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

Exploring the Structure of Higher Algebroids111This research was supported by the Polish National Science Center under the grant DEC-2012/06/A/ST1/00256.

Mikołaj Rotkiewicz222Institute of Mathematics, University of Warsaw (email: [email protected])
Abstract

The notion of a higher-order algebroid, as introduced by Jóźwikowski and Rotkiewicz in their work Higher-order analogs of Lie algebroids via vector bundle comorphisms (SIGMA, 2018), generalizes the concepts of a higher-order tangent bundle τMk:TkMM\tau^{k}_{M}:\mathrm{T}^{k}M\to M and a (Lie) algebroid. This idea is based on a (vector bundle) comorphism approach to (Lie) algebroids and the reduction procedure of homotopies from the level of Lie groupoids to that of Lie algebroids. In brief, an alternative description of a Lie algebroid (A,[,],)(A,[\cdot,\cdot],\sharp) is a vector bundle comorphism κ\kappa, defined as the dual of the Poisson map ε:TATA\varepsilon:\mathrm{T}^{\ast}A\rightarrow\mathrm{T}A^{\ast} associated with the Lie algebroid AA. The framework of comorphisms has proven to be a suitable language for describing higher-order analogues of Lie algebroids from the perspective of the role played by (Lie) algebroids in geometric mechanics. In this work, we uncover the classical algebraic structures underlying the somewhat mysterious description of higher-order algebroids through comorphisms. For the case k=2k=2, we establish a one-to-one correspondence between higher-order Lie algebroids and pairs consisting of a two-term representation (up to homotopy) of a Lie algebroid and a morphism to the adjoint representation of this algebroid.

MSC 2020:

58A20, 58A50, 17B66, 17B70

Keywords:

higher algebroids, representations up to homotopy, graded manifolds, graded bundles, VB-algebroids, Lie algebroids

1 Introduction

In [JR18] the notion of a higher algebroid was introduced, based on extensive studies of examples we would like to refer to as higher algebroids (HAs, for short) [JR13, JR15]. Our intuitive thinking was that a higher algebroid should represent a geometric and algebraic structure that generalizes higher-order tangent bundles in a similar manner as algebroids generalize tangent bundles. In the first order, the algebroid structure is defined on a vector bundle (VB, for short), with the most obvious example being the tangent bundle τM:TMM\tau_{M}:\mathrm{T}M\rightarrow M. In higher orders, τM\tau_{M} is replaced by the higher-order tangent bundle τMk:TkMM\tau^{k}_{M}:\mathrm{T}^{k}M\rightarrow M, which for k>1k>1 is no longer a vector bundle but a graded bundle, in the terminology introduced in [GR11]. It is referred to as an \mathbb{N}-graded manifold in [JR18]. In a graded bundle there exists a distinguished class of graded fiber coordinates, taking over linear coordinates, with transition functions represented as homogeneous polynomials. In a particular case of polynomials of degree one (linear maps), one gets vector bundles as a special case. From various perspectives discussed in [JR18], it became apparent that the structure of a higher algebroid should be defined on a graded bundle.

The most common way to motivate the concept of a Lie algebroid comes from the reduction of the tangent bundle T𝒢\mathrm{T}\mathcal{G} of a Lie groupoid 𝒢\mathcal{G}. As a geometric object, the Lie algebroid of 𝒢\mathcal{G} is the set 𝒜(𝒢):=TM𝒢α\mathcal{A}(\mathcal{G}):=\mathrm{T}_{M}\mathcal{G}^{\alpha}, consisting of tangent vectors in the direction of the source fibration 𝒢α\mathcal{G}^{\alpha} of 𝒢\mathcal{G} and based at MM – the base of 𝒢\mathcal{G}. The structure of the tangent bundle T𝒢\mathrm{T}\mathcal{G}, induces a certain structure on TM𝒢α\mathrm{T}_{M}\mathcal{G}^{\alpha}, leading to the notion of a Lie algebroid. Typically, the structure of a Lie algebroid is expressed by means of a bracket operation [,][\cdot,\cdot] on the space of sections of a vector bundle σ:AM\sigma:A\rightarrow M and a VB morphism :ATM\sharp:A\rightarrow\mathrm{T}M called the anchor map. However, it is obvious that this approach has no direct generalization to higher-order case, because there is no bracket operation on the space of sections of τMk:TkMM\tau^{k}_{M}:\mathrm{T}^{k}M\rightarrow M, since τMk\tau^{k}_{M} is not a vector bundle for k>1k>1 and, in particular, its sections cannot be added.

In light of the above, it is natural to consider the reduction TMk𝒢α\mathrm{T}^{k}_{M}\mathcal{G}^{\alpha} of the kthk^{\mathrm{th}}-order tangent bundle Tk𝒢\mathrm{T}^{k}\mathcal{G} as a prototype of a higher-order algebroid of order kk. The reduction map k:Tk𝒢α𝒜k(𝒢):=TMk𝒢α\mathcal{R}^{k}:\mathrm{T}^{k}\mathcal{G}^{\alpha}\rightarrow\mathcal{A}^{k}(\mathcal{G}):=\mathrm{T}^{k}_{M}\mathcal{G}^{\alpha} takes a kthk^{\mathrm{th}}-velocity [g]k[g]_{k} represented by a curve g:𝒢g:\mathbb{R}\rightarrow\mathcal{G}, lying in a single fiber of the foliation 𝒢α\mathcal{G}^{\alpha}, to the kthk^{\mathrm{th}}-velocity based at a point in MM. A natural problem arises: how to characterize the structure on 𝒜k(𝒢)\mathcal{A}^{k}(\mathcal{G}) inherited from the groupoid multiplication. In [JR15] we proposed an answer to this question by reducing the natural map κ𝒢k:TkT𝒢TTk𝒢\kappa^{k}_{\mathcal{G}}:\mathrm{T}^{k}\mathrm{T}\mathcal{G}\rightarrow\mathrm{T}\mathrm{T}^{k}\mathcal{G}.

