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

Transferring algebra structures on complexes

Claudia Miller C.M.
    Mathematics Department, Syracuse University, Syracuse, NY 13244, U.S.A.
clamille@syr.edu http://clamille.mysite.syr.edu
 and  Hamidreza Rahmati H.R.
    Department of Mathematics, University of Nebraska, Lincoln, NE 68588, USA
hrahmati2@unl.edu
Abstract.

We discuss a homological method for transferring algebra structures on complexes along suitably nice homotopy equivalences, including those obtained after an application of the Perturbation Lemma. We study the implications for the Homotopy Transfer Theorems under such homotopy equivalences.

As an application, we discuss how to use the homotopy on a Koszul complex given by a scaled de Rham map to find a new method for building a dg algebra structure on a well-known resolution, obtaining one that is both concrete and permutation invariant.

C. Miller partially supported by the National Science Foundation (DMS-1003384), Syracuse University Small Grant, and Douglas R. Anderson Faculty Scholar Fund.

Introduction

In this paper we study descent of algebra structures on complexes along a suitably nice maps, discussing a method that allows one to transfer an algebra structure on a complex to another complex. More precisely, we show that a differential graded (dg) algebra structure on one complex can be transferred to another complex that is a deformation retract of it with a homotopy that satisfies a generalized version of the Leibniz rule and another mild hypothesis that is frequently imposed; see Proposition 1.4.

For our main application, a homological tool called the Perturbation Lemma is the key. For this purpose, we also prove that the analogous dg algebra descent result holds even after applying the Perturbation Lemma as long as the complex remains a dg algebra after perturbation; see Proposition 3.5.

Such descent results turn out to be instances of the well-known Homotopy Transfer Theorem (HTT) which usually yields only an AA_{\infty}-algebra structure on the retract. Our extra hypothesis on the homotopy ensures that the product on the retract is, in fact, associative. We also make a detour to show that with these hypotheses the descended higher AA_{\infty}-operations vanish as well and that if one starts with an AA_{\infty}-algebra instead of a dg algebra then under an analogous stronger condition on the homotopy, the descended higher operations are much simpler than usual; see Propositions 2.5 and  2.7.

We are mainly interested in algebra structures on minimal free resolutions of algebras. It is worth noting that few such resolutions are known to carry dg algebra structures. And yet having one provides one with a powerful tool. Short resolutions are known to have a dg algebra structure [Her74], [BE77], [KM80], [Kus87] but counterexamples of longer ones can be found in [Avr81], [Sri92], [Sri96]. We refer the reader to [Avr98] for a full discussion.

In the second portion of the paper, we apply our descent results to obtain a new method of building dg algebra structures that are both concrete and highly symmetric on some well-known resolutions. These resolutions, constructed by Buchsbaum and Eisenbud in [BE75] using Schur modules, are the minimal graded free resolutions 𝕃a\mathbb{L}_{a} of the quotients R/𝔪aR/\mathfrak{m}^{a} of a polynomial ring RR in nn variables by the aath powers of the homogeneous maximal ideal.

These resolutions were shown to have a dg algebra structure by several authors. The first was Srinivasan in 1989 who put an explicit product using Young tableaux; see [Sri89]. Next in 1996 Peeva proved in [Pee96] that one can place a dg algebra structure on the Eliahou-Kervaire resolution, which applies here since the powers of the maximal ideal are Borel-fixed. In [Mae01] Maeda used the representation theory of the symmetric group SnS_{n} to show that in characteristic zero any SnS_{n}-invariant lift of the multiplication on the quotient ring to the resolution is automatically associative, but did not give any explicit formulas. Our goal is to find a concrete product that is also SnS_{n}-invariant; however, we pay the price of that by having more terms and requiring coefficients in the rational numbers.

In contrast to the others, we use our homotopy transfer results to define a product which is both explicit and very natural in that it is descended from a truncation of a Koszul complex and naturally SnS_{n}-invariant. One benefit of defining a product using our method is that it explains why there should really be a dg algebra structure here and is hopefully more canonical and hence useful for further applications. Another is that it enables us to define dg algebra homomorphisms between these resolutions. Our product works both in characteristic zero and in positive characteristics larger than n+an+a. For this, we use a scaled version of the deRham differential to produce a contracting homotopy for the Koszul complex on the variables that satisfies the generalized Leibniz rule; see Lemma 4.9. It is also worth mentioning that the product that we define can be constructed in a basis free fashion; see Remark 4.13.

More precisely, we prove the following results. In the first one, using the Perturbation Lemma to obtain the resolution of 𝕃a\mathbb{L}_{a} of R/𝔪aR/\mathfrak{m}^{a} as a retract of a complex with a dg algebra structure, we transfer the algebra structure as follows. For this, let 𝕊={ΛiSji,j0}\mathbb{S}=\{\Lambda^{i}\otimes S_{j}\mid i,j\geq 0\} be the totalization of the double complex defined in (4.1), and let 𝕏a\mathbb{X}_{a} be the quotient of the dg algebra 𝕊\mathbb{S} by the dg ideal tra(𝕊)={ΛiSjja}\textrm{tr}_{\geq a}(\mathbb{S})=\{\Lambda^{i}\otimes S_{j}\mid j\geq a\}.

Theorem 4.10. Let aa be a positive integer. Suppose that kk is a field of characteristic zero or positive characteristic pa+np\geq a+n. Consider the deformation retract

𝕏ap[]i----𝕃a\mathbb{X}_{a}\stackrel{{\scriptstyle[}}{{p}}_{\infty}]{i_{\infty}}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}\mathbb{L}_{a}

obtained in (4.6). Defining the product of α,β𝕃a\alpha,\beta\in\mathbb{L}_{a} by

αβ=p(i(α)i(β))\alpha\beta=p_{\infty}\left(i_{\infty}(\alpha)i_{\infty}(\beta)\right)

yields a dg algebra structure on 𝕃a\mathbb{L}_{a}. Furthermore, with this structure the map ii_{\infty} is a homomorphism of dg algebras.

The second result gives some very natural dg algebra homomorphisms between these resolutions.

Theorem 5.1. Let aa and bb be positive integers with bab\geq a and let

πb,a:𝕏b𝕏a\pi_{b,a}\colon\mathbb{X}_{b}\twoheadrightarrow\mathbb{X}_{a}

be the natural surjection. The chain map

fb,a=pπb,ai:𝕃b𝕃af_{b,a}=p_{\infty}\pi_{b,a}i_{\infty}\colon\mathbb{L}_{b}\to\mathbb{L}_{a}

is a homomorphism of dg algebras that gives a lifting of the natural surjection R/𝔪bR/𝔪aR/\mathfrak{m}^{b}\longrightarrow R/\mathfrak{m}^{a}. In particular, the Koszul complex on the variables, which is 𝕃1\mathbb{L}_{1}, is a dg algebra over 𝕃b\mathbb{L}_{b} for every positive integer bb.

Moreover, if cbac\geq b\geq a then fc,a=fb,afc,bf_{c,a}=f_{b,a}f_{c,b}.

The paper is organized as follows: In Section 1, we prove the general statement for descending dg algebra structures along a special deformation retract whose homotopy satisfies the generalized Leibniz rule; see Proposition 1.4. In Section 2, we make a detour to study the implications of the additional hypothesis of generalized Leibniz-type rules for the Homotopy Transfer Theorems, resulting in Propositions 2.5 and  2.7. In Section 3, we recall the well-known Perturbation Lemma and prove a more general version of our dg algebra descent result from Section 1, namely Proposition 3.5. Section 4 contains our main application, which is to obtain dg algebra structures on the resolutions 𝕃a\mathbb{L}_{a} of Buchsbaum and Eisenbud for all a1a\geq 1. In Section 5 we obtain dg algebra homomorphisms between these resolutions.

In this paper, we assume that the complexes consist of RR-modules for some commutative ring RR and that they are graded homologically, rather than cohomologically; the reader should be aware that most sources for AA_{\infty} algebras use the latter instead. One more note: All double complexes in this paper are considered to be anticommutative as in [Wei94], and hence their totalizations do not require any change of sign in the differentials. When we speak of a double complex (for example, when we discuss a dg algebra structure on it) we mean the totalization of it as a complex; which way we are viewing it should be clear from the context.

1. Transfer of dg algebra structures

In this section, we show how to transfer dg algebra structures along certain homotopy equivalences, namely deformation retracts whose associated homotopy behaves well with respect to products. We compare this in the next section to the Homotopy Transfer Theorem, via which the dg algebra structure descends to an AA_{\infty}-algebra structure. We also discuss there what happens when the original complex is merely an AA_{\infty}-algebra.

1.1 Definition.

A differential graded algebra over RR (dg algebra) is a complex (X,)(X,\partial) of RR-modules lying in nonnegative degrees equipped with a product given by a chain map

XRXX,(α,β)αβX\otimes_{R}X\to X,~{}(\alpha,\beta)\to\alpha\beta

giving an associative and unitary product with 1X01\in X_{0}.

The fact that the product is a chain map is equivalent to the differentials of XX satisfying the Leibniz rule:

(αβ)=(α)β+(1)|α|α(β),for allα,βX\partial(\alpha\beta)=\partial(\alpha)\beta+(-1)^{|\alpha|}\alpha\partial(\beta),~{}\text{for all}~{}\alpha,\beta\in X

where |α||\alpha| denotes the degree of α\alpha. In addition, we assume that the product is strictly graded commutative, that is, αβ=(1)|α||β|βα\alpha\beta=(-1)^{|\alpha||\beta|}\beta\alpha for all α,βX\alpha,\beta\in X and α2=0\alpha^{2}=0 if the degree of α\alpha is odd.

A homomorphism of dg algebras is a morphism of complexes ϕ:XX\phi\colon X\to X^{\prime} such that ϕ(1)=1\phi(1)=1 and ϕ(αβ)=ϕ(α)ϕ(β)\phi(\alpha\beta)=\phi(\alpha)\phi(\beta).

We now recall the definitions of the main ingredients for transferring algebra structures.

1.2 Definition.

A set of homotopy equivalence data between two chain complexes is the following set of information: quasi-isomorphisms of complexes

(X,X)p[]i----(Y,Y)(X,\partial^{X})\stackrel{{\scriptstyle[}}{{p}}]{i}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}(Y,\partial^{Y})

with ip1ip\simeq 1 via a chosen homotopy hh on XX, that is ip=1+Xh+hXip=1+\partial^{X}h+h\partial^{X}. Note that sometimes a homotopy equivalence is defined to include the condition pi1pi\simeq 1, but the version of the Perturbation Lemma in the sources we consulted do not include this condition.

It is called a deformation retract if, in addition, one has pi=1pi=1. This condition holds in our applications.

Next we see that deformation retracts that satisfy hi=0hi=0 give a simple way to transfer dg algebra structures from a complex to its summand as long as the associated homotopy satisfies the following additional property, a weakening of the Leibniz rule, which we introduce below.

1.3 Definition.

Let XX be a complex of RR-modules equipped with a product, and let h:XXh\colon X\longrightarrow X be a graded map (but not necessarily a chain map). We say that hh satisfies the generalized Leibniz rule if one has

h(αβ)h(α)X+Xh(β)h(\alpha\beta)\subseteq h(\alpha)X+Xh(\beta)

for every α\alpha and β\beta in XX.

For our application, the map hh will in fact satisfy a stronger condition, which we call scaled Leibniz rule, namely that for every α,βX\alpha,\beta\in X there are r,sRr,s\in R depending only on the degrees of α\alpha and β\beta, respectively, such that

h(αβ)=rh(α)β+sαh(β)h(\alpha\beta)=rh(\alpha)\beta+s\alpha h(\beta)

We now prove the main result of this section.

1.4 Proposition.

Let XX be a dg algebra. Consider a deformation retract

(X,X)p[]i----(Y,Y)(X,\partial^{X})\stackrel{{\scriptstyle[}}{{p}}]{i}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}(Y,\partial^{Y})

with associated homotopy hh that satisfies the generalized Leibniz rule and hi=0hi=0. The following product defines a dg algebra structure on YY

αβ= def p(i(α)i(β)) for α,βY\alpha\beta\stackrel{{\scriptstyle\text{ def }}}{{=}}p\left(i(\alpha)i(\beta)\right){\text{ for }}~{}\alpha,\beta\in Y

where the product inside parentheses is the one in XX.

