over Golod rings

Đoàn Trung Cường Institute of Mathematics, Vietnam Academy of Science and Technology, 18 Hoang Quoc Viet, 10072 Hanoi, Viet Nam. [email protected] Hailong Dao Department of Mathematics, University of Kansas, Lawrence, KS 66045-7523, USA [email protected] https://www.math.ku.edu/ hdao/ David Eisenbud Department of Mathematics, University of California, Berkley, CA, 94720 [email protected] eisenbud.github.io Toshinori Kobayashi Department of Mathematics, School of Science and Technology, Meiji University, 1-1-1 Higashi-mita, Tama-ku, Kawasaki 214-8571, Japan. [email protected] Claudia Polini Department of Mathematics, Notre Dame University, South Bend, IN [email protected]  and  Bernd Ulrich Department of Mathematics, Purdue University, W. Lafayette, IN [email protected]
Abstract.

Let (R,𝔪,k)(R,\mathfrak{m},k) be a Golod ring. We show a recurrence formula for high syzygies of kk in terms of previous ones. In the case of embedding dimension at most 22, we provided complete descriptions of all indecomposable summands of all syzygies of kk.

Key words and phrases:
syzygy, resolutions
2010 Mathematics Subject Classification:
13C13, 13D09, 13H10
HD was partly supported by Simons Foundation grant MP-TSM-00002378. DE is grateful to the National Science Foundation for partial support through grant 2001649. CP and BU were partially supported by NSF grants DMS-2201110 and DMS-2201149, respectively. DTC was funded by Vingroup Joint Stock Company and supported by Vingroup Innovation Foundation (VinIF) under the project code VINIF.2021.DA00030. This material is partly based upon work supported by the National Science Foundation under Grant No. DMS-1928930, while four of the authors were in residence at the Mathematical Sciences Research Institute in Berkeley, California, during the Special Semester in Commutative Algebra, Spring 2024.

1. Introduction

Let (R,𝔪,k)(R,\mathfrak{m},k) be a local ring of embedding dimension ee. We prove that if RR is Golod then every syzygy module is a direct sum of copies of the first e+1e+1 syzygy modules, and give a simple recurrence relation for the number of each one. This result is a surprisingly easy consequence of Golod’s 1962 description of the resolution of kk [2], but seems not to have been noticed previously.

Theorem 1.1.

Let (R,𝔪,k)(R,\mathfrak{m},k) be a local Noetherian ring of embedding dimension ee. Let KK_{\bullet} be the Koszul complex of a set of minimal generators of 𝔪\mathfrak{m}. If RR is Golod then

syze+1R(k)=i=0e1syziR(k)hei{\rm syz}_{e+1}^{R}(k)=\bigoplus_{i=0}^{e-1}{\rm syz}_{i}^{R}(k)^{h_{e-i}}

and, more generally,

syze+1+jR(k)=i=je1+jsyzi+jR(k)hei+j{\rm syz}_{e+1+j}^{R}(k)=\bigoplus_{i=j}^{e-1+j}{\rm syz}_{i+j}^{R}(k)^{h_{e-i+j}}

where hi=dimk(Hi(K)).h_{i}=\dim_{k}(H_{i}(K_{\bullet})).

The hypothesis of Theorem 1.1 applies to every local ring of embedding dimension 2 except for 0-dimensional complete intersections (over a 0-dimensional complete intersection (or any Gorenstein ring) every syzygy of kk is indecomposable).

Example 1.2.

If (R,𝔪,k)(R,\mathfrak{m},k) is an artinian ring of embedding dimension 2 and type t>1t>1 then, by Theorem 1.1,

syz3R(k)=kt1t+1𝔪.{\rm syz}_{3}^{R}(k)=k^{t}\oplus\oplus_{1}^{t+1}\mathfrak{m}.

whereas if t=1t=1 then, by Tate’s theorem [4], syz3R(k){\rm syz}_{3}^{R}(k) is an indecomposable module with 5 generators.

The first e+1 syzygies decompose in a more complex way, but when the embedding dimension is 2, we can analyze the decompositions completely:

Theorem 1.3.

Let (R,𝔪,k)(R,\mathfrak{m},k) be a Noetherian local ring. If RR has embedding dimension 2\leq 2 and is not a zero-dimensional complete intersection then every minimal RR-syzygy of kk is a direct sum of copies of kk, 𝔪\mathfrak{m} and 𝔪:=HomR(𝔪,R)=syz1R(𝔪)\mathfrak{m}^{*}:=\operatorname{Hom}_{R}(\mathfrak{m},R)={\rm syz}_{1}^{R}(\mathfrak{m}). In particular, the number of distinct indecomposable modules that appear in all syzygies of kk is at most 33.

We can further analyze the indecomposability of each syzygy:

Theorem 1.4.

Let (S,𝔫)(S,\mathfrak{n}) be a regular local ring of dimension 2, and let II be an ideal contained in 𝔫2\mathfrak{n}^{2}. Let R=S/IR=S/I and 𝔪=𝔫R\mathfrak{m}=\mathfrak{n}R.

  1. (1)

    syz1R(k)=𝔪{\rm syz}^{R}_{1}(k)=\mathfrak{m} is indecomposable if and only if xyIxy\not\in I for any x,yx,y with 𝔫=(x,y)\mathfrak{n}=(x,y) or RR is a zero-dimensional complete intersection.