In the first order, one reduces the canonical involution κ𝒢\kappa_{\mathcal{G}} which results in a relation κT𝒜(𝒢)×T𝒜(G)\kappa\subset\mathrm{T}\mathcal{A}(\mathcal{G})\times\mathrm{T}\mathcal{A}(G) and leads to an alternative definition of the structure of a Lie algebroid as a pair (A,κ)(A,\kappa) consisting of a vector bundle σ:AM\sigma:A\rightarrow M and a relation κTA×TA\kappa\subset\mathrm{T}A\times\mathrm{T}A of a special kind . This viewpoint on Lie algebroids was first introduced in [GU99]. It turns out that κ\kappa is the dual of the Poisson map ε:TATA\varepsilon:\mathrm{T}^{\ast}A\rightarrow\mathrm{T}A^{\ast} associated with the linear Poisson tensor on AA^{\ast}.333Linear Poisson structures on the dual bundle AA^{\ast} are in a one-to-one correspondence with Lie algebroid structures on AA. As ε\varepsilon is a VB morphism, the dual κ=ε\kappa=\varepsilon^{\ast} is a VB comorphism, see Definition 2.4. The comorphism approach to (Lie) algebroids is also very natural from the perspective of variational calculus. The relation κ\kappa was recognized as a ’tool’ in constructing admissible variations in geometric mechanics [GG08].

Based on the properties of the reduction of κ𝒢k\kappa^{k}_{\mathcal{G}} and its potential applications in variational calculus, we introduced higher algebroids in [JR18] as pairs (Ek,κk)(E^{k},\kappa^{k}) consisting of a graded bundle σk:EkM\sigma^{k}:E^{k}\rightarrow M of order kk, equipped with a vector bundle comorphism κk\kappa^{k}. This comorphism relates the vector bundles TkE1TkM\mathrm{T}^{k}E^{1}\rightarrow\mathrm{T}^{k}M and TEkEk\mathrm{T}E^{k}\rightarrow E^{k} and satisfies certain natural axioms.444In the comorphism approach, it is natural to consider generalizations of the notion of a Lie algebroid obtained by relaxing its axioms. In the literature on geometric mechanics, these generalizations are known as ’almost Lie’ algebroids (where the Jacobi identity is not assumed), ’skew’ algebroids (where neither the Jacobi identity nor the anchor-bracket compatibility is required), and ’general’ algebroids (where, in addition, the skew-symmetry of the bracket is not required). We recall from [JR18] a detailed formulation of these axioms in Definition 2.7.

The definition of HAs given in [JR18], which we consider to be very natural from many perspectives, also appears to be quite mysterious. The goal of the present work was to unveil the vector bundle morphisms, brackets, and other operations hidden within the comorphism description of HAs. A complete solution is achieved in the case of k=2k=2.

Our solution situates Lie HAs within the realm of representations up to homotopy (representations u.t.h., in short) of Lie algebroids, the concept introduced in [AC12]. The idea is to represent Lie algebroids using cochain complexes of vector bundles. Such a complex is given ’an action’ of a Lie algebroid represented by an AA-connection which is flat only ’up to homotopy’ governed by higher order homotopy operators. When the complex consists of only one term, this notion reduces to a genuine representation of a Lie algebroid on a vector bundle. An important example for us is the notion of the adjoint representation whose proper generalization from the field of Lie algebras to that of Lie algebroids is found within the framework of representations up to homotopy. As explained in [AC12], the adjoint representation of a Lie algebroid AA is manifested by ’an action up to homotopy’ on the two-term complex :ATM\sharp:A\rightarrow\mathrm{T}M, where \sharp is the anchor map. On the other hand, it was found in [GSM10] that 2-term representations u.t.h. have an elegant description by means of VB-algebroids — Lie algebroid objects in the category of vector bundles. It this correspondence the adjoint representation of a Lie algebroid AA is nothing more but the VB-algebroid structure on TA\mathrm{T}A – the tangent prolongation of the Lie algebroid AA. Our solution also recognizes this point of view.

Our results.

The main result is presented in Theorem 3.26 and Corollary 3.27 where we establish a one-to-one correspondence between higher algebroids of order two and morphisms between representations u.t.h. of Lie algebroids of a specific nature as presented in the diagram:

Order-two Lie higher algebroids (E2,κ2)(E^{2},\kappa^{2}) one-to-onecorrespondence Representations u.t.h. of a Lie algebroid AA on a two-term complex ACA\rightarrow C together with a morphism Φ\Phi (of a special form) to the adjoint representation ad\operatorname{ad}_{\nabla} of AA

where \nabla is a fixed linear connection on a vector bundle AMA\rightarrow M. On the left is a Lie HA structure defined on a graded bundle E2ME^{2}\rightarrow M characterised by a special type of relation denoted as κ2\kappa^{2}, which is a subset of T2E1×MTE2\mathrm{T}^{2}E^{1}\times_{M}\mathrm{T}E^{2}. In this correspondence, A=E1A=E^{1} is the reduction of the graded bundle E2E^{2} to degree 11, with the Lie algebroid structure inherited from κ2\kappa^{2}. Furthermore, the vector bundle CMC\to M is introduced as the core of E2E^{2}, as explained in Section 2. On the right-hand side, we have a representation u.t.h. of the Lie algebroid AA defined on a two-term cochain complex ACA\rightarrow C. Additionally, there is a morphism denoted as Φ\Phi that connects this representation to the adjoint representation ad\operatorname{ad}_{\nabla} of AA in the sense of [AC12, Definition 3.3], and further, Φ\Phi is of special type: the 1-form component of Φ\Phi vanishes and so Φ\Phi is a map of cochain complexes and, moreover, Φ\Phi is the identity on AA in degree 0.

