Lie-Rinehart algebras Acyclic Lie -algebroids
Abstract.
We show that there is an equivalence of categories between Lie-Rinehart algebras over a commutative algebra and homotopy equivalence classes of negatively graded Lie -algebroids over their resolutions (=acyclic Lie -algebroids). This extends to a purely algebraic setting the construction of the universal -manifold of a locally real analytic singular foliation of [28, 30]. In particular, it makes sense for the universal Lie -algebroid of every singular foliation, without any additional assumption, and for Androulidakis-Zambon singular Lie algebroids. Also, to any ideal preserved by the anchor map of a Lie-Rinehart algebra , we associate a homotopy equivalence class of negatively graded Lie -algebroids over complexes computing . Several explicit examples are given.
Introduction
The recent surge of studies about Lie -algebras or Lie -groups, their morphisms and their -oids equivalent (i.e. Lie -algebroids [6, 38, 39] and “higher groupoids” [40]) is usually justified by their use in various fields of theoretical physics and mathematics. Lie -algebras or -oids appear often where, at first look, they do not seem to be part of the story, but end up to be needed to answer natural questions, in particular questions where no higher-structure concept seems a priori involved. Among examples of such a situation, let us cite deformation quantization of Poisson manifolds [25] and many recent developments of BV operator theory, e.g. [5], deformations of coisotropic submanifolds [7], integration problems of Lie algebroids by stacky-groupoids [34], complex submanifolds and Atiyah classes [20, 10, 29]. The list could continue.
For instance, in [28]-[30], it is proven that “behind” most singular foliations there is a natural homotopy class of Lie -algebroids, called universal Lie -algebroid of , and that the latter answers natural basic questions about the existence of “good” generators and relations for a singular foliation. The present article is mainly an algebraization of [28], algebraization that allows to enlarge widely the classes of examples. More precisely, our Theorems 2.1 and 2.4 are similar to the main theorems Theorem 2.8. and Theorem 2.9 in [28]:
-
(1)
Theorem 2.1 equips any free -resolution of a Lie-Rinehart algebra with a Lie -algebroid structure (Theorem 2.8. in [28] was a statement for geometric resolutions of locally real analytic singular foliation on an open subset with compact closure). This is a sort of homotopy transfer theorem, except that no existing homotopy transfer theorem applies in the context of generic -modules (for instance, [6] deals only with projective -modules). The difficulty is that we cannot apply the explicit transfer formulas that appear in the homological perturbation lemma because there is in general no -linear section of to its projective resolutions.
-
(2)
Theorem 2.4 states that any Lie -algebroid structure that terminates in comes equipped with a unique up to homotopy Lie -algebroid morphism to any structure as in the first item (Theorem 2.8. in [28] was a similar statement for Lie -algebroids whose anchor takes values in a given singular foliation).
As in [28], an immediate corollary of the result is that any two Lie -algebroids as in the first item are homotopy equivalent, defining therefore a class canonically associated to the Lie-Rinehart algebra, that deserve in view of the second item to be called “universal”.
However:
-
(1)
While [28] dealt with Lie -algebroids over projective resolutions of finite length and finite dimension, we work here with Lie -algebroids over any free resolution -even those of infinite length and of infinite dimension in every degree.
-
(2)
In particular, since we will work in a context where taking twice the dual does not bring back the initial space, we can not work with -manifolds (those being the “dual” of Lie -algebroids): it is much complicated to deal with morphisms and homotopies.
By doing so, several limitations of [28] are overcome. While [28] only applied to singular foliations which were algebraic or locally real analytic on a relatively compact open subset, the present article associates a natural homotopy class of Lie -algebroids to any Lie-Rinehart algebra, and in particular
-
a)
to any singular foliation on a smooth manifold, (finitely generated or not). This construction still works with singular foliations in the sense of Stefan-Sussmann for instance,
-
b)
to any affine variety, to which we associate its Lie-Rinehart algebra of vector fields), and more generally to derivations of any commutative algebra,
-
c)
to singular Lie algebroids in the sense of Androulidakis and Zambon [3],
-
d)
to unexpected various contexts, e.g. Poisson vector fields of a Poisson manifold, seen as a Lie-Rinehart algebra over Casimir functions, or symmetries of a singular foliation, seen as a Lie-Rinehart algebra over functions constant on the leaves.
These Lie -algebroids are constructed on -free resolutions of the initial Lie-Rinehart algebra over . They are acyclic universal in some sense, and they also are in particular unique up to homotopy equivalence. Hence the title.
A similar algebraization of the main results of [28], using semi-models category, appeared recently in Yaël Frégier and Rigel A. Juarez-Ojeda [14]. There are strong similarities between our results and theirs, but morphisms and homotopies in [14] do not match ours. It is highly possible, however, that Theorem 2.1 could be recovered using their results. Luca Vitagliano [36] also constructed Lie -algebra structures out of regular foliations, which are of course a particular case of Lie-Rinehart algebra. These constructions do not have the same purposes. For regular foliations, our Lie -algebroid structure is trivial in the sense that it is a Lie algebroid, while his structures become trivial when a good transverse submanifold exists. Lars Kjeseth [23, 22] also has a notion of resolutions of Lie-Rinehart algebras. But his construction is more in the spirit of Koszul-Tate resolution: Definition 1. in [23] defines Lie-Rinehart algebras resolutions as resolutions of their Chevalley-Eilenberg coalgebra, not of the Lie-Rinehart algebra itself as a module. It answers a different category of questions, related to BRST and the search of cohomological model for Lie-Rinehart algebra cohomology. For instance, the construction in [22] for an affine variety and its normal bundle does not match the constructions of Section 3.3 and, in our opinion, are of independent interest.
Our construction admits an important consequence. For any ideal , and any -free resolution of a Lie-Rinehart algebra over , the tensor product computes . It is easy to check that if is a Lie-Rinehart ideal (i.e. if ), then the universal Lie -algebroid structure that we constructed goes to the quotient to (which is a complex computing ). Also, two different universal Lie -algebroid structures on two different resolutions being homotopy equivalent, they lead to homotopy equivalent Lie -algebroid structures on two complexes computing . To a Lie-Rinehart ideal is therefore associated a homotopy equivalence class of Lie -algebroids on such complexes.
This article is organized as follows: In Section 1, we fix notations and review definitions, examples, and give main properties of Lie-Rinehart algebras. Afterwards, we present the concept of Lie -algebroids, their morphisms, and homotopies of those. In Section 2, we state and prove the main results of this paper, i.e. the equivalence of categories between Lie-Rinehart algebras and homotopy classes of free acyclic Lie -algebroids, which justifies the name universal Lie -algebroid of a Lie-Rinehart algebra. Section 3 is devoted to a precise description of the universal Lie -algebroids of several Lie-Rinehart algebras. The complexity reached by the higher brackets in these examples should convince the reader that it is not a trivial structure, even for relatively simple Lie-Rinehart algebras.
Acknowledgements
We would like to thank the CNRS MITI 80Prime project GRANUM, and the Institut Henri Poincaré for hosting us in november 2021. We thank the referee for pointing relevant references and a careful reading. We acknowledge discussion with S. Lavau at early stages of the project. We would like to thank C. Ospel, P. Vanhaecke and V. Salnikov for giving the possibility to present our results at the “Rencontre Poisson à La Rochelle, 21-22 October 2021”. Last, R. Louis would like to express sincere gratitude to Université d’État d’Haïti and more precisely the department of mathematics of École Normale Supérieure (ENS), for giving a golden opportunity to meet mathematics. He also would like to acknowledge the full financial support for this Phd from Région Grand Est.
Convention.
Throughout this article, is a commutative unital algebra over a field of characteristic zero, and stands for its Lie algebra of -linear derivations. Also, stands for the derivation applied to .
An other important convention is that for an algebra over , and an -module, we will denote by the symmetric powers over the domain and its symmetric powers over .
1. Lie-Rinehart algebras and Lie -algebroids
Except for Remark 1.3, this section is essentially a review of the literature on the subject, see, e.g. [18, 19].
1.1. Lie-Rinehart algebras and their morphisms
1.1.1. Definition of Lie-Rinehart algebras
Definition 1.1.
A Lie-Rinehart algebra over is a a triple with an -module, a Lie -algebra bracket on , and a -linear Lie algebra morphism called anchor map, satisfying the the so-called Leibniz identity:
A Lie-Rinehart algebra is said to be a Lie algebroid if is a projective -module.
Definition 1.2.
Given Lie-Rinehart algebras and over algebras and respectively. A Lie-Rinehart algebra morphism over an algebra morphism is a Lie algebra morphism such that for every and :
-
(1)
-
(2)
.
When and , we say that is a Lie-Rinehart algebra morphism over .
Remark 1.3.
Let us recall some basic constructions for Lie-Rinehart algebras.
-
(1)
Restriction. Consider a Lie-Rinehart algebra over . For every Lie-Rinehart ideal , i.e. any ideal such that
the quotient space inherits a natural Lie-Rinehart algebra structure over . We call this Lie-Rinehart algebra the restriction w.r.t the Lie-Rinehart ideal . In the context of affine varieties, when is the ideal of functions vanishing on an affine subvariety , we shall denote by .
-
(2)
Localization. Assume that is unital. Let be a multiplicative subset with no zero divisor which contains the unit element. The localization module comes equipped with a natural structure of Lie-Rinehart algebra over the localization algebra . The localization map is a Lie-Rinehart algebra morphism over the localization map . Since localization exists, the notion of sheaf of Lie-Rinehart algebras [35] over a projective variety, or a scheme, makes sense.
-
(3)
Algebra extension. Assume that the algebra is unital and has no zero divisor, and let be its field of fraction. For any subalgebra with such that is for any valued in derivations of that preserves , there is natural Lie-Rinehart algebra structure over on the space .
