Transferring algebra structures on complexes
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.
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 -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 -operations vanish as well and that if one starts with an -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 of the quotients of a polynomial ring in variables by the th 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 to show that in characteristic zero any -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 -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 -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 . 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 of as a retract of a complex with a dg algebra structure, we transfer the algebra structure as follows. For this, let be the totalization of the double complex defined in (4.1), and let be the quotient of the dg algebra by the dg ideal .
Theorem 4.10. Let be a positive integer. Suppose that is a field of characteristic zero or positive characteristic . Consider the deformation retract
obtained in (4.6). Defining the product of by
yields a dg algebra structure on . Furthermore, with this structure the map is a homomorphism of dg algebras.
The second result gives some very natural dg algebra homomorphisms between these resolutions.
Theorem 5.1. Let and be positive integers with and let
be the natural surjection. The chain map
is a homomorphism of dg algebras that gives a lifting of the natural surjection . In particular, the Koszul complex on the variables, which is , is a dg algebra over for every positive integer .
Moreover, if then .
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 of Buchsbaum and Eisenbud for all . In Section 5 we obtain dg algebra homomorphisms between these resolutions.
In this paper, we assume that the complexes consist of -modules for some commutative ring and that they are graded homologically, rather than cohomologically; the reader should be aware that most sources for 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 -algebra structure. We also discuss there what happens when the original complex is merely an -algebra.
1.1 Definition.
A differential graded algebra over (dg algebra) is a complex of -modules lying in nonnegative degrees equipped with a product given by a chain map
giving an associative and unitary product with .
The fact that the product is a chain map is equivalent to the differentials of satisfying the Leibniz rule:
where denotes the degree of . In addition, we assume that the product is strictly graded commutative, that is, for all and if the degree of is odd.
A homomorphism of dg algebras is a morphism of complexes such that and .
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
with via a chosen homotopy on , that is . Note that sometimes a homotopy equivalence is defined to include the condition , 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 . This condition holds in our applications.
Next we see that deformation retracts that satisfy 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 be a complex of -modules equipped with a product, and let be a graded map (but not necessarily a chain map). We say that satisfies the generalized Leibniz rule if one has
for every and in .
For our application, the map will in fact satisfy a stronger condition, which we call scaled Leibniz rule, namely that for every there are depending only on the degrees of and , respectively, such that
We now prove the main result of this section.
1.4 Proposition.
Let be a dg algebra. Consider a deformation retract
with associated homotopy that satisfies the generalized Leibniz rule and . The following product defines a dg algebra structure on
where the product inside parentheses is the one in .
Moreover, with this structure on , the map becomes a dg algebra homomorphism.
Proof.
The Leibniz rule for holds without the assumptions that satisfies the generalized Leibniz rule and . Indeed, for any elements , one has
where the first equality is from the definition of the product, the second one holds since is a chain map, the third one is from the Leibniz rule for , the fourth holds since 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) |
for all . If one expands this expression using that satisfies the Leibniz rule and satisfies the generalized Leibniz rule, one sees that every term has a factor with , which is zero, or a factor with which is also zero since .
To verify associativity, take any elements . One has
where the first two equalities are from the definition of the product and the third one is by the equality . A similar argument shows that
and hence associativity holds since it holds for .
Finally, one can see that is the identity element of and that the product on is graded commutative since and are graded maps.
To see that the map is a dg algebra homomorphism, let . One then has
where the second equality holds as and form a deformation retract and the last one follows from (1.4.1). ∎
Note that the condition 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 -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 -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 -algebras over a field of characteristic zero. However, these are known to hold over a commutative ring as long as one makes some freeness assumptions. For simplicity we assume in this section that 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 -algebra and the Homotopy Transfer Theorem for a dg algebra. The concept of an -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 -algebra over a ring is a complex of -modules together with -multilinear maps of degree
for each , called operations or multiplications, satisfying the following relations, called the Stasheff identities.