  2. (2)

    syz2R(k){\rm syz}^{R}_{2}(k) is indecomposable if and only if 𝔪\mathfrak{m} is indecomposable and (I:𝔫)𝔫=I𝔫(I:\mathfrak{n})\mathfrak{n}=I\mathfrak{n}.

  3. (3)

    syz3R(k){\rm syz}_{3}^{R}(k) is indecomposable if and only if II is a principal ideal such that xyIxy\not\in I for any x,yx,y with 𝔫=(x,y)\mathfrak{n}=(x,y) or RR is a zero-dimensional complete intersection.

  4. (4)

    syziR(k){\rm syz}_{i}^{R}(k) for all i0i\geq 0 are indecomposable if and only if syz3R(k){\rm syz}_{3}^{R}(k) is indecomposable.

The theorems will be proved in section 5. Preparatory results, some holding in more general contexts, are given in sections 2, 3 and 4.

Our results were suggested by Macaulay 2 computations [3], performed at AIM meeting September 2023 with the help of Mahrud Sayrafi and Devlin Mallory, using their DirectSummands package.

2. syz1R(𝔪){\rm syz}^{R}_{1}(\mathfrak{m}) is 𝔪\mathfrak{m}^{*}

Theorem 2.1.

If (R,𝔪)(R,\mathfrak{m}) is a local ring of embedding dimension 2 that is not a 0-dimensional complete intersection, then syz1R(𝔪)𝔪{\rm syz}_{1}^{R}(\mathfrak{m})\cong\mathfrak{m}^{*}.

We postpone the proof until after Theorem 2.3.

Theorem 2.2.

Let SS be a ring and let IJSI\subset J\subset S be ideals. Set R=S/IR=S/I and write ():=HomS(,R)(-)^{*}:=\operatorname{Hom}_{S}(-,R) for the RR-dual. The folowing conditions are equivalent:

  1. (1)

    The dual (JR)J(JR)^{*}\to J^{*} of the natural surjection JJRJ\to JR is an isomorphism

  2. (2)

    The restriction map JIJ^{*}\to I^{*} is 0.

  3. (3)

    The natural map ExtR1(S/J,R)ExtS1(S/J,R)\operatorname{Ext}^{1}_{R}(S/J,R)\to\operatorname{Ext}^{1}_{S}(S/J,R) is an isomorphism.

If these conditions are satisfied for an ideal JJ that contains II then they are satisfied for any ideal containing JJ.

Proof.

121\Longleftrightarrow 2: Dualizing the exact sequence

0IJ(J/I=JR)00\to I\to J\to(J/I=JR)\to 0

yields the result.

131\Longleftrightarrow 3: We have a diagram

{diagram}.\begin{diagram}.

Dualizing into RR, we get the diagram

{diagram}.\begin{diagram}.

The equivalence now follows from the “five lemma”.

The last statement follows at once from condition (2). ∎

Lemma \thelem.

With notation as in Theorem 2.2, if JJ is generated by a regular sequence of length 2 then syz1R(JR)J{\rm syz}_{1}^{R}(JR)\cong J^{*}.

Proof.

Suppose that JJ is generated by x,yx,y and consider the maps

S\rTo(yx)S2\rTo(ba)R.S\rTo^{\begin{pmatrix}y\\ -x\end{pmatrix}}S^{2}\rTo^{\begin{pmatrix}b&-a\end{pmatrix}}R.

The composition is 0 if and only if ax+byIax+by\in I, and since JJ is the cokernel of (yx)\begin{pmatrix}y\\ -x\end{pmatrix}, this is the condition that (b,a)(b,-a) induces a homomorphism JRJ\to R. ∎

Theorem 2.3.

Let SS be a Noetherian local ring and let IJI\subset J be ideals. Write R=S/IR=S/I and ()=HomS(,R)(-)^{*}=\operatorname{Hom}_{S}(-,R). Assume that II has projective dimension 1\leq 1 and JJ is generated by an SS-regular sequence x,yx,y of 22 elements.

  1. (1)

    If the equivalent conditions (1)(3)(1)-(3) of Theorem 2.2 hold then syz1R(JR)(JR){\rm syz}_{1}^{R}(JR)\cong(JR)^{*} via a map that sends the restriction to JRJR of the identity RRR\to R to the relation (y,x)(y,-x) on x,yx,y over RR.

  2. (2)

    If II is a principal ideal then conditions (1)(3)(1)-(3) of Theorem 2.2 hold.

  3. (3)

    If II is not principal suppose and I(gcd(I))1JI\cdot(\gcd(I))^{-1}\subset J then conditions (1)(3)(1)-(3) of Theorem 2.2 are equivalent to the condition that gcd(I)Fitt2(I)J\gcd(I)\cdot{\rm Fitt}_{2}(I)\subset J.

Proof of Theorem 2.3.

For part (1)(1), we use section 2 with (a,b)=(x,y)(a,b)=(x,y), and the condition that J(JR)J^{*}\cong(JR)^{*} from Theorem 2.2.

For parts (2)(2) and (3)(3) we must show that the restriction map JIJ^{*}\to I^{*} is 0 whenever II is principal, and, when II is not principal, that it is 0 if and only if gcd(I)Fitt2(I)J\gcd(I)\cdot{\rm Fitt}_{2}(I)\subset J.