Following the ideas of [GSM10, DJLO15], we found that such a morphism Φ\Phi corresponds to a VB-algebroid morphism to the adjoint representation of AA represented as the VB-algebroid TA\mathrm{T}A (see Corollary 3.27). This construction makes the choice of a linear connection \nabla unnecessary. In summary, order-two Lie HAs are characterised by VB-algebroid morphisms Ψ:DTA\Psi:D\to\mathrm{T}A from a VB-algebroid DD to the tangent prolongation of a Lie algebroid AA, such that Ψ\Psi is the identity on the underlying algebroid AA, and on the core bundle, which is also identified with the vector bundle AA.

These results are obtained in a few steps which we discuss below.

Map Θk\mathrm{\Theta}^{k}.

In any order kk, we discover a canonical morphism of graded bundles  denoted by Θk:A[k]Ek\mathrm{\Theta}^{k}:A^{[k]}\rightarrow E^{k} (see Definition 3.1), which is associated with any almost Lie kthk^{\mathrm{th}}-order algebroid555In the case when k=2k=2, the existence of a morphism Θ2\mathrm{\Theta}^{2} is already guaranteed by the weaker assumption that AA is a skew algebroid. (Ek,κk)(E^{k},\kappa^{k}). Here, AA is the almost Lie algebroid (AL algebroid, for short) (E1,κ1)(E^{1},\kappa^{1}) obtained from (Ek,κk)(E^{k},\kappa^{k}) by means of the reduction to order one, and (A[k],κ[k])(A^{[k]},\kappa^{[k]}) is the kthk^{\mathrm{th}}-order prolongation of AA – a graded bundle with the HA structure naturally induced from the AL algebroid structure on AA (see (2.34) and (2.36)). The existence of this map is of crucial importance as it allows to relate properties of an abstract higher algebroid with much better recognized HA (A[k],κ[k])(A^{[k]},\kappa^{[k]}) studied in details in [JR15]. We recall that if A=𝒜(𝒢)A=\mathcal{A}(\mathcal{G}) is the Lie algebroid of a Lie groupoid 𝒢\mathcal{G} then A[k]=𝒜k(𝒢)=TMk𝒢αA^{[k]}=\mathcal{A}^{k}(\mathcal{G})=\mathrm{T}^{k}_{M}\mathcal{G}^{\alpha} is the kthk^{\mathrm{th}}-order HA of 𝒢\mathcal{G}. We conjecture (in Conjecture 3.5) that if (Ek,κk)(E^{k},\kappa^{k}) is a Lie HA then the structure map Θk:A[k]Ek\mathrm{\Theta}^{k}:A^{[k]}\rightarrow E^{k} is a morphism of HAs. We were able to prove this in the case k=2k=2.

The structure of the graded bundle of a HA (E2,κ2(E^{2},\kappa^{2}).

In general, a graded bundle of order two is obtained from its components: the vector bundle E1E^{1} (the order-one reduction of E2E^{2}), and its core vector bundle, denoted by E2widehat\widehat{E^{2}}, by gluing transition functions that are homogeneous polynomials of degree 2. In what follows, the vector bundles E1E^{1} and E2𝐰𝐢𝐝𝐞𝐡𝐚𝐭\widehat{E^{2}}, are denoted by AA and CC, respectively.

With the help of the map Θ2\mathrm{\Theta}^{2} we can recover the graded bundle  E2ME^{2}\to M as the quotient of the graded bundle  A[2]×MC[2]A^{[2]}\times_{M}C_{[2]}, see Lemma 3.6. Here, A[2]A^{[2]} is the second-order prolongation of AA and C[2]C_{[2]} denotes the graded bundle of order 2 obtained from the vector bundle CC by assigning weight two to the linear functions on CC.

Structure maps of HAs.

By focusing solely on the graded bundle structure of κ2\kappa^{2} we encounter equations (2.21). Our objective is to attribute a geometrical interpretation of the local structure functions QiaQ^{a}_{i}, QijaQ^{a}_{ij}, Qij,kμQ^{\mu}_{ij,k}, etc. present in (2.21). It turned out that the functions Qij,kμQ^{\mu}_{ij,k} do not correspond to any geometric object, highlighting that such an interpretation is not always straightforward. However, when we combine Qij,kμQ^{\mu}_{ij,k} with QijkQ^{k}_{ij} there emerges a three-argument operation, denoted by δ\delta, on the space Γ(A)\Gamma(A) of sections of the vector bundle AA with values in Γ(C)\operatorname{\Gamma}(C), where CC is the core of E2E^{2}, see (3.20) and (3.22). The meaning of the other structure functions666We will use the symbol QQ^{\cdots}_{\cdots} to refer to the structure functions QiaQ^{a}_{i}, QijaQ^{a}_{ij}, etc. present in (2.21). QQ^{\cdots}_{\cdots} proved to be more straightforward. These include: (i) a skew-symmetric bracket [,][\cdot,\cdot] on Γ(A)\operatorname{\Gamma}(A) and a VB morphism :ATM\sharp:A\to\mathrm{T}M defining a skew algebroid structure on AA, (ii) a morphism of graded bundles (the second-order anchor map) 2:E2T2M\sharp^{2}:E^{2}\rightarrow\mathrm{T}^{2}M, which is the base of the comorphism κ2\kappa^{2}, (iii) a vector bundle morphisms :AC\partial:A\rightarrow C, (iv) a map :Γ(A)×Γ(C)Γ(C)\Box:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(C)\rightarrow\operatorname{\Gamma}(C), (v) a skew symmetric map β:Γ(A)×Γ(A)Γ(C)\beta:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(A)\rightarrow\operatorname{\Gamma}(C).