-
•
The first operation is simply the differential:
-
•
The second operation satisfies the Leibniz rule:
-
•
The third one verifies that is associative up to the homotopy :
Note that the left hand side is the obstruction to associativity for and that the right hand side is the boundary of in .
-
•
More generally, for , we have
where the sums are taken over the values of with .
Note that when one applies the maps in each formula above to an element, one should use the Koszul sign rule: For graded maps and , one has
for homogeneous elements and , where denotes the degree of whether it is a map or homogenous element.
Recall that we are using homological notation; in cohomological notation the degree of would be rather than . 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 -algebra whose operations are zero for all is a dg algebra where the product is given by . Conversely, a dg algebra can be given the structure of an -algebra by setting .
However, and can usually be extended to other -algebra structures. Indeed, one can have nonzero higher operations for which the boundary of in is equal to zero and hence 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 , rather than ). 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 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 is drawn as follows.
In this notation, the properties of Leibniz rule and associativity can be drawn as follows.
To justify the first equation, note that will produce no sign when applied to the input since , but will have the sign .
We now recall the Homotopy Transfer Theorem [Kad80] that allows one to transfer a dg algebra structure along a deformation retract yielding an -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
be a deformation retract with associated homotopy where is a dg algebra. As in Remark 2.2, one considers an -algebra with equal to the differential, equal to the dg algebra product on , and for .
The Homotopy Transfer Theorem gives an -structure on as follows: First set . For , the th operation is defined as
where the left hand side is the notation for and where the sum is over , the set of all planar binary rooted trees with 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 , except for the last one, where instead is applied. The actual signs, indicated simply as 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, and are given by
where the signs in the expression for 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 -algebra with all higher operations equal to zero. Proposition 1.4 does yield a dg algebra, and so one could extend it to an -algebra by defining the higher operations 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 be a dg algebra. Consider a deformation retract
with associated homotopy that satisfies the generalized Leibniz rule and . Then the -algebra structure on obtained from the dg algebra structure on via 2.4 has trivial higher operations, that is, for all .
Proof.
Recall from 2.4 that the operations for descended from the dg algebra structure on are signed sums of elements described by planar binary rooted trees with leaves. The signs do not matter as we prove that every term equals zero. Indeed, each term always includes a factor of the form for some (this will be nested inside other maps , and and products from ). This vanishes as satisfies the generalized Leibniz rule and holds. Hence these higher operations all vanish. ∎
What if one begins with a complex that is an -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 -structure along a deformation retract yielding an -structure on the retract. We again use the tree notation introduced above.
2.6
Homotopy Transfer Theorem for -algebras. Let
be a deformation retract with associated homotopy where is an -algebra. The Homotopy Transfer Theorem gives an -structure on as follows: First set . For , the operation is defined as
where the left hand side is the notation for and where the sum is over , the set of all planar (not necessarily binary) rooted trees with 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 . Once again, the pattern is that every such product is followed by an application of , except for the last one, where instead 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 is merely an -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 on satisfies the generalized Leibniz rule for an -algebra if for every one has
where is the th factor and the other factors are . Under this hypothesis, we can show that the formulas for the descended operations via the Homotopy Transfer Theorem for -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 be an -algebra with operations for . Consider a deformation retract
with associated homotopy that satisfies the generalized Leibniz rule for an -algebra and . Then the -algebra structure on obtained from the -algebra structure on via 2.6 has operations given by
for all .
Proof.
Note that in 2.6, it follows from the construction and the properties of a deformation retract that
This covers the cases .
Recall from 2.6 that the operations for descended from the dg algebra structure on are signed sums of elements described by planar rooted trees with leaves. The signs do not matter as all of the terms vanish except one. Indeed, expanding using that and the generalized Leibniz rule for an -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 and not involving the homotopy ; 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
with associated homotopy . A perturbation is a map on of the same degree as the differential such that , that is, is again a differential. The perturbation is called small if is invertible. Most commonly, this happens when is elementwise nilpotent for then one has
where the sum is finite on each element of .