Let FF be the minimal SS-free resolution of R=S/IR=S/I, let KK be the Koszul complex resolving S/JS/J, and let γ:FK\gamma:F\to K be the comparison map lifting the identity K0=S=F0K_{0}=S=F_{0}. The restriction map is the left vertical map in the diagram

{diagram}\begin{diagram}

where κ\kappa is the RR-dual of the Koszul differential, and ϵ\epsilon is the composition of γ1R\gamma_{1}\otimes R with an isomorphism K1RK1K_{1}\otimes R\to K_{1}^{*} induced by the isomorphism 2(K1R)R\wedge^{2}(K_{1}\otimes R)\to R.

Since all the rows of this diagram are exact, the restriction map JIJ^{*}\to I^{*} is zero if and only if the composition γ1(κ,ϵ)\gamma_{1}^{*}\circ(\kappa,\epsilon) is 0. Now γ1κ\gamma_{1}^{*}\circ\kappa is the dual of the map (F1S)R=0(F_{1}\to S)\otimes R=0, so the restriction map is 0 if and only if γ1ϵ=0\gamma_{1}^{*}\circ\epsilon=0.

One checks using the cofactor expansion of the determinant that the composition γ1ϵ:F1RF1\gamma_{1}^{*}\circ\epsilon:F_{1}\otimes R\to F_{1}^{*} corresponds to the composition

(F1R)2\rTo2(F1R)\rTo2γ1R2(K1R)R,(F_{1}\otimes R)^{\otimes 2}\rTo\wedge^{2}(F_{1}\otimes R)\rTo^{\wedge^{2}\gamma_{1}\otimes R}\wedge^{2}(K_{1}\otimes R)\cong R,

so that γ1ϵ\gamma_{1}^{*}\circ\epsilon is 0 if and only if

()the 2×2 minors of γ1 are in I.(*)\ \hbox{the $2\times 2$ minors of $\gamma_{1}$ are in $I$.}

Condition (*) clearly holds if II is principal. Thus we may assume that II is not principal. Write a=gcd(I)a=\gcd(I) and suppose that I:=Ia1I^{\prime}:=I\cdot a^{-1} is minimally generated by g1,,gng_{1}^{\prime},\dots,g_{n}^{\prime}. We may harmlessly assume that gi,gjg_{i}^{\prime},g_{j}^{\prime} form a regular sequence for any iji\neq j and that gi,ag_{i}^{\prime},a form a regular sequence for any i.i. Let FF^{\prime} be the minimal SS-free resolution of S/IS/I^{\prime} and let γ:FK\gamma^{\prime}:F^{\prime}\to K be the comparison map lifting the identity K0=S=F0K_{0}=S=F_{0}^{\prime}, which exists because IJI^{\prime}\subset J. We may assume that the map γ1\gamma_{1} is multiplication by aa composed with γ1.\gamma^{\prime}_{1}. Let Δi,j\Delta_{i,j}^{\prime} be the minor of γ1\gamma_{1}^{\prime} involving columns i,j.i,j. Condition (*) means that

aΔi,jIforanyij.a\cdot\Delta_{i,j}^{\prime}\in I^{\prime}\ \ {\rm for\ any}\ i\neq j\,.

Since gi,agjg_{i}^{\prime},ag_{j}^{\prime} form a regular sequence contained in JJ, we have

(gi,agj):J=(gi,agj,aΔi,j),(g_{i}^{\prime},ag_{j}^{\prime}):J=(g_{i}^{\prime},ag_{j}^{\prime},a\Delta_{i,j}^{\prime}),

and as II^{\prime} is perfect of grade 2, it follows that aΔi,jIa\cdot\Delta_{i,j}^{\prime}\in I^{\prime} if and only if (gi,agj):IJ(g_{i}^{\prime},ag_{j}^{\prime}):I^{\prime}\subset J by the symmetry of linkage.

By a theorem of Gaeta, the link of II^{\prime} with respect to gi,agjg_{i}^{\prime},ag_{j}^{\prime} is aa times the ideal generated by the n2n-2 minors of the presentation matrix of III^{\prime}\cong I with rows ii and jj deleted. This proves the equivalence of condition (*) with the condition aFitt2(I)Ja\cdot{\rm Fitt}_{2}(I)\subset J. ∎

Proof of Theorem 2.1.

We apply Theorem 2.3 with JR=𝔪JR=\mathfrak{m}. If II is principal then the result follows from Theorem 2.3, parts (1) and (2). If II is not principal then the result holds by parts (1) and (3) unless both gcd(I)\gcd(I) and Fitt2(I){\rm Fitt}_{2}(I) are unit ideals. If gcd(I)=(1)\gcd(I)=(1) then II has codimension 2. If in addition Fitt2(I)=(1){\rm Fitt}_{2}(I)=(1) then II is a 0-dimensional complete intersection. ∎

3. The decomposition of 𝔪\mathfrak{m}

In this section (S,𝔫,k)(S,\mathfrak{n},k) denotes a regular local ring of dimension 2 and ISI\subset S denotes an ideal contained in 𝔫2\mathfrak{n}^{2}. We write R=S/IR=S/I and 𝔪=𝔫R\mathfrak{m}=\mathfrak{n}R.