-
(4)
Blow-up at the origin. Let us consider a particular case of the previous construction, when is the algebra . If the anchor map of a Lie-Rinehart algebra over takes values in vector fields on vanishing at the origin, then for all , the polynomial algebra generated by satisfies the previous condition, and comes equipped with a Lie-Rinehart algebra. Geometrically, this operation corresponds to taking the blow-up of at the origin, then looking at the -th natural chart on this blow-up: are the polynomial functions on . The family (for ) is therefore an atlas for a sheaf of Lie-Rinehart algebras (in the sense of [35]) on the blow-up of at the origin, referred to as the blow-up of at the origin.
1.1.2. Geometric and Algebraic Examples
Below is an ordered list of examples to which we intend to apply our results: Vector fields vanishing on subsets of a vector space (Example 1.5), Lie algebroids (Example 1.6), cohomology in degree of a Lie -algebroid (Example 1.7), singular foliations, or non-finitely generated generalizations of those (Example 1.8) are Lie-Rinehart algebras over the algebra of functions on a manifold . We also give an example of a Lie-Rinehart algebra over Casimir functions (Example 1.11).
Example 1.4.
For every commutative -algebra , the Lie algebra of derivations of a commutative algebra is a Lie-Rinehart algebra over , with the identity as an anchor map. Vector fields on a smooth or Stein manifold or an affine variety are therefore instances of Lie-Rinehart algebras over their respective natural algebras of functions.
Example 1.5.
Let a Lie-Rinehart algebra over an algebra and be an ideal. The sub-module is a sub-Lie-Rinehart algebra of and its anchor is given by the restriction of over . It follows easily from
Example 1.6.
Let be a smooth manifold. By Serre-Swan Theorem, Lie algebroids (in the sense of [31]) over are precisely Lie-Rinehart algebras over of the form where is a vector bundle over and is a vector bundle morphism.
Example 1.7.
(see section 1.2.2) For every Lie -algebroid over an algebra , the quotient space comes equipped with a natural Lie-Rinehart algebra over , that we call the basic Lie-Rinehart algebra of .
Example 1.8.
There are several manner to define singular foliations on a manifold . All these definitions have in common to define sub-Lie-Rinehart algebras of the Lie-Rinehart algebra of vector fields on (or , i.e. compactly supported vector fields on ). With this generality, unfortunately, there are no good definition of leaves: as a consequence, several assumptions are generally made on , and singular foliations are usually defined as sub-Lie-Rinehart algebras of (or ) satisfying one of the conditions below:
-
(1)
“singular foliation admitting leaves”: there exists a partition of into submanifolds called leaves such that for all , the image of the evaluation map is the tangent space of the leaf through (when coincides with the space of vector fields tangent to all leaves at all points, we shall speak of a “Stefan-Sussman singular foliation”)
-
(2)
“self-preserving singular foliations”: the flow of vector fields in , whenever defined, preserves ,
- (3)
-
(4)
“finitely generated singular foliations”: when is finitely generated over and closed under Lie bracket.
It is known that Condition above implies Condition , for . The converse implications are not true in general. See [30] for an overview of the matter.
Example 1.9.
Singular subalgebroids of a Lie algebroid , i.e. submodules of stable under Lie bracket, are examples of Lie-Rinehart algebras: their anchors and brackets are the restrictions of the anchors and brackets of . Locally finitely generated ones are studied in [42, 3, 41]. In particular, sections of a Lie algebroid valued in the kernel of the anchor map form a Lie-Rinehart algebra for which the anchor map is zero.
Example 1.10.
For a singular foliation (in any one of the four senses explained Example 1.8) on a manifold , consider (i.e. infinitesimal symmetries of ) and
(that can be thought of as functions constant along the leaves of ). The quotient is Lie-Rinehart algebra over .
Example 1.11.
Let be a Poisson manifold. We define to be the first Poisson cohomology of and to be the algebra of Casimir functions. The bracket of vector fields makes a Lie-Rinehart algebra over .
1.2. Lie -algebroids and their morphisms
Lie -algebras are well-known to be coderivations of degree squaring to of the graded symmetric algebra . For Lie -algebroids, the situation is more involved, because the -ary bracket is not -linear. In the finite dimensional case [38], rather than seeing it as a coderivation of the symmetric algebra, it is usual to see it as a derivation of the symmetric algebra of the dual, i.e. as a -manifold. The duality “finite rank Lie -algebroids” “Q-manifolds” is especially efficient to deal with morphisms. In the lines above, we present a co-derivation version of Lie -algebroids, which is subtle, but gives a decent description of morphisms and their homotopies as co-algebra morphisms.
1.2.1. Graded symmetric algebras
Let us fix the sign conventions and recall the definition of (negatively-graded) Lie -algebroids. For is a graded- module, we denote by the degree of a homogeneous element .
-
(1)
We denote by and call graded symmetric algebra of over the quotient of the tensor algebra over of , i.e.
by the ideal generated by , with arbitrary homogeneous elements of . We denote by the product in .
-
(2)
Similarly, we denote by and call graded symmetric algebra of over the field the quotient of the tensor algebra (over ) of , i.e.
by the ideal generated by , with arbitrary homogeneous elements of . We denote by or the product in .
The algebras and come equipped with two different “degrees” that must not be confused.
-
(1)
We define the degree of or by
for any homogeneous . With respect to this degree, and are graded commutative algebras.
-
(2)
The arity of or is defined to be . We have and where and stand for the -module of elements of arity and the -vector space of elements of arity , respectively.
Convention 1.12.
For a graded -module, elements of arity and degree in (resp. ) shall be denoted by (resp. ).
For any homogeneous elements and a permutation of , the Koszul sign is defined by:
We often write for .
For , a -shuffle is a permutation such that and , and the set of all -shuffles is denoted by . Moreover, for and two homogeneous morphisms of -graded -modules, then stands for the following morphism:
Lemma 1.13.
Both and admit natural co-commutative co-unital co-algebra structures with respect to the deconcatenation defined by:
for every .
1.2.2. Lie -algebroids as co-derivations of graded symmetric algebras
Lie -algebroids over manifolds were introduced (explicitly or implicitly) by various authors, e.g. [33], [39], and [43]. We refer to Giuseppe Bonavolontà and Norbert Poncin for a complete overview of the matter [4]. Also, [23, 37] extend theses notions to the Lie-Rinehart algebra setting.
Definition 1.14.
A negatively graded Lie -algebroid over is a collection of projective -modules, equipped with:
-
(1)
a collection of -linear maps of degree called -ary brackets
-
(2)
a -linear map called anchor map
satisfying the following axioms :
-
the higher Jacobi identity:
(1) for all and homogeneous elements ;
-
for , the bracket is -linear, while for ,
where, by convention, is extended by zero on for all .
-
on .
-
is a morphism of brackets, i.e., for all .
Remark 1.15.
The third and the fourth axiom are consequences of item and if has no zero divisors.
Convention 1.16.
From now on, we will simply say “Lie -algebroid” for “negatively graded Lie -algebroid”.
It follows from Definition 1.14 that
is a complex of projective -modules. A Lie -algebroid is said to be acyclic if this complex has no cohomology in degree .
There is an equivalent way to define Lie -algebroids in term of co-derivations. We will use such a definition to deal with morphisms of Lie -algebroids.
Remark 1.17.
[21] Recall that a co-derivation of the symmetric algebra is entirely determined by the collection indexed by of maps called its -th Taylor coefficients:
(2) |
with pr being the projection onto the term of arity , i.e. .
Definition 1.18.
A co-algebra morphism or a co-derivation of the symmetric algebra are said to be of arity , if , for or , for .
1.2.3. Lie -algebroids and Richardon-Nijenhuis brackets
The space of -multilinear maps from to admits two gradings, the degree and the arity. Elements of arity and degree shall be, by definition, the space
The Richardon-Nihenhuis bracket [24]-[26]
is defined on homogeneous elements by
(3) |
Here (also denoted by ) is the interior product defined by
(4) |
The bracket is extended by bilinearity. It is classical that this bracket is a graded Lie algebra bracket.
Let us relate the bracket with co-derivations. For a given , and a given with , we denote by the unique co-derivation with Taylor coefficients . This co-derivation has degree , and the following Lemma is easily checked:
Lemma 1.19.
[21] For every of degrees as above, we have
We can now give an alternative description of Lie -algebroids in terms of co-derivation (extending the usual [38] correspondence between Lie -algebroids and -manifolds in the finite rank case). For a good understanding of the next Proposition, see notations of Taylor coefficients in Equation (2).
Proposition 1.20.
For a collection of projective -modules, there is a one-to-one correspondence between Lie -algebroid structures on and pairs made of co-derivations of degree which satisfies , and a -linear morphism, called the anchor, such that
-
(1)
for the -th Taylor coefficient of is -multilinear,
-
(2)
for all and , we have, ,
-
(3)
on ,
-
(4)
, for all .
The correspondence consists in assigning to a Lie -algebroid the co-derivation whose -th Taylor coefficient is the -ary bracket for all .
Proof.
The higher Jacobi identity is equivalent to
For all positive integer . The statement is then an immediate consequence of Lemma 1.19. ∎
Convention 1.21.
From now, when relevant, we will sometime denote an underlying structure of Lie -algebroid on by instead.
Remark 1.22.
Notice that does not induce a co-derivation on unless .
1.2.4. Universal Lie -algebroids of Lie-Rinehart algebras
Definition 1.23.
We say that a Lie -algebroid terminates at a Lie-Rinehart algebra when it is equipped with a -linear map , called hook, such that
Example 1.24.