Moreover, with this structure on YY, the map ii becomes a dg algebra homomorphism.

Proof.

The Leibniz rule for YY holds without the assumptions that hh satisfies the generalized Leibniz rule and hi=0hi=0. Indeed, for any elements α,βY\alpha,\beta\in Y, one has

Y(αβ)\displaystyle\partial^{Y}(\alpha\beta) =(Yp)(i(α)i(β))\displaystyle=(\partial^{Y}p)(i(\alpha)i(\beta))
=(pX)(i(α)i(β))\displaystyle=(p\partial^{X})(i(\alpha)i(\beta))
=p(X(i(α))i(β)+(1)|α|i(α)X(i(β)))\displaystyle=p\left(\partial^{X}(i(\alpha))i(\beta)+(-1)^{|\alpha|}i(\alpha)\partial^{X}(i(\beta))\right)
=p(i(Y(α))i(β))+(1)|α|p(i(α)i(Y(β)))\displaystyle=p\left(i(\partial^{Y}(\alpha))i(\beta)\right)+(-1)^{|\alpha|}p\left(i(\alpha)i(\partial^{Y}(\beta))\right)
=Y(α)β+(1)|α|αY(β)\displaystyle=\partial^{Y}(\alpha)\beta+(-1)^{|\alpha|}\alpha\partial^{Y}(\beta)

where the first equality is from the definition of the product, the second one holds since pp is a chain map, the third one is from the Leibniz rule for XX, the fourth holds since ii is a chain map, the fifth is again from the definition of the product.

To prove the associativity and the last assertion, we first show that

(1.4.1) (Xh+hX)(i(α)i(β))=0(\partial^{X}h+h\partial^{X})\left(i(\alpha)i(\beta)\right)=0

for all α,βY\alpha,\beta\in Y. If one expands this expression using that X\partial^{X} satisfies the Leibniz rule and hh satisfies the generalized Leibniz rule, one sees that every term has a factor with hihi, which is zero, or a factor with hXih\partial^{X}i which is also zero since Xi=iY\partial^{X}i=i\partial^{Y}.

To verify associativity, take any elements α,β,γY\alpha,\beta,\gamma\in Y. One has

(αβ)γ\displaystyle(\alpha\beta)\gamma =p(i(α)i(β))γ\displaystyle=p\left(i(\alpha)i(\beta)\right)\gamma
=p((ip(i(α)i(β))i(γ))\displaystyle=p\left((ip(i(\alpha)i(\beta))i(\gamma)\right)
=p((1+Xh+hX)(i(α)i(β))i(γ))\displaystyle=p\left((1+\partial^{X}h+h\partial^{X})\left(i(\alpha)i(\beta)\right)i(\gamma)\right)
=p((i(α)i(β))i(γ))\displaystyle=p\left(\left(i(\alpha)i(\beta)\right)i(\gamma)\right)

where the first two equalities are from the definition of the product and the third one is by the equality ip=1+Xh+hXip=1+\partial^{X}h+h\partial^{X}. A similar argument shows that

α(βγ)=p(i(α)(i(β)i(γ)))\displaystyle\alpha(\beta\gamma)=p\left(i(\alpha)\left(i(\beta)i(\gamma)\right)\right)

and hence associativity holds since it holds for XX.

Finally, one can see that p(1)p(1) is the identity element of YY and that the product on YY is graded commutative since pp and ii are graded maps.

To see that the map ii is a dg algebra homomorphism, let α,βY\alpha,\beta\in Y. One then has

i(αβ)\displaystyle i(\alpha\beta) =ip(i(α)i(β))\displaystyle=ip\left(i(\alpha)i(\beta)\right)
=(1+Xh+hX)(i(α)i(β))\displaystyle=(1+\partial^{X}h+h\partial^{X})\left(i(\alpha)i(\beta)\right)
=i(α)i(β)\displaystyle=i(\alpha)i(\beta)

where the second equality holds as ii and pp form a deformation retract and the last one follows from (1.4.1). ∎

Note that the condition hi=0hi=0 in Proposition 1.4 holds when the deformation retract is special; see 3.3.1 for definition.

For the application we have in mind in Section 4, we need a slightly stronger result since, after we apply the Perturbation Lemma, the new homotopy need no longer satisfy the generalized Leibniz rule even if the original one does; however we show in this case that the descent still works as long as the original deformation retract is special. We give this result in Proposition 3.5.

2. Connections to Homotopy Transfer Theorems

In Proposition 1.4, we found that a dg algebra structure descends along certain deformation retracts as long as the homotopy satisfies the generalized Leibniz rule, defined in 1.3. In this section, we compare this to the result of the well-known Homotopy Transfer Theorem, via which the dg algebra structure descends to an AA_{\infty}-structure, which can have nontrivial higher products even when the descended structure is associative and hence a dg algebra. Under the aforementioned additional hypothesis on the homotopy, we compute the higher operations that arise from the descent and find them to vanish after all in Proposition 2.5. In Proposition 2.7, we also discuss what happens when the original complex is merely an AA_{\infty}-algebra under a similar, but much stronger hypothesis on the homotopy.

We note that most sources for the Homotopy Transfer Theorems work with dg algebras and AA_{\infty}-algebras over a field of characteristic zero. However, these are known to hold over a commutative ring RR as long as one makes some freeness assumptions. For simplicity we assume in this section that RR is a field of characteristic 0 (or, more generally, that we are in a setting in which the Homotopy Transfer Theorems are known to hold). However, we should point out that our transfer results Proposition 1.4 and  3.5 do not require any such hypotheses.

We begin by recalling both the definition of an AA_{\infty}-algebra and the Homotopy Transfer Theorem for a dg algebra. The concept of an AA_{\infty}-algebra was introduced by Stasheff in [Sta63] in his study of loop spaces, where the natural product is only associative up to homotopy. For some expositions of this topic, see [Kel01], [Kel06], and [LH03].

2.1 Definition.

An AA_{\infty}-algebra over a ring RR is a complex AA of RR-modules together with RR-multilinear maps of degree n2n-2

mn:AnAm_{n}\colon A^{\otimes n}\to A

for each n1n\geq 1, called operations or multiplications, satisfying the following relations, called the Stasheff identities.

  • The first operation is simply the differential:

    m1=A.m_{1}=\partial_{A}.
  • The second operation satisfies the Leibniz rule:

    m1m2=m2(m11+1m1).m_{1}m_{2}=m_{2}(m_{1}\otimes 1+1\otimes m_{1}).
  • The third one verifies that m2m_{2} is associative up to the homotopy m3m_{3}:

    m2(\displaystyle m_{2}( 1m2m21)\displaystyle 1\otimes m_{2}-m_{2}\otimes 1)
    =m1m3+m3(m111+1m11+11m1)\displaystyle=m_{1}m_{3}+m_{3}(m_{1}\otimes 1\otimes 1+1\otimes m_{1}\otimes 1+1\otimes 1\otimes m_{1})

    Note that the left hand side is the obstruction to associativity for m2m_{2} and that the right hand side is the boundary of m3m_{3} in HomR(A3,A)\operatorname{Hom}_{R}(A^{\otimes 3},A).

  • More generally, for n1n\geq 1, we have

    s=1nr,t0(1)r+stmr+1+t(1rms1t)=0\sum_{s=1}^{n}\sum_{r,t\geqslant 0}(-1)^{r+st}m_{r+1+t}(1^{\otimes r}\otimes m_{s}\otimes 1^{\otimes t})=0

    where the sums are taken over the values of r,s,tr,s,t with r+s+t=nr+s+t=n.

Note that when one applies the maps in each formula above to an element, one should use the Koszul sign rule: For graded maps ff and gg, one has

(fg)(xy)=(1)|g||x|f(x)g(y)(f\otimes g)(x\otimes y)=(-1)^{|g||x|}f(x)\otimes g(y)

for homogeneous elements xx and yy, where |w||w| denotes the degree of ww whether it is a map or homogenous element.

Recall that we are using homological notation; in cohomological notation the degree of mnm_{n} would be 2n2-n rather than n2n-2. Note also that for the signs we follow the conventions in Getzler-Jones [GJ90]; see, for example, the survey by Keller [Kel01].

2.2 Remark.

Note that an AA_{\infty}-algebra whose operations mnm_{n} are zero for all n3n\geqslant 3 is a dg algebra where the product is given by m2m_{2}. Conversely, a dg algebra can be given the structure of an AA_{\infty}-algebra by setting m3=0m_{\geq 3}=0.

However, m1m_{1} and m2m_{2} can usually be extended to other AA_{\infty}-algebra structures. Indeed, one can have nonzero higher operations for which the boundary of m3m_{3} in HomR(A3,A)\operatorname{Hom}_{R}(A^{\otimes 3},A) is equal to zero and hence AA is still associative.

The Homotopy Transfer Theorems were first proved by Kadeishvili in [Kad80] and [Kad82]. We recall them in (2.4) and (2.6). For this, we follow the exposition in Vallette’s survey [Val14]. We note that our signs are the opposite of those in his survey since his homotopy is the negative of ours (he has 1ip=Xh+hX1-ip=\partial^{X}h+h\partial^{X}, rather than ip1ip-1). This should not make a difference as the precise signs do not matter for our proofs.

2.3 Notation.

We introduce the planar rooted tree notation from Vallette’s survey to represent these products pictorially as this will make it easier to describe the Homotopy Transfer Theorem. All the diagrams are read from the top down, that is, the inputs are thought of as being entered on the top and the multi-intersections correspond to the higher products mnm_{n} being performed. Further, wherever a letter appears in such a diagram, one applies the corresponding map at that point. Again, the sign rule described in Definition 2.1 is understood to be in effect.

First, the higher operation mnm_{n} is drawn as follows.

12\cdotsnn

In this notation, the properties of Leibniz rule and associativity can be drawn as follows.

=\partial\partial+\partial=

To justify the first equation, note that 1\partial\otimes 1 will produce no sign when applied to the input xyx\otimes y since |1|=0|1|=0, but 11\otimes\partial will have the sign (1)|x|(-1)^{|x|}.

We now recall the Homotopy Transfer Theorem [Kad80] that allows one to transfer a dg algebra structure along a deformation retract yielding an AA_{\infty}-structure on the retract. For this, we will define the descended higher operations using the tree notation introduced above.

2.4

Homotopy Transfer Theorem for dg algebras. Let

(X,X)p[]i----(Y,Y)(X,\partial^{X})\stackrel{{\scriptstyle[}}{{p}}]{i}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}(Y,\partial^{Y})

be a deformation retract with associated homotopy hh where XX is a dg algebra. As in Remark 2.2, one considers XX an AA_{\infty}-algebra with m1m_{1} equal to the differential, m2m_{2} equal to the dg algebra product on XX, and mnX=0m_{n}^{X}=0 for n3n\geq 3.

The Homotopy Transfer Theorem gives an AA_{\infty}-structure on YY as follows: First set m1Y=Ym_{1}^{Y}=\partial^{Y}. For n2n\geq 2, the nnth operation mnYm_{n}^{Y} is defined as

12\cdotsnn\coloneqqPBTn{{{{\sum_{PBT_{n}}}}}}±\pmiiiihhhhiihhiiiipp

where the left hand side is the notation for mnYm_{n}^{Y} and where the sum is over PBTnPBT_{n}, the set of all planar binary rooted trees with nn leaves, and the tree diagram pictured on the right is just a representative example of such a tree. The pattern of maps appearing on each tree is meant to indicate that every product is followed by an application of hh, except for the last one, where instead pp is applied. The actual signs, indicated simply as ±\pm above, are defined in the various sources quoted, but we shall not need them for our results. Again, in applying the maps in trees, the sign rule described in Definition 2.1 is understood to be in effect.

In particular, m2Ym_{2}^{Y} and m3Ym_{3}^{Y} are given by

=iiiippand==hhiiiiiipp-iiiihhiipp

where the signs in the expression for m3Ym_{3}^{Y} are the opposite of those in [Val14] since, as we recall, his homotopy is the negative of ours.