Lemma \thelem.

Suppose that 𝔫=(x,y)\mathfrak{n}=(x,y). Each ideal II containing xyxy can be written as

  1. (1)

    I=(xy,uxa+vyb)I=(xy,ux^{a}+vy^{b}) where u,vu,v are each either 0 or units; or

  2. (2)

    (xy,xa,yb)(xy,x^{a},y^{b}),

where a,ba,b are non-negative integers.

Proof.

If II has codimension 1 then we can remove a common divisor, which we may take to be xx. In this case either I=(xy)I=(xy) or I=(xy,xa)I=(xy,x^{a}) for some a1a\geq 1 because S/(y)S/(y) is a discrete valuation ring with parameter xx.

Any element of an artinian local ring can be written as a polynomial in the generators of the maximal ideal with unit coefficients. In particular, if II has codimension 2, then any element of S/(xy,𝔫I)S/(xy,\mathfrak{n}I) can be written in the form f=uxa+vybf=ux^{a}+vy^{b} where each of uu and vv is either 0 or a unit of SS. Note that if u0u\neq 0 then, modulo xyxy, every xmx^{m} with a<ma<m is a multiple of ff and similarly for yy.

If I/(xy)I/(xy) is principal, then we may write I=(xy,uxa+vyb)I=(xy,ux^{a}+vy^{b}), and we are done. Otherwise, modulo xyxy, we may write two generators for II as {uxa+vyb,pxc+qyd}\{ux^{a}+vy^{b},px^{c}+qy^{d}\} where uu and qq are units and aa and dd are minimal. Thus we may assume that p=0p=0, and v=0v=0, and thus I:=(xy,xa,yd)I:=(xy,x^{a},y^{d}). ∎

Theorem 3.1.

The following are equivalent:

  1. (1)

    The module 𝔪\mathfrak{m} is decomposable.

  2. (2)

    We may write 𝔫=(x,y)\mathfrak{n}=(x,y) with xyIxy\in I and RR is not a zero-dimensional complete intersection.

  3. (3)

    We may write 𝔫=(x,y)\mathfrak{n}=(x,y) in such a way that I=(xy,uxa,vyb)I=(xy,ux^{a},vy^{b}) where each of u,vu,v is a unit of SS or 0 and a,b2a,b\geq 2.

Proof.

(1)(2)(1)\Rightarrow(2) In codimension 2\leq 2, a ring is Gorenstein if and only if it is a complete intersection. If 𝔪\mathfrak{m} is decomposable, then it has to decompose as 𝔪=xRyR\mathfrak{m}=xR\oplus yR where 𝔪=(x,y)\mathfrak{m}=(x,y). This implies xyIxy\in I. Every nonzero ideal of a zero-dimensional local Gorenstein ring contains the socle, and thus is indecomposable.

(2)(3)(2)\Rightarrow(3) This follows from section 3.

(3)(1)(3)\Rightarrow(1) One easily check that (I,x)(I,y)=I(I,x)\cap(I,y)=I. ∎

4. The decomposition of the syzygy of 𝔪\mathfrak{m}

Again in this section (S,𝔫,k)(S,\mathfrak{n},k) denotes a regular local ring of dimension 2 and ISI\subset S denotes an ideal contained in 𝔫2\mathfrak{n}^{2}. We write R=S/IR=S/I and 𝔪=𝔫R\mathfrak{m}=\mathfrak{n}R.

Proposition \theprop.

If 𝔪\mathfrak{m} is decomposable then syz1R(𝔪)=ka𝔪{\rm syz}_{1}^{R}(\mathfrak{m})=k^{\oplus a}\oplus\mathfrak{m}, where:

  1. (1)

    If RR is Artinian, then a=2a=2.

  2. (2)

    If depthR=0\operatorname{depth}R=0 and dimR=1\dim R=1 then a=1a=1.

  3. (3)

    If RR is Cohen-Macaulay of dimension 1 then a=0a=0.

Proof.

We may assume 𝔪=xRyR\mathfrak{m}=xR\oplus yR. In case (1) according to Theorem 3.1 we can write I=(xa,xy,yb)I=(x^{a},xy,y^{b}) and xRR/(xa1,y)xR\cong R/(x^{a-1},y) and yRR/(x,yb1)yR\cong R/(x,y^{b-1}). Thus syzR(𝔪)=(xa1,y)R(x,yb1)R{\rm syz}^{R}(\mathfrak{m})=(x^{a-1},y)R\oplus(x,y^{b-1})R. Both xa1,yb1x^{a-1},y^{b-1} are in the socle of RR, and the claim follows.

In case (2) Theorem 3.1 shows that we may write R=S/(xy,xa)R=S/(xy,x^{a}). In this case xRR/(y,xa1)xR\cong R/(y,x^{a-1}), and yRR/xRyR\cong R/xR, so syz1R(𝔪)=(y,xa1)RxRyRkxR𝔪k{\rm syz}_{1}^{R}(\mathfrak{m})=(y,x^{a-1})R\oplus xR\cong yR\oplus k\oplus xR\cong\mathfrak{m}\oplus k since xa1Rx^{a-1}R is the socle of RR.