Every Lie -algebroid terminates at the basic Lie-Rinehart algebra of Example 1.7, the projection being the hook.
Definition 1.25.
Let be a Lie-Rinehart algebra. A Lie -algebroid that terminates at through an hook is said to be universal for if
is a projective resolution of in the category of -modules.
In other words, a universal Lie -algebroid of a Lie-Rinehart algebra is a Lie -algebroid built on a projective resolution of as an -module, whose Lie-Rinehart algebra is .
1.2.5. Morphisms of Lie -algebroids and their homotopies
This section extends Section 3.4 of [28] to the infinite dimensional setting. Let and be graded -modules. A co-algebra morphism is completely determined by the collection indexed by of maps called its -th Taylor coefficients:
with pr being the projection onto the term of arity , i.e. .
The following Lemma is straightforward:
Lemma 1.26.
Let be a co-algebra morphism. The following conditions are equivalent:
-
(i)
For every , the -th Taylor coefficient of is -multilinear
-
(ii)
There exists an induced co-algebra morphism making the following diagram commutative :
We say that a co-algebra morphism is -multilinear when one of the equivalent conditions above is satisfied.
Now, let and be Lie -algebroids. Let and be their square-zero co-derivations of and respectively as in Proposition 1.20. Recall from [27] that Lie -algebra morphisms from to are defined to be co-algebra morphisms such that
(5) |
We will need two additional assumptions to turn a Lie -algebra morphism into a Lie -algebroid morphism:
Definition 1.27.
A Lie -algebroid morphism from a Lie -algebroid to a Lie -algebroid , is a Lie -morphism which is
-
(1)
-multilinear,
-
(2)
and satisfies on .
Above, is the chain map induced by (i.e. the restriction of to ).
When the Lie -algebroids and terminate at a given Lie-Rinehart algebra , we define morphisms of Lie -algebroids that terminate at as being Lie -algebroid morphisms that satisfy , where are their respective hooks. This property implies the second condition in Definition 1.27.
Example 1.28.
An -linear Lie algebroid morphism [31] is a Lie -algebroid morphism: the corresponding co-algebra morphsism is , for all .
Let us now define homotopy of Lie -algebroid morphisms. We start with a technical but important object:
Definition 1.29.
Let be a graded co-algebra morphism. A -co-derivation of degree on is a degree multilinear map which satisfies the (co)Leibniz identity:
(6) |
For a given co-algebra morphism , is entirely determined by the Taylor coefficient
which are called its -th Taylor coefficients:
(7) |
where for every subset , stands for the list and stands for the length of the list. This formula shows that the Taylor coefficients of are -multilinear if and only if induces a -co-derivation111defined as in (6) with instead of .
Proposition 1.30.
Let be a Lie -algebroid morphism. For a -multilinear -co-derivation of degree , is a -multilinear -co-derivation of degree .
Proof.
We first check that is a -co-derivation:
Subtracting a similar equation for and using (5), one obtains the -co-derivation property for . We now check that is -multilinear, for which it suffices to check that its Taylor coefficients are -multilinear by Lemma 1.26. Let be homogeneous elements. Assume (if we have more elements of degree the same reasoning holds). To verify -multilinearity it suffices to check that for all :
(8) | ||||
(9) |
Only the terms where the -ary bracket with a degree element on one-side and on the other side may forbid to go in front. There are two such terms:
where and stand for the list where and are missing respectively, and is the -th Taylor coefficient of . Since , in both terms appears, and these two terms add up to zero. ∎
Remark 1.31.
If the degree of is non-negative, then may not be -multilinear any more, since there may exist extra terms where the anchor map appears, e.g. terms of the form .
We can now define homotopies between Lie -algebroid morphisms, extending [28] from finite dimensional -manifolds to arbitrary Lie -algebroids.
Let be a vector space. Unless a topology on is chosen, the notion of -valued continuous or differentiable or smooth function on an interval does not make sense. However, we can always define the notion of a piecewise rational function as follows: we choose a finite increasing sequence of gluing points, and we require that for all the restriction of to is a finite sum of functions of the form with and a real rational function on which has no pole on . If and coincide at the gluing point , we say that is continuous. When is a space of linear maps between the vector spaces and , we shall say that a -valued map is a piecewise rational (continuous) if is a piecewise rational (continuous) -valued function for all .
Here is an important feature of such functions.
Lemma 1.32.
The derivative of a piecewise rational continuous function is defined at every point which is not a gluing point and is piecewise rational. Conversely, every piecewise rational functions admits a piecewise rational continuous primitive, unique up to a constant.
A family of co-algebra morphisms can now be defined to be piecewise rational continuous if its Taylor coefficients are piecewise rational continuous for all . For such a family , a family made of -co-derivations is said to be piecewise rational if all its Taylor coefficients are.
Remark 1.33.
In the above definitions, we do not assume the gluing points of the various Taylor coefficients or to be the same for all .
We now extend Definition 3.53 in [28] to the infinite rank case.
Definition 1.34.
Let and be Lie -algebroid morphisms from to . A homotopy between and is a pair consisting of:
-
(1)
a piecewise rational continuous path valued in Lie -algebroid morphisms between and satisfying the boundary condition:
-
(2)
a piecewise rational path , with a -co-derivations of degree from to , such that the following equation:
(10) holds for every where it is defined (that is, not a gluing point for the Taylor coefficients). More precisely, for every ,
(11) for all which is not a gluing point of the Taylor coefficient of for .
Definition 1.34 is justified by the following statement:
Proposition 1.35.
Let be a Lie -algebroid morphism from to . For all , let be a family -multilinear piecewise rational maps indexed by . Then,
-
(1)
There exists a unique piecewise rational continuous family of co-algebra morphisms such that
-
(a)
-
(b)
is a solution of the differential equation (10), where is the -co-derivation whose -th Taylor coefficient is for all .
-
(a)
-
(2)
Moreover, for all , is a Lie -algebroid homotopy between and .
Proof.
Let us show item (1). We claim that equation (10) is a differential equation that can be solved recursively. In arity zero, it reads,
(12) |
and
(13) |
is defined for all . Also, is an algebraic expression of , . But does not appear in the -th Taylor coefficient of by Equation (7). By Lemma 1.32, there exists a unique piecewise rational continuous solution such that . The construction of the Taylor coefficients of the co-algebra morphisms then goes by recursion. Recursion formulas also show that is unique.
Let us show 2), i.e. that is a a -multilinear chain map for all . The function given by
are differentiable w.r.t at all points except for a finitely many and are piecewise rational continuous. The map is a Lie -morphism because and , hence . By continuity, is constant over the interval . Since is a Lie -algebroid morphism, we have . Thus, and,
∎
Lemma 1.36.
Let and be Lie -Lie algebroids that terminate in through hooks and . Let be a homotopy. If is Lie -algebroid morphism that terminates at (i.e. ), then so is the -algebroid morphism for all .
Proof.
This is a direct consequence of Equation (13), since and are valued in the kernels of and respectively.
Last, -multilinearity of follows from the -multilinearity of , which is granted by Proposition 1.30. This completes the proof. ∎
Let us show that homotopy in the sense above defines an equivalence relation between Lie -morphisms. We have the following lemma.
Lemma 1.37.
A pair is a homotopy between Lie -algebroid morphisms and if and only if for all rational function, without poles on , the pair is a homotopy between and .
Proof.
Let be a rational function without poles on . A straightforward computation gives:
The last equation means that is a homotopy between and . The backward implication is obvious, it suffices to consider , and . ∎
Proposition 1.38.
Homotopy between Lie -morphisms is an equivalence relation. In addition, it is compatible with composition, that is, if are homotopic Lie -algebroid morphisms and are homotopic Lie -algebroid morphisms, then, so are their compositions and .
Proof.
We first show that this notion of homotopy is an equivalence relation. Let and be three Lie -morphisms of algebroids.
-
Reflexivity: The pair defines a homotopy between and .
-
Symmetry: Let be a homotopy between to . By applying Lemma 1.37 with , we obtain a homotopy between and via the pair .
-
Transitivity: Assume and and let be homotopy between and and let be a homotopy between and . By gluing and , respectively and we obtain a homotopy between and .
We then show it is compatible with composition. Let us denote by the homotopy between and , and the homotopy between and . We obtain,
Hence, and are homotopic via the pair which is easily checked to satisfy all axioms. This concludes the proof. ∎
We conclude this section with a lemma that will be useful in the sequel.
Lemma 1.39.
Let be a homotopy such that for all and for every , . Then the -th Taylor coefficient is constant on and the co-algebra morphism whose -th Taylor coefficient is for any and is a Lie -algebroid morphism.
Moreover, for a rational function with no pole on and such that , the pair is a homotopy between and .
Proof.
Since the -th Taylor coefficient of the -co-derivation depends only on for , we have by assumption for all . As a consequence is constant on . It follows from Proposition 1.35 that is a Lie -algebroid morphism since for every and
Let us prove the last part of the statement. By assumption, there exists such that for all , we have , so that and on . The function (resp. ) being piecewise rational continuous (resp. piecewise rational) on , the same holds for (resp. ) on . By gluing with a constant function (resp. with ), we see that all Taylor coefficients of (resp. ) are piecewise rational continuous (resp. piecewise rational) with finitely many gluing points. This completes the proof. ∎
Remark 1.40.
Lemma 1.39 explains how to glue infinitely many homotopies, at least when for a given , only finitely of them affects the -th Taylor coefficient.
2. Main results
2.1. The two main theorems about the universal Lie -algebroids of a Lie-Rinehart algebras
Let us now extend the main results of [28] from locally real analytic finitely generated singular foliations to arbitrary Lie-Rinehart algebras.