3.2 Definition.
Consider a set of homotopy equivalence data
with associated homotopy . Let be a small perturbation on , and let . We define the following new data
where
and set
Note that when is elementwise nilpotent, then the formulas can be rewritten as follows.
3.3 Definition.
A special deformation retract is a deformation retract that satisfies the following equations
(3.3.1) |
These ensure that the property 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
with associated homotopy , its perturbation via a small perturbation gives a set of homotopy equivalence data
with associated homotopy .
If, furthermore, the original homotopy equivalence is a special deformation retract then so is the resulting one, that is,
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 be a dg algebra. Consider a special deformation retract
with associated homotopy that satisfies the generalized Leibniz rule, and let be a small perturbation on .
If the perturbed complex from Lemma 3.4 remains a dg algebra via the same product (equivalently, if satisfies the Leibniz rule), then the following product defines a dg algebra structure on the perturbed complex
where the product inside parentheses is the one in .
Moreover, with this structure on , the map 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 is a dg algebra homomorphism, one needs to show that
Here this follows by similar reasoning due to the facts that one has
and that and . ∎
4. Application to a minimal resolution
Let be a polynomial ring over a field . In [BE75], Buchsbaum and Eisenbud introduced the minimal free resolution of the quotient of by powers of the homogeneous maximal ideal. In [Sri89], Srinivasan gives a dg algebra structure on using Young tableaux. In this section, we use the Perturbation Lemma in a simple way to obtain a dg algebra structure on that is -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 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 -modules; see Remark 4.2.
Let be a polynomial ring and let be an exterior algebra. Consider the following (anticommutative) double complex whose rows are the strands of the Koszul complex and whose columns are the tensor product over of the Koszul complex over on with the graded pieces of . We denote both it and its totalization by , as it is clear everywhere from the context which we mean. All the tensor products in the diagram are over .
(4.1.1) |
More explicitly, the horizontal differentials are given by
(4.1.2) |
and the vertical differentials are given by
where denotes the th differential in the Koszul complex .
For the totalization of this double complex (or of truncations of it), since it is anticommutative, the differentials are defined as
without adding any signs. For simplicity we write
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 be a -vector space with , and consider the symmetric and exterior algebras
Consider as a module over its enveloping algebra via the multiplication map. Its minimal graded free resolution, after identifying the two copies of with polynomial rings as in the display above, is the Koszul complex on the regular sequence . Rearranging factors, it can be expressed as
with the homological degree being the degree of the middle factor and
where is the Koszul differential on and is the Koszul differential on . Viewing graded strands, one can write this as a totalization of an anticommutative double complex of free -modules given by
Although this double complex differs from the one pictured in (4.1.1) by a sign on the horizontal maps , 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 in place of above; note that it would no longer be a resolution of over its enveloping algebra.
4.3
We introduce the complexes of Buchsbaum and Eisenbud here. They show that this is a minimal -free resolution of , where is the homogeneous maximal ideal of .
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 -modules. So one can define free -modules
in other words, with split exact sequences
The vertical differentials in the diagram induce maps on these modules, which we again denote by , to yield a complex
augmented by the evaluation map
(4.3.1) |
induced by the evaluation map from to sending to .