Finally, if RR is Cohen-Macaulay of dimension 1, then I=(xy)I=(xy) and all of the modules syziR(k){\rm syz}_{i}^{R}(k) for i1i\geq 1 are isomorphic. ∎

Lemma \thelem.

Suppose that 𝔪\mathfrak{m} is indecomposable. If RR is not a complete intersection then:

  1. (1)

    I+𝔫3I+\mathfrak{n}^{3} does not contain any element xyxy such that 𝔫=(x,y)\mathfrak{n}=(x,y).

  2. (2)

    dimk(I+𝔫3)/𝔫31\dim_{k}(I+\mathfrak{n}^{3})/\mathfrak{n}^{3}\leq 1.

  3. (3)

    There exists a choice of generators x0,y0x_{0},y_{0} of 𝔫\mathfrak{n} and a choice of minimal generators of syz1R(𝔪){\rm syz}_{1}^{R}(\mathfrak{m})

    L=(L0L1Lt):=(y0f1ftx0g1gt)L=\begin{pmatrix}L_{0}&L_{1}&\dots&L_{t}\end{pmatrix}:=\begin{pmatrix}y_{0}&f_{1}&\dots&f_{t}\\ -x_{0}&g_{1}&\dots&g_{t}\end{pmatrix}

    such that the entries of every column of the form L0+i>0λiLiL_{0}+\sum_{i>0}\lambda_{i}L_{i} generate 𝔪\mathfrak{m}.

Proof.

(1) Suppose first that dimR=1\dim R=1. If II contained an element of order 2 then since II cannot be principal we could write I=(x)+xII=(x\ell)+xI^{\prime} where \ell has order 1. By Theorem 3.1, \ell must be a unit times xx, and then without loss of generality, II^{\prime} would be generated by yby^{b}. Theorem 3.1 implies that b2b\geq 2, completing the proof in the 1-dimensional case.

Now suppose that RR is 0-dimensional. Suppose that II contains an element xy+fxy+f with order ordf3\operatorname{ord}f\geq 3 such that 𝔫=(x,y)\mathfrak{n}=(x,y). If 𝔫pI\mathfrak{n}^{p}\subset I and ordfp\operatorname{ord}f\geq p then xyIxy\in I, which is impossible, since then 𝔪\mathfrak{m} is decomposable by Theorem 3.1. Otherwise, suppose there is an expression xy+fIxy+f\in I such that 𝔫=(x,y)\mathfrak{n}=(x,y) with order ordf\operatorname{ord}f maximal and <p<p.

We may write f=xf1+yf2+gf=xf_{1}+yf_{2}+g with min{ordf1,ordf2}ordf1\min\{\operatorname{ord}f_{1},\operatorname{ord}f_{2}\}\geq\operatorname{ord}f-1 and ordg>ordf\operatorname{ord}g>\operatorname{ord}f. Thus xy+f=(x+f2)(y+f1)+(gf1f2)xy+f=(x+f_{2})(y+f_{1})+(g-f_{1}f_{2}). Note that ordgf1f2>ordf\operatorname{ord}g-f_{1}f_{2}>\operatorname{ord}f. We may replace x,yx,y by x+f2,y+f1x+f_{2},y+f_{1}, thus increasing the order of ff, a contradiction.

(2) Suppose on the contrary that ax2+bxy+cy2,ax2+bxy+cy2ax^{2}+bxy+cy^{2},a^{\prime}x^{2}+b^{\prime}xy+c^{\prime}y^{2} are linearly independent elements of I+𝔫3/𝔫3I+\mathfrak{n}^{3}/\mathfrak{n}^{3} where x,yx,y are generators of 𝔫/𝔫3\mathfrak{n}/\mathfrak{n}^{3} and the coefficients a,,ca,\dots,c^{\prime} are in kk. By taking a linear combination, we may assume that a=0a=0, in which case we are done by part (1) unless also b=0b=0. It follows that c0c\neq 0, so we may assume that c=0c^{\prime}=0. Now we are done unless b=0b^{\prime}=0, so x2x^{2} and y2y^{2} are in I+𝔫3/𝔫3I+\mathfrak{n}^{3}/\mathfrak{n}^{3}, but xyxy is not. Thus the associated graded ring of RR, and with it RR itself, is a zero-dimensional complete intersection, a contradiction. This shows that dimk(I+𝔫3)/𝔫31\dim_{k}(I+\mathfrak{n}^{3})/\mathfrak{n}^{3}\leq 1, completing the proof.

(3) By part (2) the quotient (I+𝔫3)/𝔫3(I+\mathfrak{n}^{3})/\mathfrak{n}^{3} has dimension 0 or 1. If dimk(I+𝔫3)/𝔫3=1\dim_{k}(I+\mathfrak{n}^{3})/\mathfrak{n}^{3}=1, and a corresponding generator is a perfect square 2\ell^{2} modulo 𝔫3\mathfrak{n}^{3}, then we choose generators x0=,y0x_{0}=\ell,y_{0} for 𝔫\mathfrak{n}. Otherwise we make an arbitrary choice. Also, we may choose generators h1,hth_{1}\dots,h_{t} of II such that all but possibly h1h_{1} are in 𝔫3\mathfrak{n}^{3}, and syzygies so that hi=fix+giyh_{i}=f_{i}x+g_{i}y and fi,gi𝔪2f_{i},g_{i}\in\mathfrak{m}^{2} for i>1i>1. Moreover, if h1x02h_{1}\equiv x_{0}^{2} modulo 𝔫3\mathfrak{n}^{3} we choose L1L_{1} to be the column with entries x0,0x_{0},0.