2.1.1. Existence Theorem 2.1
Here is our first main result, which states that universal Lie -algebroids over a given Lie-Rinehart algebra exist. We are convinced that it may be deduced using the methods of semi-models categories as in Theorem 4.2 in [14], but does not follow from a simple homotopy transfer argument. It extends Theorem 2.8 in [28].
Theorem 2.1.
Let be an algebra and be a Lie-Rinehart algebra over . Any resolution of by free -modules
(14) |
comes equipped with a Lie -algebroid structure whose unary bracket is and that terminates in through the hook .
Since any module admits free resolutions, Theorem 2.1 implies that:
Corollary 2.2.
Any Lie-Rinehart algebra admits a universal Lie -algebroid.
While proving Theorem 2.1, we will see that if can be equipped with a Lie algebroid bracket (i.e. a bracket whose Jacobiator is zero), then all -ary brackets of the universal Lie -algebroid structure may be chosen to be zero on :
Proposition 2.3.
Let be a free resolution of a Lie-Rinehart algebra . If admits a Lie algebroid bracket such that is a Lie-Rinehart morphism, then there exists a structure of universal Lie -algebroid of whose -ary bracket coincides with on and such that for every the -ary bracket vanishes on .
2.1.2. Universality Theorem 2.4 and corollaries
Here is our second main result. It is related to Proposition 2.1.4 in [14] (but morphisms are not the same), and extends Theorem 2.9 in [28].
Theorem 2.4.
Let be a Lie-Rinehart algebra over . Given,
-
a)
a Lie -algebroid that terminates in through the hook , and
-
b)
any Lie -algebroid universal for through the hook ,
then
-
(1)
there exists a morphism of Lie -algebroids from to over .
-
(2)
and any two such morphisms are homotopic.
Recall that a Lie -algebroid morphism as above is “a morphism of Lie -algebroid that terminates in ” (or “over ” for short) if , see Definition 1.27. Here is an immediate corollary of Theorem 2.4.
Corollary 2.5.
Any two universal Lie -algebroids of a given Lie-Rinehart algebra are homotopy equivalent. This homotopy equivalence, moreover, is unique up to homotopy.
We will prove that the morphism that appears in Theorem 2.4 can be made trivial upon choosing a “big enough” universal Lie -algebroid:
Proposition 2.6.
Let be a Lie-Rinehart algebra over . Given a Lie -algebroid structure that terminates in through a hook , then there exist a universal Lie -algebroid of through a hook such that
-
(1)
contains as a subcomplex,
-
(2)
the Lie -algebroid morphism from to announced in Theorem 2.4 can be chosen to be the inclusion map (i.e. a Lie -morphism where the only non-vanishing Taylor coefficient is the inclusion ).
The following Corollary follows immediately from Proposition 2.6:
Corollary 2.7.
Let be a Lie-Rinehart algebra over and be a Lie-Rinehart subalgebra of . Any universal Lie -algebroid of can be contained in a universal Lie -algebroid of .
2.1.3. Induced Lie -algebroids structures on .
Let be a Lie-Rinehart algebra over with anchor . We say that an ideal is a Lie-Rinehart ideal if for all . Since this assumption implies , the quotient space comes equipped with a natural Lie-Rinehart algebra structure over .
For an universal Lie -algebroid of , the quotient space comes equipped with an induced Lie -algebroid structure: the -ary brackets for go to quotient by linearity, while for , the -ary bracket goes to the quotient in view of the relation . Also, goes to the quotient to a Lie-Rinehart algebra morphism .
Definition 2.8.
Let be a Lie-Rinehart algebra over . For every Lie-Rinehart ideal , we call Lie -algebroid of the quotient Lie -algebroid , with a universal Lie -algebroid of .
Remark 2.9.
The complexes on which the Lie -algebroids of the ideal are defined compute by construction.
Moreover, for any two universal Lie -algebroids for , defined on the homotopy equivalences and , whose existence is granted by Corollary 2.5, go to the quotient and induce an homotopy equivalences between and . The following corollary is then an obvious consequence of Theorem 2.4.
Corollary 2.10.
Let be a Lie-Rinehart algebra over . Let be a Lie-Rinehart ideal. Then any two Lie -algebroids of are homotopy equivalent, and there is a distinguished class of homotopy equivalences between them.
Taking under account Remark 2.9, here is an alternative manner to restate this corollary.
Corollary 2.11.
Let be a Lie-Rinehart algebra over . Let be a Lie-Rinehart ideal. Then the complex computing comes equipped with a natural Lie -algebroid structure over , and any two such structures are homotopy equivalent in a unique up to homotopy manner.
When, in addition to being a Lie-Rinehart ideal, is a maximal ideal, then is a field and Lie -algebroids of are a homotopy equivalence class of Lie -algebras. In particular their common cohomologies, which is easily seen to be identified to comes equipped with a graded Lie algebra structure. In particular, is a Lie algebra, is a representation of this algebra, and the -ary bracket defines a class in the third Chevalley-Eilenberg cohomology of valued in . This class does not depend on any choice made in its construction by the previous corollaries and trivially extends the class called NMRLA-class in [28]. If it is not zero, then there is no Lie algebroid of rank equipped with a surjective Lie-Rinehart algebra morphism onto , where is the rank of as a module over . All these considerations can be obtained by repeating verbatim Section 4.5.1 in [28] (where non-trivial examples are given).
2.1.4. Justification of the title
Definition 2.12.
We denote by Lie--alg-oids/ the category where:
-
(1)
objects are Lie -algebroids over ,
-
(2)
arrows are homotopy equivalence classes of morphisms of Lie -algebroids over .
Theorem 2.4 means that universal Lie -algebroids over Lie-Rinehart algebra are terminal objects in the subcategory of Lie--alg-oids/ whose objects are Lie -algebroids that terminate in .
Let us re-state Corollary 2.5 differently. By associating to any Lie -algebroid its basic Lie-Rinehart algebra (see Example 1.7 and 1.24), one obtains therefore a natural functor:
-
•
from the category Lie--alg-oids/,
-
•
to the category of Lie-Rinehart algebras over .
Theorem 2.1 gives a right inverse of this functor. In particular, this functor becomes an equivalence of categories when restricted to homotopy equivalence classes of acyclic Lie -algebroids over , i.e:
Corollary 2.13.
Let be an unital commutative algebra. There is an equivalence of categories between:
-
(i)
Lie-Rinehart algebras over ,
-
(ii)
acyclic Lie -algebroids over .
The corollary justifies the title of the article.
Remark 2.14.
In the language of categories, Corollary 2.10 means that there exists a functor from Lie-Rinehart ideals of a Lie-Rinehart algebra over , to the category of Lie -algebroids, mapping a Lie-Rinehart ideal to an equivalence class of Lie -algebroids over .
2.2. An important bi-complex:
2.2.1. Description of
Let be an -module, and let and be complexes of projective -modules that terminates at :
(15) |
For every , the -th graded symmetric power of over is a projective -module, and comes with a natural grading induced by the grading on .
Definition 2.15.
Let . We call page number of and the bicomplex of -modules on the upper left quadrant defined by:
for and | (16) | ||||
for , | (17) |
together with the vertical differential defined for in any one of the two -modules (16) or (17) by
where acts as an -derivation on (and is 0 on ). The horizontal differential is given by
depending on whether is of type (16) with or the type (16) with . It is zero on elements of type (17). We denote by its associated total complex. When we shall write instead of .
The following diagram recapitulates the whole picture of :
|
(18) |
For later use, we spell out the meaning of being -closed.
Lemma 2.16.
An element in is -closed if and only if:
-
(1)
the component is valued in the kernel of ,
-
(2)
the following diagram commutes:
with being the component of in .
For , the second condition above also reads . Here is now our main technical result.
Proposition 2.17.
Let be a resolution of in the category of -modules. Then, for every ,
-
(1)
the cohomology of the complex for the total differential is zero in all degrees;
-
(2)
Moreover, a -closed element whose component on the “last column” of the diagram above is zero is the image through of some element whose two last components are also zero.
-
(3)
More generally, for all , for a -closed element of the form , one has and can be chosen in .
Proof.
Since is a projective -module for all , and is a resolution, all the lines of the above bicomplex are exact. This proves the first item. The second and the third are obtained by diagram chasing. ∎
We will need the consequence of the cone construction.
Lemma 2.18.
Let be an arbitrary complex of projective -module that terminates in a -module . There exists a projective resolution of , which contains as a sub-complex. Moreover, we can assume that admits a projective sub-module in in direct sum.
Proof.
Resolutions of an -modules are universal objects in the category of complexes of projective -modules. In particular, for every projective resolution of , there exist a (unique up to homotopy) chain map:
We apply the cone construction (see, e.g. [9], Section 1.5) to:
-
(1)
the complex
-
(2)
the direct sum of the complexes and namely,
-
(3)
the chain map obtained by mapping any to .
The differential is given by
(19) |
for all , . Since the chain given in item 3 is a quasi-isomorphism, its cone is an exact complex. We truncate the latter at degree without destroying its exactness by replacing the cone differential at degree as follows: . For a visual description, see Equation (20) below: the resolution of described in Lemma 2.18 is defined by:
(20) |
The proof of the exactness of this complex is left to the reader.
The henceforth defined complex is a resolution of , and obviously contains as a sub-chain complex of -modules. ∎
Let be a free resolution of and a subcomplex of projective -modules as in Lemma 2.18. We say that of the form preserves if is mapped by to for all possible indices. In such case, it defines by restriction to an element in the graded -module . For the sake of clarity, let us denote by and the respective differentials of the bi-complexes and and by , and , the horizontal differential resp. vertical differential, of their associated bi-complexes. Also, will stand for the restriction of to (a priori it is not valued in but in ).