Under the hypotheses in this paper, we can show that the descent actually yields an AA_{\infty}-algebra with all higher operations m3m_{\geq 3} equal to zero. Proposition 1.4 does yield a dg algebra, and so one could extend it to an AA_{\infty}-algebra by defining the higher operations m3m_{\geq 3} equal to zero, but the Homotopy Transfer Theorem (2.4) also gives a set of higher operations, which may not be the same. Here we prove that those vanish as well.

2.5 Proposition.

Let XX be a dg algebra. Consider a deformation retract

(X,X)p[]i----(Y,Y)(X,\partial^{X})\stackrel{{\scriptstyle[}}{{p}}]{i}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}(Y,\partial^{Y})

with associated homotopy hh that satisfies the generalized Leibniz rule and hi=0hi=0. Then the AA_{\infty}-algebra structure on YY obtained from the dg algebra structure on XX via 2.4 has trivial higher operations, that is, mnY=0m_{n}^{Y}=0 for all n3n\geq 3.

Proof.

Recall from 2.4 that the operations mnYm_{n}^{Y} for n3n\geq 3 descended from the dg algebra structure on XX are signed sums of elements described by planar binary rooted trees with nn leaves. The signs do not matter as we prove that every term equals zero. Indeed, each term always includes a factor of the form h(i(α)i(β))h(i(\alpha)i(\beta)) for some α,βY\alpha,\beta\in Y (this will be nested inside other maps ii, hh and pp and products from XX). This vanishes as hh satisfies the generalized Leibniz rule and hi=0hi=0 holds. Hence these higher operations mnm_{n} all vanish. ∎

What if one begins with a complex XX that is an AA_{\infty}-algebra rather than a dg algebra? First we recall the version of the Homotopy Transfer Theorem for this situation from [Kad82], namely a more general version that allows one to transfer an AA_{\infty}-structure along a deformation retract yielding an AA_{\infty}-structure on the retract. We again use the tree notation introduced above.

2.6

Homotopy Transfer Theorem for AA_{\infty}-algebras. Let

(X,X)p[]i----(Y,Y)(X,\partial^{X})\stackrel{{\scriptstyle[}}{{p}}]{i}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}(Y,\partial^{Y})

be a deformation retract with associated homotopy hh where XX is an AA_{\infty}-algebra. The Homotopy Transfer Theorem gives an AA_{\infty}-structure on YY as follows: First set m1Y=Ym_{1}^{Y}=\partial^{Y}. For n2n\geq 2, the operation mnYm_{n}^{Y} is defined as

12\cdotsnn\coloneqqPTn{{{{\sum_{PT_{n}}}}}}±\pmiiiiiihhhhiihhiiiiiiiiiipp

where the left hand side is the notation for mnYm_{n}^{Y} and where the sum is over PTnPT_{n}, the set of all planar (not necessarily binary) rooted trees with nn leaves, and the tree diagram pictured on the right is just a representative example of such a tree, where the higher products are all occurring in XX. Once again, the pattern is that every such product is followed by an application of hh, except for the last one, where instead pp is applied. Again, the actual signs are defined in the various sources quoted, but we shall not need them for our results.

Next we impose analogous but much stronger conditions on the homotopy when XX is merely an AA_{\infty}-algebra, rather than a dg algebra. We do not have any example that satisfies this condition, but include this result for completeness in case it could be useful. We say that the homotopy hh on XX satisfies the generalized Leibniz rule for an AA_{\infty}-algebra if for every n2n\geq 2 one has

h(mn(a1an))i=1nmn(Xh(ai)X)h(m_{n}(a_{1}\otimes\dots\otimes a_{n}))\subseteq\sum_{i=1}^{n}m_{n}(X\otimes\dots\otimes h(a_{i})\otimes\dots\otimes X)

where h(ai)h(a_{i}) is the iith factor and the other factors are XX. Under this hypothesis, we can show that the formulas for the descended operations via the Homotopy Transfer Theorem for AA_{\infty}-algebras (see 2.6) are much simpler than usual (they are just the ones induced by going back and forth along the homotopy equivalence).

2.7 Proposition.

Let XX be an AA_{\infty}-algebra with operations mnXm^{X}_{n} for n1n\geq 1. Consider a deformation retract

(X,X)p[]i----(Y,Y)(X,\partial^{X})\stackrel{{\scriptstyle[}}{{p}}]{i}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}(Y,\partial^{Y})

with associated homotopy hh that satisfies the generalized Leibniz rule for an AA_{\infty}-algebra and hi=0hi=0. Then the AA_{\infty}-algebra structure on YY obtained from the AA_{\infty}-algebra structure on XX via 2.6 has operations given by

mnY=pmnX(ii)m_{n}^{Y}=p\,m_{n}^{X}(i\otimes\dots\otimes i)

for all n1n\geq 1.

Proof.

Note that in 2.6, it follows from the construction and the properties of a deformation retract that

m1Y=Y=pm1Xi and m2Y=pm2X(ii).m_{1}^{Y}=\partial^{Y}=p\,m_{1}^{X}\,i~{}{\text{ \ and \ }}~{}m_{2}^{Y}=p\,m_{2}^{X}(i\otimes i).

This covers the cases n=1,2n=1,2.

Recall from 2.6 that the operations mnYm_{n}^{Y} for n3n\geq 3 descended from the dg algebra structure on XX are signed sums of elements described by planar rooted trees with nn leaves. The signs do not matter as all of the terms vanish except one. Indeed, expanding using that hi=0hi=0 and the generalized Leibniz rule for an AA_{\infty}-algebra leaves only the desired term as that is the only one given by a tree with only one (higher) operation, hence simply followed by an application of pp and not involving the homotopy hh; this term is known to be positive. ∎

3. Transfer of dg algebra structures and the Perturbation Lemma

The second main aim of this paper is to use descent along a deformation retract to find a dg algebra structure on a well known complex, which we do in the next section. Building this retract involves a homological tool called the Perturbation Lemma. In this section we extend the descent result in Proposition 1.4 to perturbations of the original setting, resulting in Proposition 3.5. One can similarly extend Propositions 2.5 and 2.7; see Remark 3.6.

The Perturbation Lemma generates new homotopy equivalences from initial ones; in general the aim is to modify the differentials of the complexes while maintaining a homotopy equivalence. For more details, the reader may consult Crainic’s exposition in [Cra04] and also [DM13] where Dyckerhoff and Murfet develop the lemma for the analogous case of matrix factorizations. The Perturbation Lemma is especially useful for double complexes where one can temporarily forget either the horizontal or the vertical differentials and add them back in later as the “perturbation"; this is the context in which we will apply it in Section 4.

We define some terminology we use in stating the Perturbation Lemma.

3.1 Definition.

Consider a set of homotopy equivalence data

(X,X)p[]i----(Y,Y)(X,\partial^{X})\stackrel{{\scriptstyle[}}{{p}}]{i}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}(Y,\partial^{Y})

with associated homotopy hh. A perturbation is a map δ\delta on XX of the same degree as the differential X\partial^{X} such that (X+δ)2=0(\partial^{X}+\delta)^{2}=0, that is, X+δ\partial^{X}+\delta is again a differential. The perturbation δ\delta is called small if 1δh1-\delta h is invertible. Most commonly, this happens when δh\delta h is elementwise nilpotent for then one has

(1δh)1=j=0(δh)j=1+(δh)+(δh)2+(1-\delta h)^{-1}=\sum_{j=0}^{\infty}(\delta h)^{j}=1+(\delta h)+(\delta h)^{2}+\cdots

where the sum is finite on each element of XX.

3.2 Definition.

Consider a set of homotopy equivalence data

(X,X)p[]i----(Y,Y)(X,\partial^{X})\stackrel{{\scriptstyle[}}{{p}}]{i}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}(Y,\partial^{Y})

with associated homotopy hh. Let δ\delta be a small perturbation on XX, and let A=(1δh)1δA=(1-\delta h)^{-1}\delta. We define the following new data

(X,X)p[]i----(Y,Y)(X,\partial_{\infty}^{X})\stackrel{{\scriptstyle[}}{{p}}_{\infty}]{i_{\infty}}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}(Y,\partial_{\infty}^{Y})

where

i=i+hAi,p=p+pAh,X=X+δ, andY=Y+pAii_{\infty}=i+hAi,~{}p_{\infty}=p+pAh,~{}\partial_{\infty}^{X}=\partial^{X}+\delta,\text{ and}~{}\partial_{\infty}^{Y}=\partial^{Y}+pAi

and set

h=h+hAhh_{\infty}=h+hAh

Note that when δh\delta h is elementwise nilpotent, then the formulas can be rewritten as follows.

i\displaystyle i_{\infty} =(1+(hδ)+(hδ)2+)i\displaystyle=(1+(h\delta)+(h\delta)^{2}+\cdots)i
p\displaystyle p_{\infty} =p(1+(δh)+(δh)2+)\displaystyle=p(1+(\delta h)+(\delta h)^{2}+\cdots)
h\displaystyle h_{\infty} =h(1+(δh)+(δh)2+)\displaystyle=h(1+(\delta h)+(\delta h)^{2}+\cdots)
Y\displaystyle\partial^{Y}_{\infty} =Y+pδi\displaystyle=\partial^{Y}+p\delta i_{\infty}
=Y+pδi\displaystyle=\partial^{Y}+p_{\infty}\delta i
3.3 Definition.

A special deformation retract is a deformation retract that satisfies the following equations

(3.3.1) hi=0,ph=0,h2=0.hi=0,~{}ph=0,~{}h^{2}=0.

These ensure that the property pi=1pi=1 is inherited by the perturbed data.

As described in [Cra04], any deformation retract can be converted into a special one by modifying the chosen homotopy in several steps; the drawback is that when seeking explicit formulas the resulting maps become more complicated. Fortunately, for our application, the deformation retracts involved are all special.

With this terminology, we are now ready to state the Perturbation Lemma.

3.4

Perturbation Lemma. Given a set of homotopy equivalence data

(X,X)p[]i----(Y,Y)(X,\partial^{X})\stackrel{{\scriptstyle[}}{{p}}]{i}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}(Y,\partial^{Y})

with associated homotopy hh, its perturbation via a small perturbation gives a set of homotopy equivalence data

(X,X)p[]i----(Y,Y)(X,\partial_{\infty}^{X})\stackrel{{\scriptstyle[}}{{p}}_{\infty}]{i_{\infty}}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}(Y,\partial_{\infty}^{Y})

with associated homotopy hh_{\infty}.

If, furthermore, the original homotopy equivalence is a special deformation retract then so is the resulting one, that is,

pi=1,h2=0,hi=0,andph=0,p_{\infty}i_{\infty}=1,\ h_{\infty}^{2}=0,\ ~{}h_{\infty}i_{\infty}=0,\ ~{}{\text{and}}~{}p_{\infty}h_{\infty}=0,

Recall that by Proposition 1.4, given a special deformation retract whose homotopy satisfies the generalized Leibniz rule, a dg algebra structure can be transferred along it. One might want to use the Perturbation Lemma to obtain new deformation retracts to which one could apply this proposition. However, even if the original homotopy satisfies the generalized Leibniz rule, the new one may no longer satisfy it. We remedy this by proving an extension of the descent results as follows.

3.5 Proposition.

Let XX be a dg algebra. Consider a special deformation retract

(X,X)p[]i----(Y,Y)(X,\partial^{X})\stackrel{{\scriptstyle[}}{{p}}]{i}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}(Y,\partial^{Y})

with associated homotopy hh that satisfies the generalized Leibniz rule, and let δ\delta be a small perturbation on XX.

If the perturbed complex (X,X)(X,\partial^{X}_{\infty}) from Lemma 3.4 remains a dg algebra via the same product (equivalently, if δ\delta satisfies the Leibniz rule), then the following product defines a dg algebra structure on the perturbed complex (Y,Y)(Y,\partial^{Y}_{\infty})

αβ= def p(i(α)i(β)) for α,βY\alpha\beta\stackrel{{\scriptstyle\text{ def }}}{{=}}p_{\infty}\left(i_{\infty}(\alpha)i_{\infty}(\beta)\right){\text{ for }}~{}\alpha,\beta\in Y

where the product inside parentheses is the one in XX.