Now consider the 2×22\times 2 matrix

L:=(L0L1):=(y0i>0λifix0i>0λigi).L^{\prime}:=\begin{pmatrix}L_{0}&L_{1}^{\prime}\end{pmatrix}:=\begin{pmatrix}y_{0}&\sum_{i>0}\lambda_{i}f_{i}\\ -x_{0}&\sum_{i>0}\lambda_{i}g_{i}\end{pmatrix}.

If λ1𝔫\lambda_{1}\in\mathfrak{n}, or dimk(I+𝔫3)/𝔫3=0\dim_{k}(I+\mathfrak{n}^{3})/\mathfrak{n}^{3}=0 then L0+L1L0L_{0}+L_{1}^{\prime}\equiv L_{0} modulo 𝔪2R2\mathfrak{m}^{2}R^{2}, and the claim follows. Thus we can assume that λ1\lambda_{1} is a unit and the determinant of LL^{\prime} is a minimal generator of II not in 𝔫3\mathfrak{n}^{3}. This is also the determinant of the 2×22\times 2 matrix (L0+L1L1)(L_{0}+L_{1}^{\prime}\ \ L_{1}^{\prime}). If the entries of L0+L1L_{0}+L_{1}^{\prime} were linearly dependent, then the determinant would factor; and by part (2) it would have to be the perfect square 2=x02\ell^{2}=x_{0}^{2}. But then, modulo 𝔪2\mathfrak{m}^{2} the matrix LL^{\prime} must be

(y0x0x00).\begin{pmatrix}y_{0}&x_{0}\\ -x_{0}&0\end{pmatrix}.

Thus the entries of L0+L1L_{0}+L_{1}^{\prime} are linearly independent as claimed. ∎

Theorem 4.1.

Write syz1R(𝔪)=kaN{\rm syz}_{1}^{R}(\mathfrak{m})=k^{\oplus a}\oplus N where NN has no kk-summands. If 𝔪\mathfrak{m} is indecomposable then:

  1. (1)

    μ(N)=μ(I)+1a\mu(N)=\mu(I)+1-a. If RR is Gorenstein then a=0a=0; otherwise a=dimk(𝔫(I:𝔫)𝔫I)a=\dim_{k}\left(\frac{\mathfrak{n}(I:\mathfrak{n})}{\mathfrak{n}I}\right).

  2. (2)

    NN is indecomposable.

Proof.

If RR were Gorenstein then syzygies of indecomposable maximal Cohen-Macaulay modules are indecomposable, so we may assume that RR is not Gorenstein.

We write μ(I)\mu(I) for the minimal number of generators of II and similarly for μ(N)\mu(N).

Proof of (1): Since μ(syz1R(𝔪))=μ(I)+1\mu({\rm syz}_{1}^{R}(\mathfrak{m}))=\mu(I)+1, we have μ(N)=μ(I)+1a\mu(N)=\mu(I)+1-a.

We fix generators x,yx,y of 𝔫\mathfrak{n} and the corresponding embedding Z:=syz1R(𝔪)R2.Z:={\rm syz}_{1}^{R}(\mathfrak{m})\subset R^{2}\,. Notice that

a=dimk(SocZ𝔪ZSocZ).a=\dim_{k}\left(\frac{\operatorname{Soc}Z}{\mathfrak{m}Z\cap\operatorname{Soc}Z}\right)\,.

Thus it suffices to prove that

SocZ𝔪ZSocZ𝔫(I:𝔫)𝔫I.\frac{\operatorname{Soc}Z}{\mathfrak{m}Z\cap\operatorname{Soc}Z}\cong\frac{\mathfrak{n}(I:\mathfrak{n})}{\mathfrak{n}I}\,.

To this end we define an RR-linear map ψ\psi as the composition of the maps

SocZZRL0=H1(x,y;R)I𝔫I.\operatorname{Soc}Z\longrightarrow\frac{Z}{RL_{0}}=H_{1}(x,y;R)\stackrel{{\scriptstyle\sim}}{{\longrightarrow}}\frac{I}{\mathfrak{n}I}\,.

Notice that ψ((f,g))=(xF+yG)+𝔫I\psi((f,g))=(xF+yG)+\mathfrak{n}I, where FF, GG are preimages of f,gf,g in SS. As SocZ=SocR2\operatorname{Soc}Z=\operatorname{Soc}R^{2}, it follows that Imψ=𝔫(I:𝔫)𝔫I\operatorname{Im}\psi=\frac{\mathfrak{n}(I:\mathfrak{n})}{\mathfrak{n}I}. Clearly Kerψ=RL0SocZ\operatorname{Ker}\psi=RL_{0}\cap\operatorname{Soc}Z.

Thus it remains to prove

RL0SocZ=𝔪ZSocZ.RL_{0}\cap\operatorname{Soc}Z=\mathfrak{m}Z\cap\operatorname{Soc}Z\,.