4.4
Next we define and to be the totalizations of the truncations at column of the anticommutative double complex
respectively, with differentials inherited from . It is well-known that there is a quasi-isomorphism, and hence a homotopy equivalence,
but we will re-derive this via the Perturbation Lemma in order to simultaneously transfer a dg algebra structure from over to (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 itself has a natural dg algebra structure. Indeed, the entire complex is a dg algebra with the obvious multiplication: for and , the product is obtained by multiplying the factors in and in independently. It satisfies the Leibniz rule and other properties of a dg algebra because the differentials and do and because homological degree in the totalization of is, in fact, given by the degree in . With this multiplication, the right truncation is clearly a dg ideal and the quotient complex
is therefore a dg algebra. Concretely, the resulting product on the left truncation is given by the multiplication on with the proviso that any terms landing in with 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 be a contractible complex of -modules, i.e., one that is homotopy equivalent to zero via a homotopy , (i.e., one that is split exact). Denote its truncation at position by
Let denote the stalk complex with this module in degree and 0 modules elsewhere. The chain maps and given in degree by and , respectively, yield a deformation retract
with associated homotopy . Indeed one can easily check that and via the homotopy . Note that one could also use instead of with appropriate and . If, furthermore, the original contracting homotopy satisfies , then the deformation retract is special: and and one always has due to the fact that in degrees .
Next we want to transfer this structure from , which has a dg algebra structure by (4.4) to the minimal free resolution of . By Proposition 3.5, it suffices to find a special deformation retract of the form
that is a perturbation of a special deformation retract whose homotopy satisfies generalized Leibniz. Note that the differential of is exactly . In 4.6, we discuss how one can find this deformation retract, given a contracting homotopy on the higher rows of , 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 and , where is obtained from by replacing the vertical differentials by 0 and is the complex 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 and .
We start by finding a deformation retract of the form
with a homotopy . For rows of except the bottom one, we use (4.5) as follows. Recall that the rows of are split exact with a contracting homotopy that we call (an explicit one is given in Remark 4.7). So each row that gets truncated has a deformation retract onto the image 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 at the bottom of the diagram is not exact and so needs to be dealt with separately in conjunction with . For this we use that there is an isomorphism defined in (4.3.1).
Putting this all together, one obtains chain maps given by
with the property that and via the homotopy . The maps and are pictured in following diagram.
(4.6.1) |
Next we apply the Pertubation Lemma, adding the missing vertical differentials of and of . More precisely, consider the perturbation on ; this is a small perturbation since the double complex is bounded. First, we check that that the differentials on obtained in this way are the original differentials on . This is because one has
where the second equality follows from the fact that vanishes on most of the diagram, the third one follows from the fact that is defined using and , as well as the commutativity of diagram (4.1.1) and the properties of , and the last one is because .
In summary, one gets a homotopy equivalence
(4.6.2) |
For later use, we calculate the new chain maps and , as well as the associated homotopy using the formulas in Definition 3.2. The map is given by
where , and this can be written as
(4.6.3) |
In contrast, the map is remarkably simpler since equals zero on most of its domain. Indeed it is given by
which can be written as
(4.6.4) |
We record also the resulting homotopy for , which is
(4.6.5) |
The map has the form pictured in the following diagram.
(4.6.6) |
We now define an explicit contracting homotopy on the rows of 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 to we need a special deformation retract, that is, we need a homotopy 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 with on the rows
of the entire diagram displayed in (4.1.1) with the property that .
Assume now that the field has characteristic zero (for the positive characteristic case, see the end of this portion).
Define as
where it is understood that when the target of the map is the zero module, that is, when or . This can also be written as a scaled de Rham differential
To address the case of positive characteristic , note that in general we only need to define a contracting homotopy for when we apply (4.5) to truncate the complex at position , and so it suffices to assume . This ensures that when necessary one has .
In Lemma 4.8, we show is a contracting homotopy with . 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 on the rows of the entire diagram 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 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 defined in (4.7) on the rows of diagram (4.1.1) in which the indices sum to a positive number.
If has characteristic zero, the maps give a contracting homotopy (that is, a null homotopy for the identity map) on the rows and satisfy .
If has positive characteristic , the same conclusions hold for with as long as .
Proof.
First we show that . At the ends of the rows one can show this easily, so we may work with basis elements of with . We compute and separately. The reader should note that we do not replace any repeated factors with , noting that the formula for the Koszul differential gives the same output for either form of input.
For any , one has
and
Thus for one has
Next, to see that , one computes for
which is zero since .