Moreover, with this structure on YY, the map ii_{\infty} becomes a dg algebra homomorphism.

Proof.

The proof is the same as that of Proposition 1.4 with the exception that to prove associativity and that ii_{\infty} is a dg algebra homomorphism, one needs to show that

(Xh+hX)(i(α)i(β))=0.(\partial^{X}_{\infty}h_{\infty}+h_{\infty}\partial^{X}_{\infty})\left(i_{\infty}(\alpha)i_{\infty}(\beta)\right)=0.

Here this follows by similar reasoning due to the facts that one has

h=h+hAh and i=i+hAi where A=(1δh)1δh_{\infty}=h+hAh~{}{\text{ and }}~{}i_{\infty}=i+hAi~{}{\text{ where }}~{}A=(1-\delta h)^{-1}\delta

and that h2=0h^{2}=0 and hi=0hi=0. ∎

3.6 Remark.

We note that the analogous generalizations of Propositions 2.5 and 2.7 hold as well, with similar proofs modified as the one above.

4. Application to a minimal resolution

Let R=k[x1,,xn]R=k[x_{1},\cdots,x_{n}] be a polynomial ring over a field kk. In [BE75], Buchsbaum and Eisenbud introduced the minimal free resolution 𝕃a\mathbb{L}_{a} of the quotient R/(x1,,xn)aR/(x_{1},\cdots,x_{n})^{a} of RR by powers of the homogeneous maximal ideal. In [Sri89], Srinivasan gives a dg algebra structure on 𝕃a\mathbb{L}_{a} using Young tableaux. In this section, we use the Perturbation Lemma in a simple way to obtain a dg algebra structure on 𝕃a\mathbb{L}_{a} that is SnS_{n}-invariant. Our approach works in characteristic zero and in positive characteristic provided that the characteristic is large enough.

We begin by recalling the definition of the resolution 𝕃a\mathbb{L}_{a} and relating it to (the totalization of) a truncation of a certain double complex in (4.1), (4.3), and (4.4). In (4.5) and (4.6) we use the Perturbation Lemma to form a deformation retract between them, as long as one has an appropriately nice associated homotopy so that one can apply Proposition 3.5. Lastly we define such a homotopy using a scaled de Rham differential in (4.7), proving its properties in Lemmas 4.8 and  4.9, culminating in Theorem 4.10.

4.1

Here we define the double complex we will be working with. This is simply a rearrangement of a Koszul complex as a double complex of free RR-modules; see Remark 4.2.

Let S=R[y1,,yn]S=R[y_{1},\dots,y_{n}] be a polynomial ring and let Λ=Re1,,en\Lambda=R\langle e_{1},\dots,e_{n}\rangle be an exterior algebra. Consider the following (anticommutative) double complex whose rows are the strands of the Koszul complex K(y1,,yn;S)K(y_{1},\dots,y_{n};S) and whose columns are the tensor product over RR of the Koszul complex over RR on x1,,xnx_{1},\dots,x_{n} with the graded pieces SaS_{a} of SS. We denote both it and its totalization by 𝕊\mathbb{S}, as it is clear everywhere from the context which we mean. All the tensor products in the diagram are over RR.

(4.1.1) ΛnSa\textstyle{\Lambda^{n}\otimes S_{a}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}κ\scriptstyle{\ \ \ \kappa}\textstyle{\cdots}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ\scriptstyle{\!\!\!\!\!\!\!\!\!\!\!\!\!\kappa}Λn1Sa\textstyle{\Lambda^{n-1}\otimes S_{a}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ\scriptstyle{\ \ \ \ \kappa}\textstyle{\cdots}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}ΛaS2\textstyle{\Lambda^{a}\otimes S_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}κ\scriptstyle{\ \ \ \kappa}\textstyle{\cdots}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ\scriptstyle{\!\!\!\!\!\!\!\kappa}Λ2Sa\textstyle{\Lambda^{2}\otimes S_{a}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}κ\scriptstyle{\ \ \ \kappa}\textstyle{\cdots}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}ΛaS1\textstyle{\Lambda^{a}\otimes S_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}κ\scriptstyle{\!\!\!\!\kappa}Λa1S2\textstyle{\Lambda^{a-1}\otimes S_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}κ\scriptstyle{\ \ \ \ \ \kappa}\textstyle{\cdots}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ\scriptstyle{\!\!\!\!\!\!\!\kappa}Λ1Sa\textstyle{\Lambda^{1}\otimes S_{a}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}κ\scriptstyle{\ \ \ \kappa}\textstyle{\cdots}ΛaS0\textstyle{\Lambda^{a}\otimes S_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ\scriptstyle{\!\!\!\!\kappa}Λa1S1\textstyle{\Lambda^{a-1}\otimes S_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ\scriptstyle{\kappa}Λa2S2\textstyle{\Lambda^{a-2}\otimes S_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ\scriptstyle{\ \ \ \ \ \kappa}\textstyle{\cdots}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ\scriptstyle{\!\!\!\!\!\!\!\!\kappa}Λ0Sa\textstyle{\Lambda^{0}\otimes S_{a}}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}Λ3S0\textstyle{\Lambda^{3}\otimes S_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ\scriptstyle{\kappa}d\scriptstyle{d}Λ2S1\textstyle{\Lambda^{2}\otimes S_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ\scriptstyle{\kappa}d\scriptstyle{d}Λ1S2\textstyle{\Lambda^{1}\otimes S_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}d\scriptstyle{d}κ\scriptstyle{\ \ \ \ \kappa}\textstyle{\cdots}Λ2S0\textstyle{\Lambda^{2}\otimes S_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ\scriptstyle{\kappa}d\scriptstyle{d}Λ1S1\textstyle{\Lambda^{1}\otimes S_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ\scriptstyle{\kappa}d\scriptstyle{d}Λ0S2\textstyle{\Lambda^{0}\otimes S_{2}}Λ1S0\textstyle{\Lambda^{1}\otimes S_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κ\scriptstyle{\kappa}d\scriptstyle{d}Λ0S1\textstyle{\Lambda^{0}\otimes S_{1}}Λ0S0\textstyle{\Lambda^{0}\otimes S_{0}}

More explicitly, the horizontal differentials κi,a:ΛiSaΛi1Sa+1\kappa_{i,a}:\Lambda^{i}\otimes S_{a}\to\Lambda^{i-1}\otimes S_{a+1} are given by

(4.1.2) κi,a(ei1eiip)=j=1a(1)jei1e^ijeiiyijp\kappa_{i,a}(e_{i_{1}}\wedge\cdots\wedge e_{i_{i}}\otimes p)=\sum_{j=1}^{a}(-1)^{j}e_{i_{1}}\wedge\cdots\wedge\widehat{e}_{i_{j}}\wedge\dots\wedge e_{i_{i}}\otimes y_{i_{j}}p

and the vertical differentials di,a:ΛiSaΛi1Sad_{i,a}\colon\Lambda^{i}\otimes S_{a}\to\Lambda^{i-1}\otimes S_{a} are given by

di,a=kosi1d_{i,a}=\textrm{kos}_{i}\otimes 1

where kosi\textrm{kos}_{i} denotes the iith differential in the Koszul complex K(x1,,xn;R)K(x_{1},\dots,x_{n};R).

For the totalization of this double complex (or of truncations of it), since it is anticommutative, the differentials are defined as

i=a(κi,a+di,a)\partial_{i}=\sum_{a}(\kappa_{i,a}+d_{i,a})

without adding any signs. For simplicity we write

=κ+d.\partial=\kappa+d.

We continue to omit the indices on the maps when there is no ambiguity.

4.2 Remark.

As an aside, we give a slightly different way of obtaining a double complex which could have been used in this section. It differs only in signs from the one pictured in (4.1.1), but comes from a well known construction.

Let VV be a kk-vector space with dimkV=n\dim_{k}V=n, and consider the symmetric and exterior algebras

S¯\displaystyle\overline{S} =S(V)k[x1,,xn]k[x1′′,,xn′′]\displaystyle=S(V)\cong k[x_{1}^{\prime},\dots,x_{n}^{\prime}]\cong k[x_{1}^{\prime\prime},\dots,x_{n}^{\prime\prime}]
Λ¯\displaystyle\overline{\Lambda} =Λ(V)ke1,,en\displaystyle=\Lambda(V)\cong k\langle e_{1},\dots,e_{n}\rangle

Consider S¯R\overline{S}\cong R as a module over its enveloping algebra S¯e=S¯kS¯\overline{S}^{\,e}=\overline{S}\otimes_{k}\overline{S} via the multiplication map. Its minimal graded free resolution, after identifying the two copies of S¯\overline{S} with polynomial rings as in the display above, is the Koszul complex Λ¯kS¯kS¯\overline{\Lambda}\otimes_{k}\overline{S}\otimes_{k}\overline{S} on the regular sequence {xi11xi′′}\{x_{i}^{\prime}\otimes 1-1\otimes x_{i}^{\prime\prime}\}. Rearranging factors, it can be expressed as

S¯kΛ¯kS¯=S¯kΛ¯kS¯=S¯kΛ¯kS¯′′\overline{S}\otimes_{k}\overline{\Lambda}\otimes_{k}\overline{S}=\underbrace{\overline{S}\otimes_{k}\overline{\Lambda}}_{\partial^{\prime}}\otimes_{k}\overline{S}=\overline{S}\otimes_{k}\underbrace{\overline{\Lambda}\otimes_{k}\overline{S}}_{\partial^{\prime\prime}}

with the homological degree being the degree of the middle factor and

=11′′=dκ\partial=\partial^{\prime}\otimes 1-1\otimes\partial^{\prime\prime}=d-\kappa

where \partial^{\prime} is the Koszul differential on x1,,xnx_{1}^{\prime},\dots,x_{n}^{\prime} and ′′\partial^{\prime\prime} is the Koszul differential on x1′′,,xn′′x_{1}^{\prime\prime},\dots,x_{n}^{\prime\prime}. Viewing graded strands, one can write this as a totalization of an anticommutative double complex of free RR-modules given by

S¯kΛ¯ikS¯jRkΛ¯ikS¯j(RkΛ¯i)R(RkS¯j)ΛiRSj\overline{S}\otimes_{k}\overline{\Lambda}^{i}\otimes_{k}\overline{S}_{j}\cong R\otimes_{k}\overline{\Lambda}^{i}\otimes_{k}\overline{S}_{j}\cong(R\otimes_{k}\overline{\Lambda}^{i})\otimes_{R}(R\otimes_{k}\overline{S}_{j})\cong\Lambda^{i}\otimes_{R}S_{j}

Although this double complex differs from the one pictured in (4.1.1) by a sign on the horizontal maps κ\kappa, one could equally well use this complex in the rest of this section; similarly, one could obtain the double complex (4.1.1) from a Koszul complex by using xi′′-x_{i}^{\prime\prime} in place of xi′′x_{i}^{\prime\prime} above; note that it would no longer be a resolution of RR over its enveloping algebra.

4.3

We introduce the complexes 𝕃a\mathbb{L}_{a} of Buchsbaum and Eisenbud here. They show that this is a minimal RR-free resolution of R/𝔪aR/\mathfrak{m}^{a}, where 𝔪\mathfrak{m} is the homogeneous maximal ideal of RR.

It is well known that the rows of the double complex (4.1.1) except the bottom one are exact; in fact, they can be viewed as the result of applying a base change to the strands of the tautological Koszul complex (see, for example, [MR18]). Hence they are contractible as they consist of free RR-modules. So one can define free RR-modules

Li,a=imκi+1,a1=kerκi,acokerκi+2,a2,L_{i,a}=\operatorname{im}\kappa_{i+1,a-1}=\ker\kappa_{i,a}\cong\nobreak{\operatorname{coker\,}}\kappa_{i+2,a-2},

in other words, with split exact sequences

Λi+2Sa2κi+2,a2Λi+1\displaystyle\Lambda^{i+2}\otimes S_{a-2}\xrightarrow[]{\;\kappa_{i+2,a-2}\;}\Lambda^{i+1} Sa1κi+1,a1Li,a0\displaystyle\otimes S_{a-1}\xrightarrow[]{\;\kappa_{i+1,a-1}\;}L_{i,a}\longrightarrow 0
0Li,aΛi\displaystyle 0\longrightarrow L_{i,a}\xrightarrow[]{\;\subseteq\;}\Lambda^{i} Saκi,aΛi1Sa+1.\displaystyle\otimes S_{a}\xrightarrow[]{\;\kappa_{i,a}\;}\Lambda^{i-1}\otimes S_{a+1}.