Lemma 2.19.
Let be a free resolution of . Let be a subcomplex made of free sub--modules such that there exists a graded free -module such that .
-
(1)
For every , a -cocycle which preserves is the image through of some element which preserves if and only if its restriction is a -coboundary.
-
(2)
In particular, if the restriction of and to makes it a resolution of , then any -cocycle which preserves is the image through of some element which preserves .
Proof.
Let us decompose the element as with, for all , in . Assume is a -cocycle which preserves .
Let us prove one direction of item 1. If is the image through of some element which preserves , then , with . Thus, the restriction of is a -coboundary.
Conversely, let us assume that is a -coboundary, i.e. for some . Take any extension of (e.g. define to be as soon as one element in is applied to it). Then is zero on . We have to check that it is a -coboundary of a map with the same property. Put . By Proposition 2.17, item 1, there exists such that . The equation is equivalent to the datum of a collection of equations
(21) |
with, and for every . Since , we have that , (with the understanding that ). Using the exactness of the horizontal differential , there exists such that . We now change to and to by putting and . One can easily check that Equation (21) still holds under these changes, i.e.,
We can therefore choose such that . We then iterate this procedure, which allows us to choose such that and . By construction, preserves , while , and . The second item follows from the first one. ∎
2.2.2. Interpretation of
Let be a chain map, and let be its natural extension to a co-algebra morphism, namely:
We denote by and the differentials of arity on and induced by and . As in Definition 1.29 and Proposition 1.30, -co-derivations of a given arity form a complex when equipped with
Proposition 2.20.
For every , and be a chain map as above. The complex of -co-derivations of arity is isomorphic to the complex obtained from by crossing its “last column”, see diagram (18).
Proof.
The chain isomorphism consists in mapping a -co-derivation of arity and degree to its Taylor coefficient, which is an element of degree of . It is routine to check that this map is a chain map. ∎
Here is an other type of interpretation for involving the Richardson-Nijenhuis bracket.
Proposition 2.21.
[15] For , is the the bi-graded complex of exterior forms on and the differential of is .
2.3. Existence: The Lie -algebroid on a free -resolution
2.3.1. Proof of Theorem 2.1
In this section, we prove Theorem 2.1.
Consider a resolution of by free -modules: such resolutions always exist[9]. To start with, we define a binary bracket . The pair will obey to the axioms of the object that we now introduce.
Definition 2.22.
[30] An almost differential graded Lie algebroid of a Lie-Rinehart algebra is a complex
of projective -modules equipped with a graded symmetric degree -bilinear bracket such that:
-
(1)
satisfies the Leibniz identity with respect to the anchor ,
-
(2)
is degree -derivation of , i.e. for all :
-
(3)
is a morphism, i.e. for all
Lemma 2.23.
Every free resolution of a Lie-Rinehart algebra comes equipped with a binary bracket that makes it an almost differential graded Lie algebroid of .
Proof.
For all , let us denote by a family of generators of the free -module . By construction is a set of generators of . In particular, there exists elements , such that for given indices , the coefficient is zero except for finitely many indices , and satisfying the skew-symmetry condition together with
(22) |
We now define:
-
(1)
an anchor map by ) for all ,
-
(2)
a degree graded symmetric operation on as follows:
-
(a)
for all .
-
(b)
for all with or .
-
(c)
we extend to using -bilinearity and Leibniz identity with respect to the anchor .
-
(a)
By construction, satisfies the Leibniz identity with respect to the anchor . Also, on . The map defined for all homogeneous by
is a graded symmetric degree operation , and . Let us check that it is -bilinear, i.e. for all :
-
(1)
if , this quantity is zero in view of
-
(2)
if , one has
since ,
-
(3)
if with , it is obvious by -linearity of on the involved spaces.
As a consequence is a degree element in the total complex . By construction has no component on the last column. Since and also , the -bilinear operator is -closed in .
Proof (of Theorem 2.1).
Lemma 2.23 gives the existence of an almost differential graded Lie algebroid with differential and binary bracket . We have to construct now the higher brackets for .
Step 1: Construction of the 3-ary bracket . (Its construction being different from the one of the higher brackets, we put it apart). We first notice that the graded Jacobiator defined for all by
is -linear in each variable, hence is a degree element in . For degree reason, its component on the last column of diagram (18) is zero, i.e. it belongs to .
Let us check that it is -closed: for this purpose we have to check that both conditions in Lemma 2.16 hold:
-
(1)
Since is a morphism from to , and since satisfies the Jacobi identity, one has for all :
- (2)
Thus, . By Proposition 2.17, item 2, Jac is a -coboundary, and, more precisely, there exists an element with such that
(23) |
We choose the -ary bracket to be .
Step 2: Recursive construction of the -ary brackets for . Let us recapitulate: , and are constructed and the lowest arity terms of satisfy
-
(1)
(since ),
-
(2)
(since and define an almost Lie algebroid structure) .
-
(3)
by definition of , and because .
However, the following term of degree and arity may not be equal to zero:
(24) |
Let us check that this term is indeed a -multilinear map: For and , the only terms of where the anchor shows up are:
The terms containing the anchor map add up to zero. When there is more elements in , the computation follows the same line. Moreover, by graded Jacobi identity of the Richardson-Nijenhuis bracket:
The term of arity in the previous expression gives . Hence, by Proposition 2.21, is a -cocycle in the complex , whose components on the last column and the column are zero. It is therefore a coboundary by Proposition 2.17 item 3: we can continue a step further and define such that:
(25) |
We choose the -ary bracket to be . We now proceed by recursion. We assume that we have constructed all the -ary brackets such as :
(26) |
for every with . The ()-ary bracket is constructed as follows. First the operator is checked to be -linear as before. Now, we have
Since satisfies Equation (26) up to order , we obtain
where we used the graded Jacobi identity of the Nijenhuis-Richardson bracket in the last step. Therefore, , seen as an element in by Remark 2.21, is a cocycle and for degree reason it has no element on the last column, and the columns in 18. The third item of Proposition 2.17 gives the existence of an -ary bracket such as
This completes the proof. ∎
2.3.2. Proof of Proposition 2.3 and Proposition 2.6
Proof (of Proposition 2.3).
This is a consequence of Proposition 2.1 and the third item of the Proposition 2.17: If the component of Jac on the column is zero, we can choose with no component on the last column and in column (see Proposition 2.17), i.e. the restriction of to is zero. Then has no component on the last column, the column and the column . so has no component in the last column, and columns as well. Hence can be chosen with no component on column , and by the third item of Proposition 2.17. The proof continues by recursion. ∎
We finish this section with a proof of Proposition 2.6.
Proof (of Proposition 2.6).
We prove this Proposition in two steps.
-
(1)
Lemma 2.18 guarantees the existence a free resolution of the Lie-Rinehart algebra such that contains and such that there exists a graded free module with .
-
(2)
Let and be as in the proof of Lemma 2.19. We construct the -ary brackets on by extending the ones of in the following way:
-
(a)
We first construct an almost Lie algebroid bracket on that extends the -ary bracket of . Since the -ary bracket is determined by its value on a basis, the existence of a free module such that allows to construct on such that its restriction to is and such that it satisfies the Leibniz identity.
As in the proof of Theorem 2.1 (to be more precise: Lemma 2.23), we see that is -linear, hence belongs to and is a -cocycle. Since is a Lie -algebroid, its restriction to is zero. Lemma 2.19 allows to change to an -ary bracket with on . Hence defines a graded almost Lie algebroid bracket, whose restriction to is still .
-
(b)
Since is an extension of , its Jacobiator of the -ary bracket preserves . Also, its restriction is the Jacobiator of , and the latter is the -coboundary of in view of the higher Jacobi identity of . Since is a -cocycle, Lemma 2.19 assures that Jac is the image through of some element which preserves and whose restriction to is . The proof continues by recursion: at the -th step, we use Lemma 2.19 to construct an -ary bracket for that extends the -ary bracket of .
-
(a)
By construction, the inclusion map is a morphism for the -ary brackets for all . ∎
2.4. Universality: Proof of Theorem 2.4
Let be a Lie-Rinehart algebra. We consider a universal Lie -algebroid of : its existence is granted by Theorem 2.1, proved in Section 2.3.1. Let be an arbitrary Lie -algebroid that terminates in through a hook . Let (resp. be the co-derivations of (resp. of ) associated to the Lie -algebroid structures (resp. ) that terminate in .
Let us show that:
-
(1)
there is a Lie -algebroid morphism from to ,
-
(2)
Any such two Lie -morphisms are homotopic.
Altogether, these two points prove Theorem 2.4. The Taylor coefficients of the required Lie -algebroid morphisms and homotopies will be constructed by induction. These inductions rely on Lemmas 2.24 and 2.26 below.
Lemma 2.24.
Let be a co-algebra morphism such that
-
(1)
is -multilinear,
-
(2)
on ,
For every such that222 being a -co-derivation, its component of arity is zero for if only if its -th Taylor coefficient is zero for . for every , then the map given by:
-
(1)
is a -co-derivation of degree ,
-
(2)
is -multilinear,
-
(3)
and the induced -co-derivation satisfies:
Remark 2.25.
Proof.
A straightforward computation yields:
Now, preserves arity i.e. and so does . Taking into account the assumption for every , we obtain:
All the other terms disappear for arity reasons. Hence is a -co-derivation.