There is also another interesting structure map ψ:Γ(A)×Γ(A)𝔛(M)\psi:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(A)\to\mathfrak{X}(M) (see (3.33)), which becomes relevant when studying tensorial properties of the aforementioned structure maps. Moreover, the symmetric part of ψ\psi, denoted by ψsym{\psi}^{\mathrm{sym}}, together with the VB morphisms 1=\sharp^{1}=\sharp and 2widehat\widehat{\sharp^{2}} (the core of 2\sharp^{2}), allows us to recover the second-order anchor map 2\sharp^{2}, see Lemma 3.12 and Theorem 3.13. This resolves the problem of presenting axioms of a skew order-two HA entirely in terms of VB morphisms and VB differential operators.

Definitions of all these structure maps are given in Subsection 3.2. Most of them are obtained through algebroid lifts Γ(A)𝔛(E2)\operatorname{\Gamma}(A)\rightarrow\mathfrak{X}(E^{2}), ssαs\mapsto{s}^{\langle{{\alpha}}\rangle}, associated with the HA (E2,κ2)(E^{2},\kappa^{2}) and Lie brackets of vector fields on E2E^{2}. The definition of algebroid lifts, as seen in (2.29), relies on the characteristic property of a VB comorphism: unlike a typical VB morphism, it induces a map between the spaces of sections.

We also introduce a structure map ω:Γ(A)×Γ(A)×Γ(A)Γ(C)\omega:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(A)\rightarrow\operatorname{\Gamma}(C), being some modifications of the map δ\delta, see (3.25). While it carries the same information as δ\delta, it has better algebraic properties and helps to formulate our results more concisely. Additionally, we define maps ξ\xi, ε\varepsilon, ε0\varepsilon_{0}, ε1\varepsilon_{1} in Definition 3.10, which appear in the Leibniz-type formulas for above-mentioned structure maps (Theorem 3.13). The vanishing of these maps is also included as an axiom of AL HAs or Lie HAs, see Theorems 3.16 and 3.20. Moreover, for greater precision in formulating certain results, we found it useful to decompose some of these maps, such as δ\delta, ω\omega, and ψ\psi, into their symmetric and anti-symmetric parts.

Most of the structure maps mentioned above are multi-differential operators on certain vector bundles. We provide a detailed description of the tensorial properties of these structure maps and prove that a system of such maps allows to reconstruct the skew HA (E2,κ2)(E^{2},\kappa^{2}), see Theorem 3.13. This approach can also be extended to kthk^{\mathrm{th}}-order HAs for k>2k>2, as discussed in Remark 3.9.

In the next step, we characterize the axioms of an almost Lie HA (Theorem 3.16) and Lie HA (Theorem 3.20) in terms of the above mentioned structure maps. In other words, we formulate necessary and sufficient conditions that the structure maps β\beta, \square, ω\omega etc. should satisfy for the related HAs to be, respectively, almost Lie and Lie. Throughout our analysis, we heavily rely on Theorem 2.11, which provides characterizations of AL and Lie axioms for higher algebroids through algebroid lifts.

HAs over a point.

In case when the base MM is a point we find a complete description of order-two skew and Lie higher algebroids, see Theorem 3.15: An order-two skew HA over a point has to split, meaning E2=𝔤[1]×C[2]E^{2}=\mathfrak{g}_{[1]}\times C_{[2]} where 𝔤=E1\mathfrak{g}=E^{1} and C=E2widehatC=\widehat{E^{2}}. Furthermore, in the Lie case, there is a one-to-one correspondence between such Lie HAs and Lie module morphisms :𝔤C\partial:{\mathfrak{g}}\to{C} from the adjoint module of the Lie algebra 𝔤\mathfrak{g} to the 𝔤\mathfrak{g}-module CC.

Main result.

The reformulated axioms of Lie HAs and the description of order-two HAs over a point by means of representations of Lie algebras may suggest a relation between HAs and representations of Lie algebroids. Note however that there is no concept of the adjoint representation within the framework of representations on VBs. It is the setting of representations u.t.h. of Lie algebroids in which the correct generalization of the concept of the adjoint representation of a Lie algebra is possible. The construction of a Lie algebroid representation out of a Lie HA (E2,κ2)(E^{2},\kappa^{2}) imitates the construction of the adjoint representation given in [AC12]. It is obtained by means of the structure maps of a Lie HA mentioned earlier. There exist also an obvious map Φ\Phi between the complexes ACA\to C and ATMA\to\mathrm{T}M which, thanks to the properties of the structure maps of a Lie HA, turns out to be a morphism to the adjoint representation of AA. Conversely, if a representation u.t.h. of AA on a two-term complex of the form ACA\to C is given, and a morphism of complexes Φ:(AC)(ATM)\Phi:(A\to C)\rightarrow(A\to\mathrm{T}M) is given that also serves as a morphism of representations, then we can extract the structure maps from it and construct a skew HA structure on the graded bundle described in Lemma 3.6. It can be then verified that these maps satisfy the axioms of AL and Lie HA given in Theorems 3.16 and  3.20.

Examples.

Given a Lie algebroid AA, there are two natural morphisms to the adjoint representation of a Lie algebroid AA. One is the identity morphism Φ\Phi on the adjoint representation. The other one is obtained from the double of a vector bundle, described in [AC12], which is a representation of the Lie algebroid AA on a 2-term complex of the form EidEE\xrightarrow{\operatorname{id}}E. We illustrate HAs corresponding to these two cases in Examples 3.30 and 3.29, respectively.

Organization of the paper.