The vertical differentials dd in the diagram induce maps on these modules, which we again denote by dd, to yield a complex

𝕃a:0Ln1,adn1Ln2,adn2L0,a𝜀R0\mathbb{L}_{a}\colon 0\to L_{n-1,a}\xrightarrow[]{\;d_{n-1}\;}L_{n-2,a}\xrightarrow[]{\;d_{n-2}\;}\dots\to L_{0,a}\xrightarrow[]{\;\varepsilon\;}R\to 0

augmented by the evaluation map

(4.3.1) ε:L0,a=Λ0SaSaR\varepsilon\colon L_{0,a}=\Lambda^{0}\otimes S_{a}\cong S_{a}\to R

induced by the evaluation map from S=R[y1,,yn]S=R[y_{1},\dots,y_{n}] to R=k[x1,,xn]R=k[x_{1},\dots,x_{n}] sending yiy_{i} to xix_{i}.

4.4

Next we define tra(𝕊)\textrm{tr}_{\geq a}(\mathbb{S}) and tra1(𝕊)\textrm{tr}_{\leq a-1}(\mathbb{S}) to be the totalizations of the truncations at column aa of the anticommutative double complex 𝕊\mathbb{S}

{ΛiSjja}and{ΛiSjja1},\displaystyle\{\Lambda^{i}\otimes S_{j}\mid j\geq a\}~{}\text{and}~{}\{\Lambda^{i}\otimes S_{j}\mid j\leq a-1\},

respectively, with differentials inherited from 𝕊\mathbb{S}. It is well-known that there is a quasi-isomorphism, and hence a homotopy equivalence,

tra1(𝕊)𝕃a\textrm{tr}_{\leq a-1}(\mathbb{S})\simeq\mathbb{L}_{a}

but we will re-derive this via the Perturbation Lemma in order to simultaneously transfer a dg algebra structure from tra1(𝕊)\textrm{tr}_{\leq a-1}(\mathbb{S}) over to 𝕃a\mathbb{L}_{a} (by obtaining a special deformation retract rather than just any homotopy equivalence).

To set up for this, we first argue as in [W0̈4] that the left truncation tra1(𝕊)\textrm{tr}_{\leq a-1}(\mathbb{S}) itself has a natural dg algebra structure. Indeed, the entire complex 𝕊\mathbb{S} is a dg algebra with the obvious multiplication: for αΛiSa\alpha\in\Lambda^{i}\otimes S_{a} and βΛjSb\beta\in\Lambda^{j}\otimes S_{b}, the product is obtained by multiplying the factors in Λ\Lambda and in SS independently. It satisfies the Leibniz rule and other properties of a dg algebra because the differentials κ\kappa and dd do and because homological degree in the totalization of 𝕊\mathbb{S} is, in fact, given by the degree in Λ\Lambda. With this multiplication, the right truncation tra(𝕊)\textrm{tr}_{\geq a}(\mathbb{S}) is clearly a dg ideal and the quotient complex

𝕏a= def tra1(𝕊)𝕊/tra(𝕊)\mathbb{X}_{a}\stackrel{{\scriptstyle\text{ def }}}{{=}}\textrm{tr}_{\leq a-1}(\mathbb{S})\cong\mathbb{S}/\textrm{tr}_{\geq a}(\mathbb{S})

is therefore a dg algebra. Concretely, the resulting product on the left truncation tra1(𝕊)\textrm{tr}_{\leq a-1}(\mathbb{S}) is given by the multiplication on 𝕊\mathbb{S} with the proviso that any terms landing in ΛiSj\Lambda^{i}\otimes S_{j} with jaj\geq a are taken to be zero.

For the next step, we first need a tool for converting a split exact sequence to a deformation retract from any truncation to the image of the differential at the truncation.

4.5

Let (X,X)(X,\partial^{X}) be a contractible complex of RR-modules, i.e., one that is homotopy equivalent to zero via a homotopy ss, (i.e., one that is split exact). Denote its truncation at position cc by

trc(X)=Xnn+1XXc+1c+1XXc0\textrm{tr}_{\geq c}(X)=\cdots\longrightarrow X_{n}\xrightarrow[]{\;\partial_{n+1}^{X}\;}\cdots\longrightarrow X_{c+1}\xrightarrow[]{\;\partial_{c+1}^{X}\;}X_{c}\longrightarrow 0

Let imcX\operatorname{im}\partial^{X}_{c} denote the stalk complex with this module in degree cc and 0 modules elsewhere. The chain maps ii and pp given in degree cc by sc1s_{c-1} and cX\partial_{c}^{X}, respectively, yield a deformation retract

trc(X)p[]i----imXc\displaystyle\textrm{tr}_{\geq c}(X)\stackrel{{\scriptstyle[}}{{p}}]{i}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}\operatorname{im}\partial^{X}_{c}

with associated homotopy h=s|trcXh=-s|_{\textrm{tr}_{\geq c}X}. Indeed one can easily check that pi=1pi=1 and ip1ip\simeq 1 via the homotopy hh. Note that one could also use cokerc+1X\nobreak{\operatorname{coker\,}}\partial^{X}_{c+1} instead of imcX\operatorname{im}\partial^{X}_{c} with appropriate ii and pp. If, furthermore, the original contracting homotopy satisfies s2=0s^{2}=0, then the deformation retract is special: h2=0h^{2}=0 and hi=0hi=0 and one always has ph=0ph=0 due to the fact that p=0p=0 in degrees ncn\neq c.

Next we want to transfer this structure from 𝕏a=tra1(𝕊)\mathbb{X}_{a}=\textrm{tr}_{\leq a-1}(\mathbb{S}), which has a dg algebra structure by (4.4) to the minimal free resolution 𝕃a\mathbb{L}_{a} of R/𝔪aR/\mathfrak{m}^{a}. By Proposition 3.5, it suffices to find a special deformation retract of the form

(𝕏a,𝕏a)][----(𝕃a,d)(\mathbb{X}_{a},\partial^{\mathbb{X}_{a}})\stackrel{{\scriptstyle[}}{{]}}{}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}(\mathbb{L}_{a},d)

that is a perturbation of a special deformation retract whose homotopy satisfies generalized Leibniz. Note that the differential of 𝕏a\mathbb{X}_{a} is exactly κ+d\kappa+d. In 4.6, we discuss how one can find this deformation retract, given a contracting homotopy on the higher rows of 𝕊\mathbb{S}, which we define in 4.7 and whose required properties we establish in Lemmas 4.8 and 4.9.

4.6

Here is overview of how we obtain such a deformation retract using the Perturbation Lemma; see (3.4). First we form a deformation retract between two complexes 𝕏a\mathbb{X}_{a}^{\circ} and 𝕃a\mathbb{L}_{a}^{\circ}, where 𝕏a\mathbb{X}_{a}^{\circ} is obtained from 𝕏a\mathbb{X}_{a} by replacing the vertical differentials dd by 0 and 𝕃a\mathbb{L}_{a}^{\circ} is the complex 𝕃a\mathbb{L}_{a} with differentials set equal to zero. We do this via (4.5) using the homotopy from 4.7. Second we use the Perturbation Lemma to reinsert the original differentials on each, which has the effect of modifying the maps ii and pp.

We start by finding a deformation retract of the form

(𝕏a,κ)p[]i----(𝕃a,0)(\mathbb{X}_{a}^{\circ},\kappa)\stackrel{{\scriptstyle[}}{{p}}]{i}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}(\mathbb{L}_{a}^{\circ},0)

with a homotopy hh. For rows of 𝕏a\mathbb{X}_{a}^{\circ} except the bottom one, we use (4.5) as follows. Recall that the rows of 𝕊\mathbb{S} are split exact with a contracting homotopy that we call σ\sigma (an explicit one is given in Remark 4.7). So each row that gets truncated has a deformation retract onto the image Li,aL_{i,a} of the next horizontal differential; see diagram (4.6.1). Note that some of the lower rows will remain intact and hence are homotopy equivalent to zero; see (4.6.1). On the other hand, the row Λ0S0\Lambda^{0}\otimes S_{0} at the bottom of the diagram is not exact and so needs to be dealt with separately in conjunction with R=(𝕃a)0R=(\mathbb{L}_{a})_{0}. For this we use that there is an isomorphism ε:Λ0S0R\varepsilon\colon\Lambda^{0}\otimes S_{0}\to R defined in (4.3.1).

Putting this all together, one obtains chain maps given by

i\displaystyle i ={σon Li,aε1on R\displaystyle=\begin{cases}\sigma&{\text{on }}L_{i,a}\\ \varepsilon^{-1}&{\text{on }}R\end{cases}
p\displaystyle p ={κon ΛiSa1 for i>0εon Λ0S00else,\displaystyle=\begin{cases}\kappa&{\textrm{on }}\Lambda^{i}\otimes S_{a-1}{\textrm{ for }}i>0\\ \varepsilon&{\textrm{on }}\Lambda^{0}\otimes S_{0}\\ 0&{\textrm{else,}}\end{cases}

with the property that pi=1pi=1 and ip1ip\simeq 1 via the homotopy h=σ|𝕏ah=-\sigma|_{\mathbb{X}_{a}^{\circ}}. The maps ii and pp are pictured in following diagram.

(4.6.1) 0\textstyle{0}ΛnSa1\textstyle{\Lambda^{n}\otimes S_{a-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{\ \ p}Ln1,a\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces L_{n-1,a}}i\scriptstyle{\ \ i}\textstyle{\ \vspace{10mm}\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Λn1Sa1\textstyle{\Lambda^{n-1}\otimes S_{a-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{\ \ p}Ln2,a\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces L_{n-2,a}}i\scriptstyle{\ \ i}\textstyle{\vdots}\textstyle{\vdots}\textstyle{\vdots}\textstyle{\vdots}ΛaS1\textstyle{\Lambda^{a}\otimes S_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Λ2Sa1\textstyle{\Lambda^{2}\otimes S_{a-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{\ \ p}L1,a\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces L_{1,a}}i\scriptstyle{\ \ i}ΛaS0\textstyle{\Lambda^{a}\otimes S_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Λa1S1\textstyle{\Lambda^{a-1}\otimes S_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Λ1Sa1\textstyle{\Lambda^{1}\otimes S_{a-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{\ \ p}L0,a\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces L_{0,a}}i\scriptstyle{\ \ i}Λa1S0\textstyle{\Lambda^{a-1}\otimes S_{0}}\textstyle{\vdots}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Λ0Sa1\textstyle{\Lambda^{0}\otimes S_{a-1}}R\textstyle{R\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i\scriptstyle{\ \ i}\textstyle{\vdots}Λ2S1\textstyle{\Lambda^{2}\otimes S_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}Λ2S0\textstyle{\Lambda^{2}\otimes S_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Λ1S1\textstyle{\Lambda^{1}\otimes S_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}Λ1S0\textstyle{\Lambda^{1}\otimes S_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Λ0S1\textstyle{\Lambda^{0}\otimes S_{1}}Λ0S0\textstyle{\Lambda^{0}\otimes S_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}

Next we apply the Pertubation Lemma, adding the missing vertical differentials dd of 𝕏a\mathbb{X}_{a} and dd of 𝕃a\mathbb{L}_{a}. More precisely, consider the perturbation δ=d\delta=d on 𝕏a\mathbb{X}^{\circ}_{a}; this is a small perturbation since the double complex 𝕏a\mathbb{X}_{a} is bounded. First, we check that that the differentials on 𝕃a\mathbb{L}_{a} obtained in this way are the original differentials on 𝕃a\mathbb{L}_{a}. This is because one has

𝕃a+pδi=0+p(1+(dh)+(dh)2+)di=pdi=dpi=d\partial^{\mathbb{L}_{a}^{\circ}}+p_{\infty}\delta i=0+p(1+(dh)+(dh)^{2}+\cdots)di=pdi=dpi=d

where the second equality follows from the fact that pp vanishes on most of the diagram, the third one follows from the fact that pp is defined using κ\kappa and ε\varepsilon, as well as the commutativity of diagram (4.1.1) and the properties of ε\varepsilon, and the last one is because pi=1pi=1.