Note that this proof works in positive characteristic as long as the maps are defined, which is guaranteed by the hypotheses. ∎
4.9 Lemma.
If has characteristic zero, the maps defined in (4.7) satisfy the scaled Leibniz rule. More precisely, when , with positive, the maps satisfy
and when one has that , , and are all 0.
If has positive characteristic , the same conclusion holds as long as .
Proof.
Without loss of generality, we may assume that
with .
Then one has
where the relevant terms are 0 when either or 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 defined in (4.6.5) obtained from the homotopy 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 on satisfies the scaled Leibniz rule for products that land in for and since otherwise the product is zero and the result is trivial; this explains why we need only take in the statement below.
4.10 Theorem.
Let be a positive integer. Suppose that is a field of characteristic zero or positive characteristic . Consider the deformation retract obtained in (4.6)
with the associated homotopy defined in (4.6.5) using from (4.7), where and are defined as in (4.6.3) and (4.6.4).
Defining the product of by
yields a dg algebra structure on . Furthermore, with this structure the map is a homomorphism of dg algebras.
4.11 Remark.
The product given in the theorem above can be described explicitly, using the definitions of and , as follows.
4.12 Remark.
Because of the symmetric way in which the maps , and are defined, the dg algebra structure defined on 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 for any to their respective minimal free resolutions and , and these maps turn out to be dg algebra morphisms (for the dg algebra structures placed on them in the previous section). Since is simply the Koszul complex on the variables, this yields that is a dg algebra over for each .
To set up the statement, recall from (4.6.2) that for any there is a homotopy equivalence
pictured in (4.6.6) that is used in Theorem 4.10 to place a dg algebra structure on for which is a dg algebra homomorphism. Although the value of varies below, it should be clear from the context which and maps are being applied. Recall also from (4.4) that we can view as the quotient of the dg algebra by the dg ideal
In this way, inherits the dg algebra structure from . Therefore, if , the inclusion of dg ideals gives a natural quotient map
which has the effect of sending the columns to zero for . This is clearly a homomorphism of dg algebras.
5.1 Theorem.
Let and be positive integers with . The chain map
is a homomorphism of dg algebras that gives a lifting of the natural surjection . In particular, the Koszul complex on the variables, which is , is a dg algebra over for every positive integer .
Moreover, if then .
Proof.
First note that is a chain map since it is a composition of chain maps. Also, is the identity map on ; thus gives a lifting of the natural surjection .
Next we show that is a homomorphism of dg algebras. Clearly, if then is the identity map. So we may assume that . Let and be homogeneous elements of . If either sits in degree 0, then as is a homomorphism of -modules. So we may assume that and for some . Since and are homomorphisms of dg algebras, one has
We pause to compute the composition
(5.1.1) |
and so we have
From (5.1.1) we also get an alternate formula for as follows which we use in the next part
(5.1.2) |
where the other terms disappear as they are in the portion of the domain where equals 0. Not that in the last line can be replaced by .
Next we compute
by the definition of the product in . This is equal to because
(5.1.3) |
where the first equality is by (5.1.2), the second one is by the definitions of and , the third one is by the definitions of , , and , the fourth one is because since the homotopy satisfies and , and the fifth is because . Note that when we apply the definitions of , , and we are using that the terms are in for .
Note that for , of course, the map is an isomorphism of complexes, hence it is trivial that is a homomorphism of dg algebras since and always are.
5.2 Remark.
In the proof above, the following more explicit formula for the map was derived; see 5.1.2.
(5.2.1) |
As a consequence we see that for
(5.2.2) |
since the map has image in 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 -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 , 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, -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 -algebra, Soobshch. Akad. Nauk Gruzin. SSR 108 (1982), no. 2, 249–252 (1983). MR 720689
- [Kel01] Bernhard Keller, Introduction to -infinity algebras and modules, Homology Homotopy Appl. 3 (2001), no. 1, 1–35. MR 1854636
- [Kel06] by same author, -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 -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, -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 -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