Let us prove that it is -linear. It suffices to check -linearity of . Let us choose homogeneous elements and let us assume that is the only term of degree : The proof in the case where there is more than one such an homogeneous element of degree is identical. We choose and we compute for some . The only terms in the previous expression which are maybe non-linear in are those for which the -ary brackets of a term containing with or appear (since and all other brackets are -linear). There are two such terms. The first one appears when we apply first, and then : this forces to appear, and the non-linear term is then:
(27) |
with the permutation that let goes in front and leave the remaining terms unchanged. There is a second term that appears when one applies first, then . Since it is a co-morphism, is the product of several terms among which only one is of degree , namely the term
Applying to this term yields the non-linear term
(28) |
where and are as in Proposition 1.30. Since , we see that the terms (27) and (28) containing an anchor add up to zero.
Let us check that is a chain map, in the sense that it satisfies item 3). Considering again , we have that , for all . Since , one has
By consequent, the -linear map satisfies item 3). ∎
Lemma 2.26.
Let be -linear Lie -algebroid morphisms and let . If for every , then
-
(1)
is a -co-derivation
-
(2)
is -multilinear
-
(3)
and the induced -co-derivation satisfies:
Remark 2.27.
Proof.
For all , one has:
Since has arity and for all , we obtain
(29) |
This proves the first item. Since both and are -multilinear, is -multilinear. Which proves the second item. Since, and are Lie -morphisms:
(30) |
By looking at the component of arity , one obtains, This proves the third item. ∎
Lemma 2.28.
Under the assumptions of Lemma 2.26, there exists
-
(1)
a Lie -morphism of algebroids
-
(2)
and a homotopy joining to , compatible with the hooks,
such that
-
(1)
the components of arity less or equal to of vanish,
-
(2)
for every .
Proof.
Let us consider . By assumption, for all , so that in view of Lemma 2.26
-
(1)
is a -co-derivation.
-
(2)
The restriction of to corresponds to a closed element of degree in .
Proposition 2.17 implies that there exists a degree -linear map , of arity , hence an element in , such that,
(31) |
We denote its extension to a -co-derivation of degree by . We now consider the following differential equation for :
(32) |
where is the unique -co-derivation of degree whose unique non zero Taylor coefficient is . The existence of a solution for the differential equation (32) is granted by Proposition 1.35. By considering the component of arity in Equation (32), we find
Hence:
Therefore, applying to the previous relation, one finds
This completes the proof. ∎
Proof (of Theorem 2.4).
Let us prove item 1. We construct the Taylor coefficients of the Lie -algebroid by recursion.
The Taylor coefficient of arity is obtained out of classical properties of projective resolutions of -modules. Given any complex which terminates in through , for every free resolution of , there exists a chain map as in Equation (15), and any two such chain maps are homotopic. We still denote by its extension to an arity co-morphism .
To construct the second Taylor coefficient, let us consider the map:
(33) |
This map is in fact -bililinear, i.e. belongs to , hence to , see Equation (18). Let us check that it is a -cocycle:
-
A.
If either one of the homogeneous elements or is not of degree , a straightforward computation gives:
-
B.
If both are of degree :
By Proposition 2.17 item 2), there exists , of degree , so that
(34) |
Since, is a chain map, Relation (34) can be rewritten in terms of and as follows
(37) |
The construction of the morphism announced in Theorem 2.4 is then done by recursion. The recursion assumption is that we have already defined a -multilinear co-morphism with
The co-morphism with Taylor coefficients and satisfies the recursion assumption for .
Assume now that we have a co-morphism that satisfies this assumption for some , and consider the map . Lemma 2.24 implies that it is a -multilinear -coderivation, and that it corresponds to a -closed element in . Since it has no component on the last column for degree reason, Proposition 2.17 implies that is a coboundary: That is to say that there is a -co-derivation (of arity and degree ) which can be seen as a map such that:
Consider now the co-morphism whose Taylor coefficients are those of in arity and in arity :
(38) |
This is easily seen to satisfy the recursion relation for . This concludes the recursion. The Taylor coefficients obtained by recursion define a Lie -algebroid which is compatible by construction with the hooks .
By continuing this procedure we construct a Lie -morphism from to . This proves the first item of Theorem 2.4.
Let us prove the second item in Theorem 2.4. Notice that in the proof of the existence of the Lie -morphism between and obtained in the first item, we made many choices, since we have chosen a coboundary at each step of the recursion.
Let be two Lie -morphisms between and . The arity component of the co-morphisms and restricted to are chain maps:
which are homotopy equivalent in the usual sense because is a projective resolution of : said differently, there exists a degree -linear map such that
(39) |
Let us consider the following differential equation:
(40) |
with being a -co-derivation of degree whose Taylor coefficient of arity is . This equation does admit solutions in view of Proposition 1.35.
By looking at the the component arity of Equation (40) on , one has:
Hence, is a solution such that . By construction, is homotopic to via the pair over , and its arity Taylor coefficient coincides with the Taylor coefficient of .
From there, the construction goes by recursion using Lemma 2.28. Indeed, this lemma allows to construct recursively a sequence of Lie -algebroids morphism and homotopies (with ) between and such that: is zero for and . By Lemma 2.28, all these homotopies are compatible with the hooks. These homotopies are glued in a homotopy such that for every , the components of arity of the Lie -algebroids morphism are constant and equal to for . By Lemma 1.39, these homotopies can be glued to a homotopy on . Explicitly, since maps to and by Lemma 1.39, the pair is a homotopy between and . This proves the second item of the Theorem 2.4.
∎
3. Examples of universal Lie -algebroid structures of a Lie-Rinehart algebra
3.1. New constructions from old ones
In this section, we explain how to construct universal Lie -algebroids of some Lie-Rinehart algebra which is derived from a second one through one of natural constructions as in Section 1.1 (localization, germification, restriction), when a universal Lie -algebroid of the latter is already known.
3.1.1. Localization
Localisation is an useful algebraic tool. When is an algebra of functions, it corresponds to study local properties of a space, or germs of functions.
Let be a Lie-Rinehart algebra over a unital algebra . Let be a multiplicative closed subset containing no zero divisor. We recall from item 2, Remark 1.3 that the localization of at comes equipped with a natural structure of Lie-Rinehart algebra over the localization algebra . Recall that for a homomorphism of -modules, there is a well-defined homomorphism of -modules,
that can be considered as a -module homomorphism
called the localization of .
Given a Lie -algebroid structure of . The triplet is a Lie -algebroid structure that terminates at through the hook where
-
(1)
;
-
(2)
The anchor map is defined by
for ;
-
(3)
, for all ;
-
(4)
The binary bracket is more complicated because of the anchor map: we set
for (with the understanding that on with );
-
(5)
.
One can check that these operations above are well-defined and for all the map is indeed a derivation on . The previously defined structure is also a Lie -algebroid that we call localization of the Lie -algebroid with respect to .
Proposition 3.1.
Let be a unital algebra and a multiplicative subset containing no zero divisor. The localization of a universal Lie -algebroid of a Lie-Rinehart algebra is a universal Lie -algebroid of .
Proof.
The object described above is also a Lie -algebroid terminating in . It is universal because localization preserves exact sequences [1]. ∎
3.1.2. Restriction
When is the ring of functions of an affine variety , to every subvariety corresponds its zero locus, which is an ideal . A Lie -algebroid or a Lie-Rinehart algebra over may not restrict to a Lie-Rinehart algebra over : it only does so when one can quotient all brackets by , which geometrically means that the anchor map takes values in vector fields tangent to . We can then “restrict”, i.e. replace by . This operation has already been defined in Section 2.1.3, and here is an immediate consequence of Corollary 2.11:
Proposition 3.2.
Let be Lie-Rinehart ideal, i.e. an ideal such that . The quotient of a universal Lie -algebroid of with respect to an ideal is a Lie -algebroid that terminates in . It is universal if and only if is exact, i.e. if .
3.1.3. Germification
Let be an affine variety and its coordinates ring. Denote by the germs of regular functions at . Since is a local ring, and since [16], where and is the localization w.r.t the complement of , Proposition 3.1 implies the following statement:
Proposition 3.3.
Let be an affine variety with functions . For every point and any Lie-Rinehart algebra over , the germ at of the universal Lie -algebroid of is the universal Lie -algebroid of the germ of at .
Here, the germ at of a Lie-Rinehart algebra or a Lie -algebroid is its localization w.r.t the complement of .
3.1.4. Sections vanishing on a codimension subvariety
Let be an arbitrary Lie-Rinehart algebra over . For any ideal , is also a Lie-Rinehart algebra (see Example 1.5). When are functions on a variety , are functions vanishing on a subvariety and is a -module of sections over , corresponds geometrically to sections vanishing along . It is not an easy task. In codimension , i.e. when is generated by one element, the construction can be done by hand.
Proposition 3.4.
Let be a Lie-Rinehart algebra over a commutative algebra . Let be a Lie -algebroid that terminates in through a hook . For any element , the -module is closed under the Lie bracket, so the triple is a Lie-Rinehart algebra over . A Lie -algebroid hooked in through can be defined as follows:
-
(1)
The brackets are given by
-
(a)
,
-
(b)
the -ary bracket:
(41) for all , with the understanding that on ,
-
(c)
for all ,
-
(a)
-
(2)
,
-
(3)
.
Proof.
We leave it to the reader. ∎
Proposition 3.5.
If is not a zero-divisor in , and is a universal Lie -algebroid of , then the Lie -structure described in the four items of 3.1 is the universal Lie -algebroid of .
Proof.
If is not a zero-divisor in (i.e if is an injective endomorphism of ), then the kernel of coincides with the kernel of , i.e. with the image of , so that is a resolution of . ∎
3.1.5. Algebra extension and blow-up
Recall that for a unital algebra with no zero divisor, derivations of induce derivations of its field of fractions .
Proposition 3.6.