In summary, one gets a homotopy equivalence

(4.6.2) (𝕏a,𝕏a=κ+d)p[]i----(𝕃a,𝕃a=d)(\mathbb{X}_{a},\partial^{\mathbb{X}_{a}}\!\!=\kappa+d)\stackrel{{\scriptstyle[}}{{p}}_{\infty}]{i_{\infty}}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}(\mathbb{L}_{a},\partial^{\mathbb{L}_{a}}=d)

For later use, we calculate the new chain maps ii_{\infty} and pp_{\infty}, as well as the associated homotopy hh_{\infty} using the formulas in Definition 3.2. The map ii_{\infty} is given by

i=(1+(hδ)+(hδ)2+)ii_{\infty}=(1+(h\delta)+(h\delta)^{2}+\cdots)i

where δ=d\delta=d, and this can be written as

(4.6.3) i={(1+(σd)+(σd)2+)σ on Li,aε1 on R.i_{\infty}=\begin{cases}(1+(-\sigma d)+(-\sigma d)^{2}+\cdots)\sigma&{\text{ on }}L_{i,a}\\ \varepsilon^{-1}&{\text{ on }}R.\end{cases}

In contrast, the map pp_{\infty} is remarkably simpler since pp equals zero on most of its domain. Indeed it is given by

p=p(1+(δh)+(δh)2+)p_{\infty}=p(1+(\delta h)+(\delta h)^{2}+\cdots)

which can be written as

(4.6.4) p={κon ΛiSa1 for i>0εon Λ0Sj for all j0else.p_{\infty}=\begin{cases}\kappa&{\textrm{on }}\Lambda^{i}\otimes S_{a-1}{\textrm{ for }}i>0\\ \varepsilon&{\textrm{on }}\Lambda^{0}\otimes S_{j}{\textrm{ for all }}j\\ 0&{\textrm{else.}}\end{cases}

We record also the resulting homotopy for ip1i_{\infty}p_{\infty}\simeq 1, which is

(4.6.5) h=h(1+(δh)+(δh)2+)=σ(1+(dσ)+(dσ)2+)\displaystyle\begin{aligned} h_{\infty}&=h(1+(\delta h)+(\delta h)^{2}+\cdots)\\ &=-\sigma(1+(-d\sigma)+(-d\sigma)^{2}+\cdots)\end{aligned}

The map pp_{\infty} has the form pictured in the following diagram.

(4.6.6) 0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ΛnSa1\textstyle{\Lambda^{n}\otimes S_{a-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{\ \ p_{\infty}}Ln1,a\textstyle{L_{n-1,a}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}\textstyle{\ \vspace{10mm}\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Λn1Sa1\textstyle{\Lambda^{n-1}\otimes S_{a-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{\ \ p_{\infty}}Ln2,a\textstyle{L_{n-2,a}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ΛaS1\textstyle{\Lambda^{a}\otimes S_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Λ2Sa1\textstyle{\Lambda^{2}\otimes S_{a-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{\ \ p_{\infty}}L1,a\textstyle{L_{1,a}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}ΛaS0\textstyle{\Lambda^{a}\otimes S_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Λa1S1\textstyle{\Lambda^{a-1}\otimes S_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Λ1Sa1\textstyle{\Lambda^{1}\otimes S_{a-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{\ \ p_{\infty}}L0,a\textstyle{L_{0,a}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}Λa1S0\textstyle{\Lambda^{a-1}\otimes S_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Λ0Sa1\textstyle{\Lambda^{0}\otimes S_{a-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{\ \ p_{\infty}}R\textstyle{R}\textstyle{\vdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Λ2S1\textstyle{\Lambda^{2}\otimes S_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}Λ2S0\textstyle{\Lambda^{2}\otimes S_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Λ1S1\textstyle{\Lambda^{1}\otimes S_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots}Λ1S0\textstyle{\Lambda^{1}\otimes S_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Λ0S1\textstyle{\Lambda^{0}\otimes S_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p_{\infty}}Λ0S0\textstyle{\Lambda^{0}\otimes S_{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p_{\infty}}

We now define an explicit contracting homotopy σ\sigma on the rows of 𝕊\mathbb{S} that can be used to complete the argument in 4.6. This turns out to be nothing but a scaled version of the de Rham differential.

4.7

In view of Proposition 3.5, in order to transfer the dg algebra structure from 𝕏a\mathbb{X}_{a} to 𝕃a\mathbb{L}_{a} we need a special deformation retract, that is, we need a homotopy hh with the properties listed in (3.3.1). As explained in 4.6 in view of (4.5), this comes down to finding a contracting homotopy σ\sigma with σ2=0\sigma^{2}=0 on the rows

Λi1Sm+1ΛiSmΛi+1Sm1\cdots\to\Lambda^{i-1}\otimes S_{m+1}\to\Lambda^{i}\otimes S_{m}\to\Lambda^{i+1}\otimes S_{m-1}\to\cdots

of the entire diagram 𝕊\mathbb{S} displayed in (4.1.1) with the property that i+m>0i+m>0.

Assume now that the field kk has characteristic zero (for the positive characteristic case, see the end of this portion).

Define σi,m:ΛiSmΛi+1Sm1\sigma_{i,m}\colon\Lambda^{i}\otimes S_{m}\to\Lambda^{i+1}\otimes S_{m-1} as

σi,m\displaystyle\sigma_{i,m} (et1etiyp1ypm)\displaystyle(e_{t_{1}}\wedge\cdots\wedge e_{t_{i}}\otimes y_{p_{1}}\cdots y_{p_{m}}) =1i+mj=1mepjet1etiyp1y^pjypm\displaystyle=\frac{1}{i+m}\sum_{j=1}^{m}e_{p_{j}}\wedge e_{t_{1}}\wedge\cdots\wedge e_{t_{i}}\otimes y_{p_{1}}\cdots\hat{y}_{p_{j}}\cdots y_{p_{m}}

where it is understood that σi,m=0\sigma_{i,m}=0 when the target of the map is the zero module, that is, when m=0m=0 or i=ni=n. This can also be written as a scaled de Rham differential

σi,m=1i+mj=1nejyj.\sigma_{i,m}=\frac{1}{i+m}\ \sum_{j=1}^{n}e_{j}\otimes\frac{\partial}{\partial y_{j}}.

To address the case of positive characteristic pp, note that in general we only need to define a contracting homotopy σi,m\sigma_{i,m} for mam\leq a when we apply (4.5) to truncate the complex at position a1a-1, and so it suffices to assume pn+ap\geq n+a. This ensures that when necessary one has 1i+mk\frac{1}{i+m}\in k.

In Lemma 4.8, we show σ\sigma is a contracting homotopy with σ2=0\sigma^{2}=0. In view of Lemma 1.4, we also require the homotopy to satisfy the generalized Leibniz property; this also comes down to the same property for σ\sigma on the rows of the entire diagram 𝕊\mathbb{S} in (4.1.1), which we verify in Lemma 4.9.

A contracting homotopy was defined previously by Srinivasan in [Sri89]. However, it does not satisfy either required property. The map σ\sigma defined above is more symmetric (it is invariant under permutations of the variables) and hence ends up having its square equal to zero and satisfying the generalized Leibniz rule, in fact, the stronger scaled Leibniz rule, as we see in the next two results.

4.8 Lemma.

Consider the maps σ\sigma defined in (4.7) on the rows of diagram (4.1.1) in which the indices sum to a positive number.

If RR has characteristic zero, the maps σ\sigma give a contracting homotopy (that is, a null homotopy for the identity map) on the rows and satisfy σ2=0\sigma^{2}=0.

If RR has positive characteristic pp, the same conclusions hold for σi,m\sigma_{i,m} with ma1m\leq a-1 as long as pn+ap\geq n+a.

Proof.

First we show that κσ+σκ=1\kappa\sigma+\sigma\kappa=1. At the ends of the rows one can show this easily, so we may work with basis elements of ΛiSm\Lambda^{i}\otimes S_{m} with m,i>0m,i>0. We compute κσ\kappa\sigma and σκ\sigma\kappa separately. The reader should note that we do not replace any repeated factors ejeje_{j}\wedge e_{j} with 0, noting that the formula for the Koszul differential κ\kappa gives the same output for either form of input.

For any α=et1etiyp1ypmΛiSm\alpha=e_{t_{1}}\wedge\cdots\wedge e_{t_{i}}\otimes y_{p_{1}}\cdots y_{p_{m}}\in\Lambda^{i}\otimes S_{m}, one has

κi+1,m1σi,m(α)\displaystyle\kappa_{i+1,m-1}\sigma_{i,m}(\alpha)
=1i+mj=1m[α+u=1i(1)uepjet1e^tuetiytuyp1y^pjypm]\displaystyle=\frac{1}{i+m}\sum_{j=1}^{m}[\alpha+\sum_{u=1}^{i}(-1)^{u}e_{p_{j}}\wedge e_{t_{1}}\wedge\cdots\hat{e}_{t_{u}}\cdots\wedge e_{t_{i}}\otimes y_{t_{u}}y_{p_{1}}\cdots\hat{y}_{p_{j}}\cdots y_{p_{m}}]
=1i+m[mα+j=1mu=1i(1)uepjet1e^tuetiytuyp1y^pjypm]\displaystyle=\frac{1}{i+m}[m\alpha+\sum_{j=1}^{m}\sum_{u=1}^{i}(-1)^{u}e_{p_{j}}\wedge e_{t_{1}}\wedge\cdots\hat{e}_{t_{u}}\cdots\wedge e_{t_{i}}\otimes y_{t_{u}}y_{p_{1}}\cdots\hat{y}_{p_{j}}\cdots y_{p_{m}}]

and

σi1,m+1κi,m(α)\displaystyle\sigma_{i-1,m+1}\kappa_{i,m}(\alpha)
=u=1i(1)u+11i+m[α+j=1mepjet1e^tuetiytuyp1y^pjypm]\displaystyle=\sum_{u=1}^{i}(-1)^{u+1}\frac{1}{i+m}[\alpha+\sum_{j=1}^{m}e_{p_{j}}\wedge e_{t_{1}}\wedge\cdots\hat{e}_{t_{u}}\cdots\wedge e_{t_{i}}\otimes y_{t_{u}}y_{p_{1}}\cdots\hat{y}_{p_{j}}\cdots y_{p_{m}}]
=1i+m[iα+j=1mu=1i(1)u+1epjet1e^tuetiytuyp1y^pjypm]\displaystyle=\frac{1}{i+m}[i\alpha+\!\sum_{j=1}^{m}\sum_{u=1}^{i}(-1)^{u+1}e_{p_{j}}\wedge e_{t_{1}}\wedge\cdots\hat{e}_{t_{u}}\cdots\wedge e_{t_{i}}\otimes y_{t_{u}}y_{p_{1}}\!\cdots\hat{y}_{p_{j}}\cdots y_{p_{m}}]

Thus for m,i>0m,i>0 one has

(κi+1,m1σi,m+σi1,m+1κi,m)(α)=(mi+m)α+(ii+m)α=α\displaystyle(\kappa_{i+1,m-1}\sigma_{i,m}+\sigma_{i-1,m+1}\kappa_{i,m})(\alpha)=(\frac{m}{i+m})\alpha+(\frac{i}{i+m})\alpha=\alpha

Next, to see that σ2=0\sigma^{2}=0, one computes for m2m\geq 2

σi+1,m1σi,m(α)=1(i+m)2j=1mu=1ujmepuepjet1etiyp1y^puy^pjypm\displaystyle\sigma_{i+1,m-1}\sigma_{i,m}(\alpha)=\frac{1}{(i+m)^{2}}\sum_{j=1}^{m}\sum_{{u=1}\atop{u\neq j}}^{m}e_{p_{u}}\wedge e_{p_{j}}\wedge e_{t_{1}}\wedge\cdots\wedge e_{t_{i}}\otimes y_{p_{1}}\cdots\hat{y}_{p_{u}}\cdots\hat{y}_{p_{j}}\cdots y_{p_{m}}

which is zero since epuepj=epjepue_{p_{u}}\wedge e_{p_{j}}=-e_{p_{j}}\wedge e_{p_{u}}.