Section 2 begins by collecting notations and fundamental constructions concerning graded bundles, double vector bundles and VB-algebroids. We also introduce a functor, denoted by λ\lambda, which is a generalization of the linearisation functor discovered in [BGG16] and which is used in the definition of the morphism Θk\mathrm{\Theta}^{k} (Definition 3.1). We recall also basic definitions from [JR18, JR15] (VB comorphism, higher-order algebroid, prolongations of AL algebroids) and give a definition of algebroid lifts in a slightly different way than in [JR18, Definition 4.8], more convenient for computations which we perform in Section 4. Theorem 2.11 extends Proposition 4.9 from [JR18] to the AL case. In Lemma 2.14 we express the compatibility of algebroid lifts obtained by means of HAs κk\kappa^{k} and the reduction of κk\kappa^{k} to a lower weight. We also list a few canonical inclusions used in the paper and describe their relationships.

Section 3 is devoted to a detailed analysis of mathematical structures standing behind a comorphism κk\kappa^{k} that defines a HA structure. W begin with the definition and properties of the map Θk\mathrm{\Theta}^{k} , which connects an arbitrary HA (Ek,κk)(E^{k},\kappa^{k}) with the kthk^{\mathrm{th}}-order prolongation of its first-order reduction (E1,κ1)(E^{1},\kappa^{1}). We provide coordinate formulas for Θ2\mathrm{\Theta}^{2} and Θ3\mathrm{\Theta}^{3}, see Example 3.3.

From this point on, we focus solely on the case k=2k=2. In Lemma 3.6 we find an explicit construction of a graded bundle E2ME^{2}\to M which hosts an order-two HA (E2,κ2)(E^{2},\kappa^{2}). Subsequently, we introduce several canonical maps associated with this HA referring to them as "the structure maps of (E2,κk)(E^{2},\kappa^{k})". Most of these maps are differential operators defined on (the product of) the spaces of sections Γ(A)\operatorname{\Gamma}(A) or Γ(C)\operatorname{\Gamma}(C) with values in Γ(C)\operatorname{\Gamma}(C). The term "structure functions" is reserved to functions QiaQ^{a}_{i}, QijaQ^{a}_{ij}, etc. which are given in Example 2.8 as a local representation of a general order-two HA. These functions depend on the chosen coordinate system on the graded bundle E2E^{2}. Although we work in the case k=2k=2 we present analogs of the structure maps in any order, see Remark 3.9. In Theorem 3.13 we provide an equivalent description of skew, order-two HAs in terms of the aforementioned structure maps. In Theorem 3.15 we discuss the special case when the base MM of E2E^{2} is a single point (an order-two analog of a (Lie) algebra) and give a characterization of such structures. It turns out that the Lie condition is very rigorous and all such Lie HAs correspond to morphisms :CA\partial:C\to A of Lie algebra modules, where AA represents the adjoint module of the Lie algebra AA.

We subsequently examine the conditions in Definition 2.7 characterizing AL and Lie HA and translate them to the level of the structure maps, see Theorem 3.16 and Theorem 3.20. Moving forward, in Lemma 3.25 we recognize that data describing order-two Lie HAs gives rise to a representation u.t.h. of the Lie algebroid AA on the structure map :AC\partial:A\to C considered as a two-term complex of vector bundles and also induces a morphism Φ\Phi to the adjoint representation of AA. Remarkably, this data is also sufficient for recovering a HA structure (E2,κ2)(E^{2},\kappa^{2}), as demonstrated in Theorem 3.26. Furthermore, we formulate VB-algebroid version of this correspondence in Corollary 3.27 and illustrate the obtained relationship in Examples 3.30 and 3.29. We also briefly recall the correspondence between representations u.t.h. and VB-algebroids —- providing the necessary facts on this subject to demonstrate our results.

In Appendix 4 we give proofs for various results, including part (a) of Theorem 3.13, Theorem 3.15, Conjecture 3.5 in the case k=2k=2 and complete the proof of Theorem 3.16 , where more detailed calculations, including those in coordinates, are carried out. Some of these calculations are supported by additional lemmas. One can also find there a brief recollection on representations up to homotopy, guided by [AC12]. For more in-depth information, interested readers should refer to the existing literature [AC12, GSM10, BGV18, GSJLMM18].

Historical remarks.

The studies on HAs, as understood in this paper, were initiated by M. Jóźwikowski and the author of present manuscript in [JR13], and continued in [JR15, JR18]. Prior to this, higher-order analogues of Lie algebroids was the subject of [Vor10] by T. Voronov who proposed that such analogues should be QQ-manifolds of spacial kind generalizing Vaintrob approach to Lie algebroids [Vai97]. The most recent studies are due to A. Bruce, K. Grabowska and J. Grabowski [BGG16] whose idea was to imitates the canonical inclusion TkMTTk1M\mathrm{T}^{k}M\subset\mathrm{T}\mathrm{T}^{k-1}M on the abstract level of graded bundles  having TkM\mathrm{T}^{k}M as a prominent example of kthk^{\mathrm{th}}-order analogue of a Lie algebroid. As we pointed in [JR18], all these approaches lead to different mathematical objects. This distinctiveness is further evident from the classification of order-two (Lie) HAs given in this work.

Acknowledgments.

The author is grateful to Michał Jóźwikowski for his insightful comments on the organization and editing of this work, and for many fruitful discussions regarding the research.

2 Preliminaries

2.1 Graded bundles 

We shall review basic constructions associated with graded bundles  that will be used in the present work. For further details, we refer to [GR11] and additional works [BGR16, JR18, Vor02].