Let be an unital algebra with no zero divisor, its field of fractions, and an algebra with . For every Lie-Rinehart algebra over whose anchor map takes values in derivations of preserving , then
-
(1)
any Lie -algebroid structure that terminates at extends for all to a Lie -algebroid structure on ,
-
(2)
and this extension is a Lie -algebroid that terminates at the Lie-Rinehart algebra .
Proof.
Since they are -linear, the hook , the anchor , and the brackets for are extended to -linear maps. Since the image of is the image of , it is made of derivations preserving , which is easily seen to allow an extension of to using the Leibniz identity. ∎
Remark 3.7.
Of course, the Lie -algebroid structure obtained on is not in general the universal Lie -algebroid of , because the complex may not be a resolution of (see Example 3.10).
Remark 3.8.
Since any module over a field is projective, any Lie-Rinehart algebra over a field is a Lie algebroid. If we choose therefore, the Lie-Rinehart algebra is a Lie algebroid, so is homotopy equivalent to any of its universal Lie -algebroid. Unless is a Lie algebroid itself, the Lie -algebroid in Proposition 3.6 will not be homotopy equivalent to a Lie -algebroid whose underlying complex is of length one, and is therefore not a universal Lie -algebroid of .
Example 3.9.
For the coordinate ring of , the blow-up of at the origin is covered by affine charts: in the -th affine chart , the coordinate ring is
By Remark 1.3, and Proposition 3.6 for any Lie-Rinehart algebra whose anchor map takes values in vector fields vanishing at , we obtain a Lie -algebroid of that we call blow-up of at in the chart . Proposition 3.6 then says that the blow-up at of the universal Lie -algebroid of , in each chart, is a Lie -algebroid that terminates in the blow-up of (as defined in remark 1.3). It may not be the universal one, see Example 3.10.
Example 3.10 (Universal Lie--algebroids and blow-up: a counter example).
Consider the polynomial function in variables . Let us consider the singular foliation as in Example 3.2.1. Its generators are , for .
Let us consider its blow-up in the chart . Geometrically speaking, is the -module generated by the blown-up vector fields , and . The vector fields , belong to the -module generated by the vector fields . Since they are independent, the singular foliation is a free -module. Its universal Lie -algebroid can be concentrated in degree , it is then a Lie algebroid.
On the other hand, the blow-up of the universal Lie -algebroid of is not homotopy equivalent to a Lie algebroid. Indeed, the pull-back Lie -algebroid verifies by construction that for and that for every in the inverse image of zero, and such a complex can not be homotopy equivalent333We are in fact proving that is not an exact functor in this case. to a complex of length .
This example tells us that the blow-up of the universal Lie -algebroid of a Lie-Rinehart algebra may not be the universal Lie -algebroid of its blow-up.
3.2. Universal Lie -algebroids of some singular foliations
3.2.1. Vector fields annihilating a Koszul function
This universal Lie -algebroid was already described in Section 3.7 of [28], where the brackets were simply checked to satisfy the higher Jacobi identities - with many computations left to the reader. Here, we give a theoretical explanation of the construction presented in [28].
Let be the algebra of all polynomials on . A function is said to be a Koszul polynomial, if the Koszul complex
(42) |
is exact in all degree, except in degree . By virtue of a theorem of Koszul [13], see [17] Theorem 16.5 , is Koszul if is a regular sequence.
From now on, we choose a Koszul function, and consider the singular foliation
(43) |
The Koszul complex (42) truncated of its degree term gives a free resolution of , with , , and .
Remark 3.11.
Exactness of the Koszul complex implies in particular that is generated by the vector fields:
(44) |
In [28], this resolution is equipped with a Lie -algebroid structure, whose brackets we now recall.
Proposition 3.12.
A universal Lie -algebroid of is given on the free resolution by defining the following -ary brackets:
(45) |
and the anchor map given for all by
(46) |
Above, for every multi-index of length , stands for the -vector field and . Also, is a multi-index obtained by concatenation of multi-indices . For every , is the signature of the permutation which brings to the first slots of . Last, for , we define .
To understand this structure, let us first define a sequence of degree graded symmetric polyderivations on (by convention, -vector fields are of degree ) by:
(47) |
We extend them to a graded poly-derivation of .
Lemma 3.13.
The poly-derivations are -multilinear and equip with a (graded symmetric) Poisson -algebra structure. Also, .
Proof.
For degree reason, for all and all . This implies the required -multilinearity. It is clear that the higher Jacobi identities hold since brackets of generators are elements in , and all brackets are zero when applied an element in . ∎
Proof (of Proposition 3.12).
The brackets introduced in Proposition 3.12 are modifications of the Poisson -algebra described in Lemma 3.13. By construction, when all arguments are generators of the form for some of cardinal . By -multilinearity, this implies when , or when and no argument is a bivector-field, or when and the argument is not a bivector field. As a consequence, all higher Jacobi identities hold when applied to -vector fields with .
Let us see what happens when one of the arguments is a bivector field, i.e. in the case where we deal with at least an element of degree . Let us assume that there is one such element, i.e. with . Then, in view of the higher Jacobi identity for the Poisson -brackets gives:
(48) | ||||
(49) | ||||
(50) | ||||
(51) | ||||
(52) |
In lines (48)-(49)-(50) above, we have for all the terms involved. This is not the case for (51)-(52). Indeed:
and
since . Hence the quantities in lines (52) and (51) add up, when we re-write them in term of the new brackets , to yield precisely the higher Jacobi identity for this new bracket. It is then not difficult to see this is still the case if there is more than one bivector field, by using many times the same computations. ∎
3.2.2. Restriction to of vector fields annihilating
We keep the convention and notations of the previous section. Let us consider the restriction of the Lie-Rinehart algebra to the zero-locus of a Koszul polynomial . Since all vector fields in are tangent to , this restriction is now a Lie-Rinehart algebra over (see Example 1.1).
Proposition 3.14.
Let be a Koszul Polynomial. The restriction of the universal Lie -algebroid of Proposition 3.12 to the zero-locus of is a universal Lie -algebroid of the Lie-Rinehart algebra .
Since the image of its anchor map are vector fields tangent to , it is clear that the universal Lie -algebroid of Proposition 3.12 restricts to . To prove Proposition 3.14, it suffices to check that the restriction to of the Koszul complex is still exact, except in degree . This is a simple lemma, whose proof is left to the reader.
Lemma 3.15.
The restriction to the zero locus of of the Koszul complex (42), namely the complex
is a free resolution of in the category of -modules.
3.2.3. Vector fields vanishing on subsets of a vector space
Let be the algebra of smooth or holomorphic or polynomial or formal functions on , and be an ideal. Then , i.e. vector fields of the form: , with , is a Lie-Rinehart algebra.
Remark 3.16.
Geometrically, when corresponds to functions vanishing on a sub-variety , must be interpreted as vector fields vanishing along .
Let us describe a Lie -algebroid that terminates at , then discuss when it is universal. Let be generators of . Consider the free graded algebra generated by variables of degree . The degree derivation squares to zero. The -module of elements degree in is made of all sums with . Consider the complex of free -modules
(53) |
Proposition 3.17.
The complex (53) comes equipped with a Lie -algebroid structure that terminates in through the anchor map given by for all , and .
Proof.
First, one defines a -linear Poisson--algebra structure on the free algebra generated by (in degree ) and (in degree ) and by:
(54) |
all other brackets of generators being equal to . Since the brackets of generators take values in , and since an -ary bracket where an element of appears is zero, this is easily seen to be a Poisson -structure. The general formula is
(55) |
where for every list , where is the Koszul sign, and where for a list containing , stands for the list from which the element is crossed out, as in Equation (45).
The -module generated by , i.e.
the complex (53) is easily seen to be stable under the brackets
for all , so that we can define on a sequence of brackets
by letting them coincide with the previous brackets on the generators, i.e. is given by Equation (55) for all . The brackets are then extended by derivation, -linearity or Leibniz identity with respect to the given anchor map, depending on the degree.
In particular,
for . For , on .
For , we still have on .
Let us verify that all required axioms are satisfied. For , Equation (55) specializes to:
which proves that the anchor map is a morphism when compared with the relation:
The higher Jacobi identities are checked on generators as follows:
-
(1)
When there are no degree generators, it follows from the higher Jacobi identities of the Poisson -structure (54) and the -multilinearity of all Lie -algebroid brackets involved.
-
(2)
When generators of degree are involved, the higher Jacobi identities are obtained by doing the same procedure as in the proof of Proposition 3.13, that is, we first consider the higher Jacobi identities for the Poisson -structure (54), and we put aside the terms where is applied to these degree generators. We then check that the latter terms are exactly the terms coming from an anchor map when the -ary bracket is applied to generators of degree and the -ary brackets of the remaining generators.
∎
Proposition 3.18.
When , and when is an affine variety defined by a regular sequence , then the Lie -algebroid described in Proposition 3.17 is the universal Lie -algebroid of the singular foliation of vector fields vanishing along .
Proof.
For a regular sequence , equipped with the derivation is a free -resolution of the ideal of functions vanishing along . Since is a flat -module, the sequence
(56) |
is a free -resolution of the singular foliation . The Lie -algebroid structure of Proposition 3.17 is therefore universal. ∎
Example 3.19.
As a special case of the Proposition 3.18, let us consider a complete intersection defined by one function, i.e. an affine variety whose ideal is generated by a regular polynomial . One has a free resolution of the space of vector fields vanishing on given as follows:
where is a degree variable, so that . The universal Lie -algebroid structure over that resolution is given on the set of generators by :
and for every . It is a Lie algebroid structure. Notice that this construction could be also be recovered using Section 3.1.4.