The right hand side is in the left hand side because 𝔪ZRL0\mathfrak{m}Z\subset RL_{0}. As to the converse, the indecomposibility of 𝔪\mathfrak{m} implies that I𝔫2I\not=\mathfrak{n}^{2}, hence 𝔪L00\mathfrak{m}L_{0}\not=0. Therefore RL0SocZ𝔪L0𝔪ZRL_{0}\cap\operatorname{Soc}Z\subset\mathfrak{m}L_{0}\subset\mathfrak{m}Z.

Proof of (2): Let LL be the minimal syzygy matrix chosen as in section 4. Note that syz1R(𝔪)/RL0{\rm syz}_{1}^{R}(\mathfrak{m})/RL_{0} is the Koszul homology of x0,y0x_{0},y_{0} in RR, so it is annihilated by 𝔪\mathfrak{m}. Since, by definition, syz1R(𝔪)=Nka{\rm syz}_{1}^{R}(\mathfrak{m})=N\oplus k^{a} we have 𝔪syz1R(𝔪)=𝔪N\mathfrak{m}\,{\rm syz}_{1}^{R}(\mathfrak{m})=\mathfrak{m}N. Finally, since L0L_{0} is a minimal generator of syz1R(𝔪){\rm syz}_{1}^{R}(\mathfrak{m}), it cannot be contained in 𝔪syz1R(𝔪)\mathfrak{m}\,{\rm syz}_{1}^{R}(\mathfrak{m}). Putting this together, we have

𝔪/(0:𝔪)𝔪L0=𝔪syz1R(𝔪)=𝔪N.\mathfrak{m}/(0:\mathfrak{m})\cong\mathfrak{m}\,L_{0}=\mathfrak{m}\,{\rm syz}_{1}^{R}(\mathfrak{m})=\mathfrak{m}\,N.

Since NN does not have kk as a direct summand, the indecomposability of NN would follow from the indecomposability of 𝔪N𝔪/(0:𝔪)\mathfrak{m}N\cong\mathfrak{m}/(0:\mathfrak{m}), so it suffices to treat the cases where the maximal ideal 𝔪/(0:𝔪)\mathfrak{m}/(0:\mathfrak{m}) of S/(I:𝔫)S/(I:\mathfrak{n}) is decomposable. By Theorem 3.1 this is the case if and only if I:𝔫=(xy,uxα,vyβ)I:\mathfrak{n}=(xy,ux^{\alpha},vy^{\beta}) where each of u,vu,v is a unit or 0, and both of α,β\alpha,\beta are 2\geq 2.

Suppose first that dimR=1\dim R=1. In this case we may assume that I:𝔫=(xy,uxα)I:\mathfrak{n}=(xy,ux^{\alpha}). Assume that α3\alpha\geq 3. Set I:=𝔫(I:𝔫)I^{\prime}:=\mathfrak{n}\,(I:\mathfrak{n}). We have II(xy,uxα)I^{\prime}\subset I\subset(xy,ux^{\alpha}). By Theorem 3.1 the ideal II contains no product of two elements that generate the maximal ideal of SS so no element of the form xy+λxαxy+\lambda x^{\alpha} can be in II. Thus I𝔫(xy)+(uxα)=(x2y,xy2,uxα)=:I′′I\subset\mathfrak{n}(xy)+(ux^{\alpha})=(x^{2}y,xy^{2},ux^{\alpha})=:I^{\prime\prime}. If u=0u=0 then I=II=I^{\prime}. If uu is a unit then I′′/IkI^{\prime\prime}/I^{\prime}\cong k. Because xα1I:𝔫x^{\alpha-1}\notin I:\mathfrak{n}, the ideal II cannot equal I′′I^{\prime\prime} , so in any case I=II=I^{\prime}. In this case a=dimk(I/𝔫I)a=\dim_{k}(I^{\prime}/\mathfrak{n}I) is the minimal number of generators of II. Thus NN is cyclic, hence indecomposable.

Now, suppose that dimR=0\dim R=0 or I:𝔫=(xy,x2)I:\mathfrak{n}=(xy,x^{2}). We first show that the module NN can be generated by 2 elements.

If I:𝔫=(xy,x2)I:\mathfrak{n}=(xy,x^{2}), as in the proof above I=x(x,y)2I(x2y,xy2,x2)=x(y2,x)I^{\prime}=x(x,y)^{2}\subset I\subset(x^{2}y,xy^{2},x^{2})=x(y^{2},x). Thus either I=II=I^{\prime} and as above NN is cyclic or I=x(y2,x)I=x(y^{2},x) in which case a=1a=1 hence μ(N)=2\mu(N)=2.

If α=β=2\alpha=\beta=2 then 𝔫3I𝔫2\mathfrak{n}^{3}\subset I\subset\mathfrak{n}^{2}. By section 4, either I=𝔫3I=\mathfrak{n}^{3} hence a=μ(I)a=\mu(I), thus NN is cyclic, or I=𝔫3+(q)I=\mathfrak{n}^{3}+(q), where q𝔫2𝔫3q\in\mathfrak{n}^{2}\setminus\mathfrak{n}^{3}. In the latter case a=μ(I)1a=\mu(I)-1, so that μ(N)=2\mu(N)=2.