Note that this proof works in positive characteristic as long as the maps σi,m\sigma_{i,m} are defined, which is guaranteed by the hypotheses. ∎

4.9 Lemma.

If RR has characteristic zero, the maps σ\sigma defined in (4.7) satisfy the scaled Leibniz rule. More precisely, when αΛiSa\alpha\in\Lambda^{i}\otimes S_{a} , βΛjSb\beta\in\Lambda^{j}\otimes S_{b} with i+a+j+bi+a+j+b positive, the maps σ\sigma satisfy

σ(αβ)=1i+a+j+b((i+a)σ(α)β+(1)i(j+b)ασ(β))\sigma(\alpha\beta)=\frac{1}{i+a+j+b}\left((i+a)\sigma(\alpha)\beta+(-1)^{i}(j+b)\alpha\sigma(\beta)\right)

and when i=a=j=b=0i=a=j=b=0 one has that σ(αβ)\sigma(\alpha\beta), σ(α)\sigma(\alpha), and σ(β)\sigma(\beta) are all 0.

If RR has positive characteristic pp, the same conclusion holds as long as pi+a+j+bp\geq i+a+j+b.

Proof.

Without loss of generality, we may assume that

α=et1etiyp1ypaandβ=es1esjyq1yqb.\alpha=e_{t_{1}}\wedge\cdots\wedge e_{t_{i}}\otimes y_{p_{1}}\cdots y_{p_{a}}~{}\text{and}~{}\beta=e_{s_{1}}\wedge\cdots\wedge e_{s_{j}}\otimes y_{q_{1}}\cdots y_{q_{b}}.

with i+a+j+b>0i+a+j+b>0.

Then one has

σ(αβ)\displaystyle\sigma(\alpha\beta) =σ(et1eties1esjyp1ypayq1yqb)\displaystyle=\sigma(e_{t_{1}}\wedge\cdots\wedge e_{t_{i}}\wedge e_{s_{1}}\wedge\cdots\wedge e_{s_{j}}\otimes y_{p_{1}}\cdots y_{p_{a}}y_{q_{1}}\cdots y_{q_{b}})
=1i+a+j+b(u=1aepuet1esjyp1y^pjypayq1yqb\displaystyle=\frac{1}{i+a+j+b}\left(\sum_{u=1}^{a}e_{p_{u}}\wedge e_{t_{1}}\wedge\cdots\wedge e_{s_{j}}\otimes y_{p_{1}}\cdots\hat{y}_{p_{j}}\cdots y_{p_{a}}y_{q_{1}}\cdots y_{q_{b}}\right.
+v=1beqvet1esjyp1ypayq1y^qvyqb)\displaystyle\hskip 71.13188pt+\left.\sum_{v=1}^{b}e_{q_{v}}\wedge e_{t_{1}}\wedge\cdots\wedge e_{s_{j}}\otimes y_{p_{1}}\cdots y_{p_{a}}y_{q_{1}}\cdots\hat{y}_{q_{v}}\cdots y_{q_{b}}\right)
=1i+a+j+b(u=1a(epuet1etiyp1y^pjypa)(β)\displaystyle=\frac{1}{i+a+j+b}\left(\sum_{u=1}^{a}(e_{p_{u}}\wedge e_{t_{1}}\wedge\cdots\wedge e_{t_{i}}\otimes y_{p_{1}}\cdots\hat{y}_{p_{j}}\cdots y_{p_{a}})(\beta)\right.
+v=1b(1)i(α)(eqves1esjyq1y^qvyqb))\displaystyle\hskip 71.13188pt+\left.\sum_{v=1}^{b}(-1)^{i}(\alpha)(e_{q_{v}}\wedge e_{s_{1}}\wedge\cdots\wedge e_{s_{j}}\otimes y_{q_{1}}\cdots\hat{y}_{q_{v}}\cdots y_{q_{b}})\right)
=1i+a+j+b((i+a)σ(α)β+(1)i(j+b)ασ(β))\displaystyle=\frac{1}{i+a+j+b}\left((i+a)\sigma(\alpha)\beta+(-1)^{i}(j+b)\alpha\sigma(\beta)\right)

where the relevant terms are 0 when either i+ai+a or j+bj+b is 0. ∎

Next we put together all the ingredients from this section to obtain our main application of our homotopy descent results. The proof is an application of Proposition 3.5 to the deformation retract obtained in (4.6) with the homotopy hh_{\infty} defined in (4.6.5) obtained from the homotopy σ\sigma on the rows of the diagram (4.1.1) satisfying the scaled Leibniz rule and hence the generalized Leibniz rule; see (4.7), (4.8), and (4.9). See also the overview in the paragraph before (4.6). One note: one need only check the homotopy hh on 𝕏a\mathbb{X}_{a} satisfies the scaled Leibniz rule for products that land in ΛiSm\Lambda^{i}\otimes S_{m} for ini\leq n and m<am<a since otherwise the product is zero and the result is trivial; this explains why we need only take pa+np\geq a+n in the statement below.

4.10 Theorem.

Let aa be a positive integer. Suppose that kk is a field of characteristic zero or positive characteristic pa+np\geq a+n. Consider the deformation retract obtained in (4.6)

𝕏ap[]i----𝕃a\mathbb{X}_{a}\stackrel{{\scriptstyle[}}{{p}}_{\infty}]{i_{\infty}}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}\mathbb{L}_{a}

with the associated homotopy hh_{\infty} defined in (4.6.5) using σ\sigma from (4.7), where ii_{\infty} and pp_{\infty} are defined as in (4.6.3) and (4.6.4).

Defining the product of α,β𝕃a\alpha,\beta\in\mathbb{L}_{a} by

αβ=p(i(α)i(β))\alpha\beta=p_{\infty}\left(i_{\infty}(\alpha)i_{\infty}(\beta)\right)

yields a dg algebra structure on 𝕃a\mathbb{L}_{a}. Furthermore, with this structure the map ii_{\infty} is a homomorphism of dg algebras.

4.11 Remark.

The product given in the theorem above can be described explicitly, using the definitions of ii_{\infty} and pp_{\infty}, as follows.

Consider elements α,β𝕃a\alpha,\beta\in\mathbb{L}_{a}. If one of them is in (𝕃a)0=R(\mathbb{L}_{a})_{0}=R then their product is the one coming from the RR-module structure of each (𝕃a)i(\mathbb{L}_{a})_{i}. If both have positive degree, then

αβ=κ(α~β~)\alpha\beta=\kappa(\widetilde{\alpha}\widetilde{\beta})

where κ\kappa is defined in (4.1.2) and

α~\displaystyle\widetilde{\alpha} =(1+(σd)+(σd)2+)σ(α)\displaystyle=(1+(-\sigma d)+(-\sigma d)^{2}+\cdots)\sigma(\alpha)
β~\displaystyle\widetilde{\beta} =(1+(σd)+(σd)2+)σ(β)\displaystyle=(1+(-\sigma d)+(-\sigma d)^{2}+\cdots)\sigma(\beta)

where the scaled de Rham differential σ\sigma is defined in 4.7.

4.12 Remark.

Because of the symmetric way in which the maps κ\kappa, dd and hh are defined, the dg algebra structure defined on 𝕃a\mathbb{L}_{a} in Theorem 4.10 is invariant under the action of the symmetric group on the polynomial ring.

4.13 Remark.

One may note that our algebra structure is, in fact, basis free, although we do not describe it in a basis free way. It is well known that the differentials in the complex from which we descend our structure are so, and one can see that the homotopy is as well, as it is just a scaled version of the de Rham map.

5. Comparison maps

In this section we use the results from the previous sections to obtain comparison maps lifting the natural surjections R/𝔪bR/𝔪aR/\mathfrak{m}^{b}\longrightarrow R/\mathfrak{m}^{a} for any bab\geq a to their respective minimal free resolutions 𝕃b\mathbb{L}_{b} and 𝕃a\mathbb{L}_{a}, and these maps turn out to be dg algebra morphisms (for the dg algebra structures placed on them in the previous section). Since 𝕃1\mathbb{L}_{1} is simply the Koszul complex KK on the variables, this yields that KK is a dg algebra over 𝕃b\mathbb{L}_{b} for each b1b\geq 1.

To set up the statement, recall from (4.6.2) that for any cc there is a homotopy equivalence

𝕏cp[]i----𝕃c\mathbb{X}_{c}\stackrel{{\scriptstyle[}}{{p}}_{\infty}]{i_{\infty}}{\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\leftarrow$\cr\vrule width=0.0pt,height=2.58333pt$\hfil\scriptstyle\relbar$\cr}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\cr\vrule width=0.0pt,height=2.58333pt$\scriptstyle\relbar$}}}\joinrel\mathrel{\raise 3.87498pt\hbox{\oalign{$\scriptstyle\relbar$\hfil\cr$\scriptstyle\vrule width=0.0pt,height=1.80832pt\smash{\rightarrow}$\cr}}}}\mathbb{L}_{c}

pictured in (4.6.6) that is used in Theorem 4.10 to place a dg algebra structure on 𝕃c\mathbb{L}_{c} for which ii_{\infty} is a dg algebra homomorphism. Although the value of cc varies below, it should be clear from the context which ii_{\infty} and pp_{\infty} maps are being applied. Recall also from (4.4) that we can view 𝕏c\mathbb{X}_{c} as the quotient 𝕊/trc(𝕊)\mathbb{S}/\textrm{tr}_{\geq c}(\mathbb{S}) of the dg algebra 𝕊\mathbb{S} by the dg ideal

trc(𝕊)={ΛiSjjc}\textrm{tr}_{\geq c}(\mathbb{S})=\{\Lambda^{i}\otimes S_{j}\mid j\geq c\}

In this way, 𝕏c\mathbb{X}_{c} inherits the dg algebra structure from 𝕊\mathbb{S}. Therefore, if bab\geq a, the inclusion of dg ideals trb(𝕊)tra(𝕊)\textrm{tr}_{\geq b}(\mathbb{S})\hookrightarrow\textrm{tr}_{\geq a}(\mathbb{S}) gives a natural quotient map

πb,a:𝕏b=𝕊/trb(𝕊)𝕊/tra(𝕊)=𝕏a\pi_{b,a}\colon\mathbb{X}_{b}=\mathbb{S}/\textrm{tr}_{\geq b}(\mathbb{S})\twoheadrightarrow\mathbb{S}/\textrm{tr}_{\geq a}(\mathbb{S})=\mathbb{X}_{a}

which has the effect of sending the columns ΛiSj\Lambda^{i}\otimes S_{j} to zero for ajb1a\leq j\leq b-1. This is clearly a homomorphism of dg algebras.

5.1 Theorem.

Let aa and bb be positive integers with bab\geq a. The chain map

fb,a=pπb,ai:𝕃b𝕃af_{b,a}=p_{\infty}\pi_{b,a}i_{\infty}\colon\mathbb{L}_{b}\to\mathbb{L}_{a}

is a homomorphism of dg algebras that gives a lifting of the natural surjection R/𝔪bR/𝔪aR/\mathfrak{m}^{b}\longrightarrow R/\mathfrak{m}^{a}. In particular, the Koszul complex on the variables, which is 𝕃1\mathbb{L}_{1}, is a dg algebra over 𝕃b\mathbb{L}_{b} for every positive integer bb.

Moreover, if cbac\geq b\geq a then fc,a=fb,afc,bf_{c,a}=f_{b,a}f_{c,b}.

Proof.