3.3. Universal Lie -algebroids associated to an affine variety
This section is mainly programmatic. It explains how we can attach the purely algebraic objects “universal Lie -algebroids” to the geometric objects “affine varieties”. Their relations will be studied in a future article. For any affine variety over , with ring of function , derivations of (i.e. vector fields on ) are a Lie-Rinehart algebra over denoted . Its universal Lie--algebroid can therefore be constructed.
Definition 3.20.
We call universal Lie -algebroid of an affine variety any Lie--algebroid associated to its Lie-Rinehart algebra of vector fields on .
Let us state a few results about the universal Lie -algebroid of an affine variety . For every point , let be the algebra of germs of functions at . The germ at of the Lie-Rinehart algebra of vector fields on is easily checked to coincide with the Lie-Rinehart algebra of derivations of .
Here is an immediate consequence of Proposition 3.3.
Proposition 3.21.
Let be an affine variety. For every , the germ at of the universal Lie -algebroid of is the universal Lie -algebroid of .
Let us choose . As stated in Section 3.1.2, the Lie -algebroid structure of over restricts at (i.e. goes to the quotient with respect to the ideal ) if and only if , (i.e is a Lie-Rinehart ideal). This is the case, in particular, if is an isolated singular point. In that case, we obtain a Lie -algebroid over . Since this quotient is the base field , we obtain in fact a Lie -algebra. By Section 3.1.2 again, it is a Lie -algebra on . We call it the Lie -algebra of the isolated singular point . To describe this structure, let us start with the following Lemma.
Lemma 3.22.
Let be a point of an affine variety . The universal Lie -algebroid of can be constructed on a resolution , with free -modules of finite rank for all , which is minimal in the sense that for all .
Proof.
Since Noetherian property is stable by passage to localization, the ring is a Noetherian local ring. Proposition 8.2 in [32] assures that admits a free minimal resolution by free finitely generated -modules. Since is a local ring with maximal idea , we can assume that this resolution is minimal. In view of Theorem 2.1, there exists a Lie -algebroid structure over this resolution, and the latter is universal for . ∎
By Theorem 2.1, a resolution of as in Lemma 3.22 comes equipped with a universal Lie -algebroid structure for . The quotient with respect to is a Lie -algebra of the isolated singular point with trivial -ary bracket. Using Corollary 2.11 and its subsequent discussion, we can prove the next statement.
Proposition 3.23.
For any universal Lie -algebroid structure on a resolution of as in Lemma 3.22, the quotient with respect to the ideal is a representative of the Lie -algebra of the isolated singular point , with trivial -ary bracket, on a graded vector space canonically isomorphic to ( being a -module through evaluation at ).
In particular, its -ary bracket is a graded Lie bracket on which does not depend on any choice made in the construction, and its -ary bracket is a Chevalley-Eilenberg cocycle whose class is also canonical.
This discussion leads to the natural question:
Question. How is the geometry of an affine variety related to its universal Lie -algebroid?
This will be the topic of a forthcoming article.
References
- [1] Gathmann Andreas. Commutative Algebra. Class Notes TU Kaiserslautern 2013/14.
- [2] Iakovos Androulidakis and Marco Zambon. Smoothness of holonomy covers for singular foliations and essential isotropy. Mathematische Zeitschrift, 275(3):921–951, 2013.
- [3] Iakovos Androulidakis and Marco Zambon. Integration of Singular Algebroids, https://arxiv.org/abs/2008.07976. arXiv, 2018.
- [4] Giuseppe Bonavolontà and Norbert Poncin. On the category of Lie -algebroids. J. Geom. Phys., 73:70–90, 2013.
- [5] Ricardo Campos. BV formality. Adv. Math., 306:807–851, 2017.
- [6] Ricardo Campos. Homotopy equivalence of shifted cotangent bundles. J. Lie Theory, 29(3):629–646, 2019.
- [7] Alberto S. Cattaneo and Giovanni Felder. Coisotropic submanifolds in Poisson geometry and branes in the Poisson sigma model. Lett. Math. Phys., 69:157–175, 2004.
- [8] Dominique Cerveau. Distributions involutives singulières. Ann. Inst. Fourier (Grenoble), 29(3):xii, 261–294, 1979.
- [9] Weibel Charles. An introduction to homological algebra / Charles A. Weibel. Cambridge studies in advanced mathematics. Cambridge University Press, Cambridge New York (N. Y.) Melbourne, 2014.
- [10] Zhuo Chen, Mathieu Stiénon, and Ping Xu. From Atiyah classes to homotopy Leibniz algebras. Comm. Math. Phys., 341(1):309–349, 2016.
- [11] Pierre Dazord. Feuilletages à singularités. Indag. Math., 47:21–39, 1985.
- [12] Claire Debord. Holonomy Groupoids of Singular Foliations. J. Differential Geom., 58(3):467–500, 07 2001.
- [13] David Eisenbud. Commutative algebra. With a view toward algebraic geometry, volume 150. Berlin: Springer-Verlag, 1995.
- [14] Yaël Frégier and Rigel A. Juarez-Ojeda. Homotopy theory of singular foliations. arXiv 1811.03078, 2018.
- [15] Alfred Frölicher and Albert Nijenhuis. Theory of vector-valued differential forms. I: Derivations in the graded ring of differential forms. Nederl. Akad. Wet., Proc., Ser. A, 59:338–350, 351–359, 1956.
- [16] Robin Hartshorne. Algebraic Geometry. Graduate Texts in Mathematics, 52. Springer New York, New York, NY, 1st ed. 1977. edition, 1977.
- [17] Matsumura Hideyuki. Commutative ring theory / Hideyuki Matsumura,… ; translated by M. Reid. Cambridge studies in advanced mathematics. Cambridge University Press, Cambridge, [edition with corrections] edition, 1989.
- [18] Johannes Huebschmann. Lie-Rinehart algebras, Gerstenhaber algebras and Batalin-Vilkovisky algebras. Ann. Inst. Fourier (Grenoble), 48(2):425–440, 1998.
- [19] Johannes Huebschmann. Lie-Rinehart algebras, descent, and quantization, volume 43 of Fields Inst. Commun. Amer. Math. Soc., Providence, RI, 2004.
- [20] Maxim Kapranov. Rozansky-Witten invariants via Atiyah classes. Compositio Math., 115(1):71–113, 1999.
- [21] Christian Kassel. Quantum Groups, volume 155 of Graduate Texts in Mathematics. Springer, 2012.
- [22] Lars Kjeseth. A homotopy lie-rinehart resolution and classical brst cohomology. Homology, Homotopy and Applications, 3(8):165–192, 2001.
- [23] Lars Kjeseth. Homotopy rinehart cohomology of homotopy lie-rinehart pairs. Homology, Homotopy and Applications, 3(7):139–163, 2001.
- [24] Ivan Kolář, Peter W. Michor, and Jan Slovák. Natural operations in differential geometry. Springer-Verlag, Berlin, 1993.
- [25] Maxim Kontsevich. Deformation quantization of Poisson manifolds. Lett. Math. Phys., 66(3):157–216, 2003.
- [26] Yvette Kosmann-Schwarzbach. Derived Brackets. Letters in mathematical physics, 69(1):61–87, 2004.
- [27] Tom Lada and Jim Stasheff. Introduction to SH Lie algebras for physicists. International journal of theoretical physics, 32(7):1087–1103, 1993.
- [28] Camille Laurent-Gengoux, Sylvain Lavau, and Thomas Strobl. The universal Lie -algebroid of a singular foliation. Doc. Math., 25:1571–1652, 2020.
- [29] Camille Laurent-Gengoux, Mathieu Stiénon, and Ping Xu. Poincaré-Birkhoff-Witt isomorphisms and Kapranov dg-manioflds. Advances in Mathematics, 2021.
- [30] Sylvain Lavau. Lie -algebroids and singular foliations. Ph-D 2017.
- [31] Kirill C. H. Mackenzie. General theory of Lie groupoids and Lie algebroids, volume 213 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 2005.
- [32] Jon Peter May. Cohen-Macauley and regular local rings. On-line 06-2021, https://www.math.uchicago.edu/ may/MISC/RegularLocal.pdf.
- [33] Hisham Sati, Urs Schreiber, and Jim Stasheff. Twisted differential string and fivebrane structures. Comm. Math. Phys., 315(1):169–213, 2012.
- [34] Hsian-Hua Tseng and Chenchang Zhu. Integrating Lie algebroids via stacks. Compos. Math., 142(1):251–270, 2006.
- [35] Joel Villatoro. On sheaves of Lie-Rinehart algebras. Arxiv.org/pdf/2010.15463, 2020.
- [36] Luca Vitagliano. On the strong homotopy associative algebra of a foliation. Commun. Cont. Math., 2015.
- [37] Luca Vitagliano. Representations of homotopy Lie-Rinehart algebras. Math. Proc. Camb. Phil. Soc., 158:155–191, 2015.
- [38] Theodore Voronov. Higher derived brackets and homotopy algebras. J. Pure Appl. Algebra, 202(1-3):133–153, 2005.
- [39] Theodore Voronov. -manifolds and higher analogs of Lie algebroids. In XXIX Workshop on Geometric Methods in Physics, volume 1307 of AIP Conf. Proc., pages 191–202. Amer. Inst. Phys., Melville, NY, 2010.
- [40] Pavol Ševera and Michal Širaň. Integration of differential graded manifolds. Int. Math. Res. Not. IMRN, (20):6769–6814, 2020.
- [41] Marco Zambon. Holonomy transformations for Lie subalgebroids, arxiv/2103.10409. arXiv, 2018.
- [42] Marco Zambon and Iakovos Androulidakis. Singular subalgebroids, arxiv/1805.02480.pdf. arXiv, 2018.
- [43] Pavol Ševera. Some title containing the words “homotopy” and “symplectic”, e.g. this one, arxiv:0105080. arXiv, 2001.