A fundamental example of a graded bundle is the kthk^{\mathrm{th}}-order tangent bundle τMk:TkMM\tau^{k}_{M}:\mathrm{T}^{k}M\rightarrow M. The elements of TkM\mathrm{T}^{k}M are kthk^{\mathrm{th}}-order tangency classes [γ]k[\gamma]_{k} of curves γ\gamma in MM.777[γ]k[\gamma]_{k} is also called the kk-velocity represented by the curve γ\gamma Then T1M=TM\mathrm{T}^{1}M=\mathrm{T}M is the tangent bundle of MM but for k>1k>1, τMk\tau^{k}_{M} is not a vector bundle; however, the fibers are still equipped with a special structure, namely, a natural action of the monoid of real numbers (,)(\mathbb{R},\cdot),

h:×TkMTkM,(s,[γ]k)[γs]kh:\mathbb{R}\times\mathrm{T}^{k}M\rightarrow\mathrm{T}^{k}M,\quad(s,[\gamma]_{k})\mapsto[\gamma_{s}]_{k}

where γs(t)=γ(st)\gamma_{s}(t)=\gamma(st). Thus, in the terminology of [GR11], TkM\mathrm{T}^{k}M is a homogeneity structure, i.e., a manifold equipped with a smooth action of (,)(\mathbb{R},\cdot). On the other hand, local coordinates (xa)(x^{a}) on MM induce adapted coordinates 888f(α)𝒞(TkM)f^{({\alpha})}\in\mathcal{C}^{\infty}(T^{k}M), given by f(α)([γ]k)=dkdtkt=0f(γ(t))f^{({\alpha})}([\gamma]_{k})=\frac{\mathrm{d}^{k}}{\mathrm{d}t^{k}}_{t=0}f(\gamma(t)), denotes the (α)({\alpha})-lift of f𝒞(M)f\in\mathcal{C}^{\infty}(M), see [Mor70]. Hence, xa,(α)=(xa)(α)x^{a,({\alpha})}=(x^{a})^{({\alpha})} denotes the (α)({\alpha})-lift of the coordinate function xax^{a}. (xa,(α))0αk(x^{a,({\alpha})})_{0\leq{\alpha}\leq k} on TkM\mathrm{T}^{k}M which are naturally graded by numbers 0,1,,k0,1,\ldots,k. On T2MT^{2}M they transform as

xa=xa(x),x˙a=xaxbx˙b,x¨a=xaxbx¨b+2xaxbxcx˙bx˙c,x^{a^{\prime}}=x^{a^{\prime}}(x),\quad\dot{x}^{a^{\prime}}=\frac{\partial x^{a^{\prime}}}{\partial x^{b}}\,\dot{x}^{b},\quad\ddot{x}^{a^{\prime}}=\frac{\partial x^{a^{\prime}}}{\partial x^{b}}\,\ddot{x}^{b}+\frac{\partial^{2}x^{a^{\prime}}}{\partial x^{b}\partial x^{c}}\dot{x}^{b}\dot{x}^{c},

where x˙a=(xa)(1)\dot{x}^{a}=(x^{a})^{(1)}, x¨a=(xa)(2)\ddot{x}^{a}=(x^{a})^{(2)}. In general, the gradation of coordinates leads to the concept of a graded bundle  i.e., a smooth fiber bundle σk:EkM\sigma^{k}:E^{k}\rightarrow M in which we are given a distinguished class of fiber coordinates, called graded coordinates. Each graded coordinate is assigned its weight and transition functions preserve this gradation. An important assumption is made that weights are non-negative integers. (The index kk in EkE^{k} indicates that all weights are k\leq k, in which case we say that the graded bundle σk\sigma^{k} is of order kk. Graded bundles of order 11 are nothing more than vector bundles.)

It has been shown that both the concept of a homogeneity structure and a graded bundle are equivalent [GR11]. A graded bundle associated with a homogeneity structure (E,h)(E,h) can be conveniently encoded by means of the weight vector field defined as Δ(p)=ddt|t=1ht(p)\Delta(p)=\frac{\mathrm{d}}{\mathrm{d}t}\big{|}_{t=1}h_{t}(p). In graded coordinates (xa,ywi)(x^{a},y^{i}_{w}) 999This notation means that (xa)(x^{a}) are functions defined (locally) on the base MM of the graded bundle h0:EMh_{0}:E\rightarrow M while (ywi)(y^{i}_{w}) are fiber coordinates in this bundle. Moreover, the (abundant) notation yi=ywiy^{i}=y^{i}_{w} indicates that the function yiy^{i} has weight ww, i.e., is a homogeneous function (with respect to hh) of weight ww. we have Δ=iwiywiywi\Delta=\sum_{i}w^{i}y^{i}_{w}\partial_{y^{i}_{w}}. A morphism ff between graded bundles EE and FF, colloquially described as a map preserving the gradation of coordinates, can be given a short, precise meaning as a smooth map f:EFf:E\rightarrow F such that the corresponding weight vector fields, ΔE\Delta_{E} and ΔF\Delta_{F}, are ff-related. Equivalently, this can be described as a smooth map intertwining the corresponding homogeneity structures, i.e., fhtE=htFff\circ h_{t}^{E}=h_{t}^{F}\circ f for every tt\in\mathbb{R}, where ht=h(t,)h_{t}=h(t,\cdot).

In this work, we frequently encounter multi-graded structures like TTkM\mathrm{T}\mathrm{T}^{k}M, TEk\mathrm{T}E^{k} (the tangent bundle of a graded bundle of order kk) or TkE\mathrm{T}^{k}E (kthk^{\mathrm{th}}-order tangent bundle of EE, where σ:EM\sigma:E\rightarrow M is a vector bundle). In all these examples, there are present two (compatible) graded bundle structures. Such structures can be described as (F;Δ1,Δ2)(F;\Delta_{1},\Delta_{2}) – a manifold FF equipped with two weight vector fields Δ1,Δ2\Delta_{1},\Delta_{2}, and the condition of compatibility can be expressed as [Δ1,Δ2]=0[\Delta_{1},\Delta_{2}]=0. Equivalently, the last condition can be stated as ht1hs2=hs2ht1h_{t}^{1}\circ h_{s}^{2}=h_{s}^{2}\circ h_{t}^{1} for any t,st,s\in\mathbb{R}, where (F,hi)(F,h^{i}) are the homogeneity structures with weight vector field Δi\Delta_{i}, for i=1,2i=1,2. Moreover, the bases of the graded bundles (F,Δi)(F,\Delta_{i}), where i=1,2i=1,2, carry induced graded bundle structures. In this paper, we shall mostly encounter the case when one of these graded bundle structures has order 1 (like in TEk\mathrm{T}E^{k} ot TkET^{k}E) and will refer to them as weighted vector bundles.101010Weighted structures, e.g. weighted algebroids, where intensively studied in [BGG15a, BGG16]. They can be presented as a diagram like