Finally, without loss of generality, we can assume that α3\alpha\geq 3. In this case following the same argument as in dimR=1\dim R=1, we can show that (x2y,xy2,xα+1,yβ+1)I(x2y,xy2,xα,yβ)(x^{2}y,xy^{2},x^{\alpha+1},y^{\beta+1})\subset I\subsetneq(x^{2}y,xy^{2},x^{\alpha},y^{\beta}). Thus I=(x2y,xy2,xα+1,yβ+1,uxα+vyβ)I=(x^{2}y,xy^{2},x^{\alpha+1},y^{\beta+1},ux^{\alpha}+vy^{\beta}) where each of u,vu,v is a unit or 0, α3\alpha\geq 3, and β2\beta\geq 2. Thus aμ(I)1a\geq\mu(I)-1, so μ(N)2\mu(N)\leq 2.

Since 𝔪\mathfrak{m} is indecomposable, I𝔫2I\not=\mathfrak{n}^{2} and thus 0:𝔪𝔪0:\mathfrak{m}\subsetneq\mathfrak{m}. Every minimal set of generators of syz1R(𝔪){\rm syz}_{1}^{R}(\mathfrak{m}) contains an element of the form L0+i>0λiLiL_{0}+\sum_{i>0}\lambda_{i}L_{i}, whose annihilator is exactly 0:𝔪0:\mathfrak{m} by section 4(3). This generator cannot be among the minimal generators of kak^{a}, so every minimal set of generators of NN contains such an element.

If N=ABN=A\oplus B with A,B0A,B\not=0 then AA and BB must be cyclic. We may assume that AA is minimally generated by an element of the form L0+i>0λiLiL_{0}+\sum_{i>0}\lambda_{i}L_{i} and thus AR/(0:𝔪)A\cong R/(0:\mathfrak{m}). In particular 𝔪A𝔪B=𝔪N𝔪A\mathfrak{m}A\oplus\mathfrak{m}B=\mathfrak{m}N\cong\mathfrak{m}A. This implies that that 𝔪B=0\mathfrak{m}B=0 for instance because the number of generators of 𝔪B\mathfrak{m}B is 0. Since NN does not have kk as a direct summand, B=0B=0, and we are done. ∎

5. Proofs of Theorems 1.1, 1.3, and 1.4

Proof of Theorem 1.1.

The formula for the differential dd in the minimal RR-free resolution of kk, in terms of the Massey operations μ\mu [1, Theorem 5.2.2] has the form:

d(avi1vip)=davi1vip±a(j=1pμ(vi1..vij)vij+1vip).d(a\otimes v_{i_{1}}\otimes...\otimes v_{i_{p}})=da\otimes v_{i_{1}}\otimes...\otimes v_{i_{p}}\pm a\left(\sum_{j=1}^{p}\mu(v_{i_{1}}..v_{i_{j}})\otimes v_{i_{j+1}}\otimes...\otimes v_{i_{p}}\right).

The only term that does not preserve the last tensor factor, vipv_{i_{p}}, is the one that maps to the Koszul complex alone. Thus the truncation of the complex beyond the Koszul complex splits as a direct sum of complexes, and the summands are the resolutions of

He(K)kk=He(K)ksyz0(k),He1(K)ksyz1(k),,H1(K)ksyze1(k)H_{e}(K)\otimes_{k}k=H_{e}(K)\otimes_{k}{\rm syz}_{0}(k),\ H_{e-1}(K)\otimes_{k}{\rm syz}_{1}(k),\dots,H_{1}(K)\otimes_{k}{\rm syz}_{e-1}(k)

where ee is the embedding dimension. ∎

Proof of Theorem 1.3.

From Theorem 2.1 we have the first assertion. All indecomposable summands will be those appearing in k,𝔪,syz1R(m)k,\mathfrak{m},{\rm syz}_{1}^{R}(m), so the last assertion follows from section 4 and Theorem 4.1. ∎

Proof of Theorem 1.4.

Part (1) follows from Theorem 3.1. For part (2), notice that if 𝔪\mathfrak{m} is decomposable then syz1R(𝔪){\rm syz}^{R}_{1}(\mathfrak{m}) is decomposable. Now the assertion follows from section 4 and Theorem 4.1.

For part (3), we may assume that RR is Gorenstein: if syz2R(𝔪){\rm syz}_{2}^{R}(\mathfrak{m}) is indecomposable then this is true by 1.2, and for the opposite implication, it is part of the assumptions. Since 𝔪\mathfrak{m} is a maximal Cohen-Macaulay RR-module and RR is Gorenstein, 𝔪\mathfrak{m} is indecomposable if and only if one or all of its syzygies are indecomposable. Part (3) now follows from part (1). ∎

References

  • [1] Avramov, L. L. Infinite free resolutions [mr1648664]. In Six lectures on commutative algebra, Mod. Birkhäuser Class. Birkhäuser Verlag, Basel, 2010, pp. 1–118.
  • [2] Golod, E. S. Homologies of some local rings. Dokl. Akad. Nauk SSSR 144 (1962), 479–482.
  • [3] Grayson, D. R., and Stillman, M. E. Macaulay2, a software system for research in algebraic geometry. Available at http://www2.macaulay2.com.
  • [4] Tate, J. Homology of Noetherian rings and local rings. Illinois J. Math. 1 (1957), 14–27.