First note that fb,af_{b,a} is a chain map since it is a composition of chain maps. Also, (fb,a)0(f_{b,a})_{0} is the identity map on RR; thus fb,af_{b,a} gives a lifting of the natural surjection R/𝔪bR/𝔪aR/\mathfrak{m}^{b}\longrightarrow R/\mathfrak{m}^{a}.

Next we show that fb,af_{b,a} is a homomorphism of dg algebras. Clearly, if b=ab=a then fb,af_{b,a} is the identity map. So we may assume that b>ab>a. Let α\alpha and β\beta be homogeneous elements of 𝕃b\mathbb{L}_{b}. If either sits in degree 0, then fb,a(αβ)=fb,a(α)fb,a(β)f_{b,a}(\alpha\beta)=f_{b,a}(\alpha)f_{b,a}(\beta) as fb,af_{b,a} is a homomorphism of RR-modules. So we may assume that αLi,b\alpha\in L_{i,b} and βLj,b\beta\in L_{j,b} for some 0i,jn0\leq i,j\leq n. Since πb,a\pi_{b,a} and ii_{\infty} are homomorphisms of dg algebras, one has

fb,a(αβ)\displaystyle f_{b,a}(\alpha\beta) =pπb,ai(αβ)\displaystyle=p_{\infty}\pi_{b,a}i_{\infty}(\alpha\beta)
=p(πb,ai(α)trb,ai(β))\displaystyle=p_{\infty}(\pi_{b,a}i_{\infty}(\alpha)\textrm{tr}_{b,a}i_{\infty}(\beta))

We pause to compute the composition

(5.1.1) πb,ai=πb,a(1+(hδ)++(hδ)b1)i=((hδ)ba++(hδ)b1)i\displaystyle\begin{aligned} \pi_{b,a}i_{\infty}&=\pi_{b,a}(1+(h\delta)+\cdots+(h\delta)^{b-1})i\\ &=((h\delta)^{b-a}+\cdots+(h\delta)^{b-1})i\end{aligned}

and so we have

fb,a(αβ)\displaystyle f_{b,a}(\alpha\beta) =p([(hδ)ba++(hδ)b1]i(α)[(hδ)ba++(hδ)b1]i(β))\displaystyle=p_{\infty}\!\!\left([(h\delta)^{b-a}+\cdots+(h\delta)^{b-1}]i(\alpha)\,[(h\delta)^{b-a}+\cdots+(h\delta)^{b-1}]i(\beta)\right)

From (5.1.1) we also get an alternate formula for fb,af_{b,a} as follows which we use in the next part

(5.1.2) fb,a=pπb,ai=p((hδ)ba++(hδ)b1)i=p(hδ)bai\displaystyle\begin{aligned} f_{b,a}&=p_{\infty}\pi_{b,a}i_{\infty}\\ &=p_{\infty}((h\delta)^{b-a}+\cdots+(h\delta)^{b-1})i\\ &=p_{\infty}(h\delta)^{b-a}i\end{aligned}

where the other terms disappear as they are in the portion of the domain where pp_{\infty} equals 0. Not that in the last line pp_{\infty} can be replaced by pp.

Next we compute

fb,a(α)fb,a(β)\displaystyle f_{b,a}(\alpha)f_{b,a}(\beta) =p(i(fb,a(α))i(fb,a(β)))\displaystyle=p_{\infty}(i_{\infty}(f_{b,a}(\alpha))\,i_{\infty}(f_{b,a}(\beta)))

by the definition of the product in 𝕃a\mathbb{L}_{a}. This is equal to fb,a(αβ)f_{b,a}(\alpha\beta) because

(5.1.3) ifb,a=ip(hδ)bai=(1+(hδ)++(hδ)a1)ip[hδ(hδ)ba1]i=(1+(hδ)++(hδ)a1)σκ(σ)δ(hδ)ba1i=(1+(hδ)++(hδ)a1)(σ)δ(hδ)ba1i=(1+(hδ)++(hδ)a1)(hδ)bai=((hδ)ba++(hδ)b1)i\displaystyle\begin{aligned} i_{\infty}f_{b,a}&=i_{\infty}p_{\infty}(h\delta)^{b-a}i\\ &=(1+(h\delta)+\cdots+(h\delta)^{a-1})ip[h\delta(h\delta)^{b-a-1}]i\\ &=(1+(h\delta)+\cdots+(h\delta)^{a-1})\sigma\kappa(-\sigma)\delta(h\delta)^{b-a-1}i\\ &=(1+(h\delta)+\cdots+(h\delta)^{a-1})(-\sigma)\delta(h\delta)^{b-a-1}i\\ &=(1+(h\delta)+\cdots+(h\delta)^{a-1})(h\delta)^{b-a}i\\ &=((h\delta)^{b-a}+\cdots+(h\delta)^{b-1})i\end{aligned}

where the first equality is by (5.1.2), the second one is by the definitions of ii_{\infty} and pp_{\infty}, the third one is by the definitions of ii, pp, and hh, the fourth one is because σκσ=σ\sigma\kappa\sigma=\sigma since the homotopy σ\sigma satisfies σκ=1κσ\sigma\kappa=1-\kappa\sigma and σ2=0\sigma^{2}=0, and the fifth is because h=σh=-\sigma. Note that when we apply the definitions of ii, pp, and hh we are using that the terms are in ΛjS\Lambda^{j}\otimes S for j>0j>0.

Last we compute the composition

fb,afc,b\displaystyle f_{b,a}f_{c,b} =(pπb,ai)fc,b\displaystyle=\left(p_{\infty}\pi_{b,a}i_{\infty}\right)f_{c,b}
=pπb,a((hδ)cb++(hδ)c1)i\displaystyle=p_{\infty}\pi_{b,a}((h\delta)^{c-b}+\cdots+(h\delta)^{c-1})i
=p(hδ)cai=fc,a\displaystyle=p_{\infty}(h\delta)^{c-a}i=f_{c,a}

where the second equality is by (5.1.3), the third is from the definitions of the maps, and the last is by (5.1.2). as desired. ∎

Note that for a=1a=1, of course, the map p:𝕏1𝕃1p_{\infty}\colon\mathbb{X}_{1}\to\mathbb{L}_{1} is an isomorphism of complexes, hence it is trivial that fb,af_{b,a} is a homomorphism of dg algebras since πb,a\pi_{b,a} and ii_{\infty} always are.

5.2 Remark.

In the proof above, the following more explicit formula for the map fb,a:𝕃b𝕃af_{b,a}\colon\mathbb{L}_{b}\to\mathbb{L}_{a} was derived; see 5.1.2.

(5.2.1) fb,a=p(hδ)baif_{b,a}=p(h\delta)^{b-a}i

As a consequence we see that for j>0j>0

(5.2.2) im(fb,a)j𝔪ba(La)j\operatorname{im}(f_{b,a})_{j}\subseteq\mathfrak{m}^{b-a}\mathbb{(}L_{a})_{j}

since the map δ=d=kos1\delta=d=\textrm{kos}\otimes 1 has image in 𝔪\mathfrak{m} times the next free module as kos is the differential in the Koszul complex on the variables. This may also be seen in an elementary way using long exact sequences of Tor modules.

Acknowledgements

We thank Srikanth Iyengar for suggesting the Perturbation Lemma to us, Ben Briggs for discussions about AA_{\infty}-algebras, and Luchezar Avramov for some helpful comments. We also Daniel Murfet for asking us questions that encouraged us to explore the connections to the Homotopy Transfer Theorem.

References

  • [Avr81] Luchezar L. Avramov, Obstructions to the existence of multiplicative structures on minimal free resolutions, Amer. J. Math. 103 (1981), no. 1, 1–31. MR 601460
  • [Avr98] by same author, Infinite free resolutions, Six lectures on commutative algebra (Bellaterra, 1996), Progr. Math., vol. 166, Birkhäuser, Basel, 1998, pp. 1–118. MR MR1648664 (99m:13022)
  • [BE75] David A. Buchsbaum and David Eisenbud, Generic free resolutions and a family of generically perfect ideals, Advances in Math. 18 (1975), no. 3, 245–301. MR 0396528
  • [BE77] by same author, Algebra structures for finite free resolutions, and some structure theorems for ideals of codimension 33, Amer. J. Math. 99 (1977), no. 3, 447–485. MR 453723
  • [Cra04] Marius Crainic, On the perturbation lemma, and deformations, arXiv:math/0403266.
  • [DM13] Tobias Dyckerhoff and Daniel Murfet, Pushing forward matrix factorizations, Duke Math. J. 162 (2013), no. 7, 1249–1311. MR 3079249
  • [GJ90] Ezra Getzler and John D. S. Jones, AA_{\infty}-algebras and the cyclic bar complex, Illinois J. Math. 34 (1990), no. 2, 256–283. MR 1046565
  • [Her74] Jürgen Herzog, Komplexe auflösungen und dualität in der lokalen algebra, Universität Regensburg, Habilitationsschrift (1974).
  • [Kad80] T. V. Kadeišvili, On the theory of homology of fiber spaces, Uspekhi Mat. Nauk 35 (1980), no. 3(213), 183–188, International Topology Conference (Moscow State Univ., Moscow, 1979). MR 580645
  • [Kad82] T. V. Kadeishvili, The algebraic structure in the homology of an A()A(\infty)-algebra, Soobshch. Akad. Nauk Gruzin. SSR 108 (1982), no. 2, 249–252 (1983). MR 720689
  • [Kel01] Bernhard Keller, Introduction to AA-infinity algebras and modules, Homology Homotopy Appl. 3 (2001), no. 1, 1–35. MR 1854636
  • [Kel06] by same author, AA-infinity algebras, modules and functor categories, Trends in representation theory of algebras and related topics, Contemp. Math., vol. 406, Amer. Math. Soc., Providence, RI, 2006, pp. 67–93. MR 2258042
  • [KM80] Andrew R. Kustin and Matthew Miller, Algebra structures on minimal resolutions of Gorenstein rings of embedding codimension four, Math. Z. 173 (1980), no. 2, 171–184. MR 583384
  • [Kus87] Andrew R. Kustin, Gorenstein algebras of codimension four and characteristic two, Comm. Algebra 15 (1987), no. 11, 2417–2429. MR 912779
  • [LH03] Kenji Lefèvre-Hasegawa, Sur les AA_{\infty}-catégories, PhD thesis, Université Denis Diderot - Paris 7, 2003, arXiv:math/0310337.
  • [Mae01] Takashi Maeda, Minimal algebra resolution associated with hook representations, J. Algebra 237 (2001), no. 1, 287–291. MR 1813892
  • [MR18] Claudia Miller and Hamidreza Rahmati, Free resolutions of Artinian compressed algebras, J. Algebra 497 (2018), 270–301. MR 3743182
  • [Pee96] Irena Peeva, 0-Borel fixed ideals, J. Algebra 184 (1996), no. 3, 945–984. MR 1407879
  • [Sri89] Hema Srinivasan, Algebra structures on some canonical resolutions, J. Algebra 122 (1989), no. 1, 150–187. MR 994942
  • [Sri92] by same author, The nonexistence of a minimal algebra resolution despite the vanishing of Avramov obstructions, J. Algebra 146 (1992), no. 2, 251–266. MR 1152904
  • [Sri96] by same author, A grade five Gorenstein algebra with no minimal algebra resolutions, J. Algebra 179 (1996), no. 2, 362–379. MR 1367854
  • [Sta63] James Dillon Stasheff, Homotopy associativity of HH-spaces. I, II, Trans. Amer. Math. Soc. 108 (1963), 275-292; ibid. 108 (1963), 293–312. MR 0158400
  • [Val14] Bruno Vallette, Algebra + homotopy = operad, Symplectic, Poisson, and noncommutative geometry, Math. Sci. Res. Inst. Publ., vol. 62, Cambridge Univ. Press, New York, 2014, pp. 229–290. MR 3380678
  • [W0̈4] Samuel Wüthrich, Homology of powers of regular ideals, Glasg. Math. J. 46 (2004), no. 3, 571–584. MR 2094811
  • [Wei94] Charles A. Weibel, An introduction to homological algebra, Cambridge Studies in Advanced Mathematics, vol. 38, Cambridge University Press, Cambridge, 1994. MR 1269324