Fk\textstyle{F^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σk\scriptstyle{\sigma^{k}}πk\scriptstyle{\pi^{k}}F0\textstyle{F^{0}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π0\scriptstyle{\pi^{0}}F¯k\textstyle{\underline{F}^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ¯k\scriptstyle{\underline{\sigma}^{k}}F¯0.\textstyle{\underline{F}^{0}.} (2.1)

where kk indicates the order of the graded bundle σk:FkF0\sigma^{k}:F^{k}\rightarrow F^{0}; σk\sigma^{k} is a VB morphism and πk\pi^{k} is a morphism of graded bundles. In the special case k=1k=1, we recover the notion of a double vector bundle (DVB, in short), e.g. [Mac05].

Given a graded bundle σk:EkM\sigma^{k}:E^{k}\rightarrow M of order kk and an integer 0jk0\leq j\leq k we may consider a natural projection, denoted by σjk:EkEj\sigma^{k}_{j}:E^{k}\rightarrow E^{j}, where EjE^{j} is a graded bundle of order jj over MM obtained from EkE^{k} by removing from the atlas for EkE^{k} all coordinates of weights greater than jj. The graded bundle EjE^{j} obtained this way is denoted by Ek[Δj]E^{k}[\Delta\leq j] [BGR16] and called the reduction of EkE^{k} to order jj. Taking j=k1,k2,,0j=k-1,k-2,\ldots,0 we arrive at the tower of affine bundle projections

Ekσk1kEk1σk2k1Ek2σk3k2σ12E1σ01M=E0.E^{k}\xrightarrow{\sigma^{k}_{k-1}}E^{k-1}\xrightarrow{\sigma^{k-1}_{k-2}}E^{k-2}\xrightarrow{\sigma^{k-2}_{k-3}}\ldots\xrightarrow{\sigma^{2}_{1}}E^{1}\xrightarrow{\sigma^{1}_{0}}M=E^{0}.

We have σjk=σjj+1σk1k\sigma^{k}_{j}=\sigma^{j+1}_{j}\circ\ldots\circ\sigma^{k}_{k-1}, and we write shortly σj\sigma^{j} for σ0j\sigma^{j}_{0}.

A complementary construction is obtained by setting to zero all fiber coordinates in the bundle σk:EkM\sigma^{k}:E^{k}\rightarrow M of weight less than a given number 1<jk1<j\leq k. The resulting submanifold, denoted by Ek[Δj]E^{k}[\Delta\geq j], is a graded subbundle of EkE^{k} with the same base MM. In case j=kj=k, Ek[Δk]E^{k}[\Delta\geq k] is called the core of EkE^{k} and denoted by Ekwidehat\widehat{E^{k}}. The core can be endowed with a natural VB structure. This way we obtain a functor widehat:𝒢[k]𝒱\widehat{\cdot}:\mathcal{GB}[k]\rightarrow\mathcal{VB}, where 𝒢[k]\mathcal{GB}[k] is the category of graded bundles of order kk, and 𝒱=𝒢[1]\mathcal{VB}=\mathcal{GB}[1] is the category of vector bundles. In the case of multi-graded structures (F;Δ1,,Δn)(F;\Delta_{1},\ldots,\Delta_{n}), we write F𝒢[k1,,kn]F\in\mathcal{GB}[k_{1},\ldots,k_{n}], indicating that (F,Δi)𝒢[ki](F,\Delta_{i})\in\mathcal{GB}[k_{i}] and [Δi,Δj]=0[\Delta_{i},\Delta_{j}]=0 for iji\neq j. The core of the graded bundle (F,Δ1++Δn)(F,\Delta_{1}+\ldots+\Delta_{n}) is denoted in the same way as Fwidehat\widehat{F}. (It will be usually clear which weight field of FF we are referring to.)

There is an obvious graded bundle structure on the product E×FE\times F of the graded bundles EE and FF, defined by htE×F=htE×htFh^{E\times F}_{t}=h^{E}_{t}\times h^{F}_{t} where tt\in\mathbb{R}. If E,FE,F have the same base MM, then E×MFE\times_{M}F is a graded subbundle of E×FE\times F.

Given a positive integer kk and a vector bundle EME\rightarrow M we write E[k]E_{[k]} for the graded bundle (E,kΔ)(E,k\cdot\Delta), where Δ\Delta is the Euler vector field of EE. Then, for example, E[1]×MF[2]E_{[1]}\times_{M}F_{[2]} refers to a graded bundle of order two. It is the graded bundle associated with the graded vector bundle E[1]F[2]E_{[1]}\oplus F_{[2]}, where EE, FF are VBs over MM.

2.2 Double vector bundles and VB-algebroids

As we already mentioned, a structure of a DVB on a manifold DD is a pair of VBs σE:DE\sigma_{E}:D\to E and σA:DA\sigma_{A}:D\to A such that for any xDx\in D and t,st,s\in\mathbb{R} holds

tE(sAx)=sA(tEx)t\cdot_{E}(s\cdot_{A}x)=s\cdot_{A}(t\cdot_{E}x)

where E\cdot_{E} (respectively, A\cdot_{A}) denotes the multiplication by scalars in the vector bundle σE\sigma_{E} (resp., σA\sigma_{A}). The bases EE and AA carry induced VB structures over a common base MM giving rise to a diagram

D\textstyle{D\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σA\scriptstyle{\sigma_{A}}σE\scriptstyle{\sigma_{E}}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σMA\scriptstyle{\sigma^{A}_{M}}E\textstyle{E\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σME\scriptstyle{\sigma^{E}_{M}}M.\textstyle{M.} (2.2)

There is also a third vector bundle over MM, known as the core of the DVB (D,σE,σA)(D,\sigma_{E},\sigma_{A}), defined as the intersection of the kernels of the VB morphisms σE\sigma_{E} and σA\sigma_{A}, C=kerσEkerσAC=\ker\sigma_{E}\cap\ker\sigma_{A}. From the perspective of graded manifolds, DVBs are ×\mathbb{Z}\times\mathbb{Z}-graded manifolds admitting coordinates only in weights (0,0)(0,0), (1,0)(1,0), (0,1)(0,1) and (1,1)(1,1). From this perspective, the core CC is the core of the graded bundle (D,ΔE+ΔA)(D,\Delta_{E}+\Delta_{A}), where ΔE\Delta_{E} (resp., ΔA\Delta_{A}) is the Euler vector field of the vector bundle σE\sigma_{E} (resp., σA\sigma_{A}), and it will be denoted simply as C=DwidehatC=\widehat{D}.

There is a well-defined action D×MCDD\times_{M}C\rightarrow D, denoted by (d,c)d+c(d,c)\mapsto d\bm{\boldsymbol{+}}c, which arises from the affine bundle structure of DD over its order-one reduction, E×MAE\times_{M}A. A section cΓ(C)c\in\operatorname{\Gamma}(C) gives a so-called core section cc^{\dagger} of the VB σE:DE\sigma_{E}:D\to E, given by σ(em)=em+c(m)\sigma(e_{m})=e_{m}\bm{\boldsymbol{+}}c(m) where mMm\in M, em(σME)1(m)e_{m}\in(\sigma^{E}_{M})^{-1}(m). A section sΓE(D)s\in\operatorname{\Gamma}_{E}(D) is called linear if it is a VB morphism from (E,σME)(E,\sigma^{E}_{M}) to (D,σA)(D,\sigma_{A}). The subspace of linear sections (resp. core sections) is denoted by ΓE(D)\operatorname{\Gamma}^{\ell}_{E}(D) (resp., ΓEc(D)\operatorname{\Gamma}^{c}_{E}(D)).

A decomposition of a DVB DD as in (2.2) is a DVB morphism from DD to its split form D¯:=E×MA×MC\bar{D}:=E\times_{M}A\times_{M}C which is the identity on each component: the side bundles EE, AA and the core CC. Decompositions are in bijective correspondence with inclusions (also referred to as decompositions) :A×MED\sum:A\times_{M}E\to D, which are DVB morphisms inducing the identity on the side bundles AA and EE. Decompositions are also in bijective correspondence with horizontal lifts θA:Γ(A)ΓE(D)\theta_{A}:\operatorname{\Gamma}(A)\to\operatorname{\Gamma}^{\ell}_{E}(D), which are defined as splittings of the short exact sequence

0Hom(E,C)ΓE(D)Γ(A)00\to\operatorname{Hom}(E,C)\to\operatorname{\Gamma}^{\ell}_{E}(D)\to\operatorname{\Gamma}(A)\to 0 (2.3)

of 𝒞(M)\mathcal{C}^{\infty}(M)-modules where sΓE(D)s\in\operatorname{\Gamma}^{\ell}_{E}(D) projects to its base map, which turns out to be a section of σMA\sigma^{A}_{M}.

The foundation on DVBs were laid by J. Pradines [Pra75]. Double structures such as DVBs, as well as double Lie groupoids and algebroids where extensively studied by K. C. H. Mackenzie and his collaborators (see [Mac05] and references therein). In this paper, we shall deal with VB-algebroids – a pair of an algebroid and a VB structures, in compatibility, defined on a common manifold.

The compatibility condition can be stated in various equivalent ways, presenting such a structure as a Lie algebroid object in the category of vector bundles (the origins of the notion of VB-algebroids) or as a vector bundle object in the category of Lie algebroids (LA-vector bundles). See [GSM10] for definitions and the equivalence of both concepts.

Following the ideas from [GR09], one can formulate the compatibility condition as follows: a VB-algebroid structure on a manifold DD is a pair of VBs σA:DA\sigma_{A}:D\rightarrow A, σE:DE\sigma_{E}:D\rightarrow E, and a Lie algebroid structure on the vector bundle σE\sigma_{E}, such that for each tt\in\mathbb{R} the map xtAxx\mapsto t\cdot_{A}x, xDx\in D, is an algebroid morphism, see [BGV18, Definition 2.10].

It follows that (D,σE,σA)(D,\sigma_{E},\sigma_{A}) is a DVB; σMA:AM\sigma^{A}_{M}:A\to M carries an induced algebroid structure. Moreover, the anchor map D:DTE\sharp^{D}:D\to\mathrm{T}E is a DVB morphism, and ΓEc(D)ΓE(D)\operatorname{\Gamma}^{c}_{E}(D)\oplus\operatorname{\Gamma}^{\ell}_{E}(D) is a graded Lie algebra, concentrated in degrees 1-1 (the space of core sections) and 0 (the space of linear sections), with respect to the Lie bracket on Γ(σE)\operatorname{\Gamma}(\sigma_{E}).

2.3 Linearisation of graded bundles and the functor λ\lambda

We define a functor λ:𝒢[k,1]𝒢[k1,1]\lambda:\mathcal{GB}[k,1]\rightarrow\mathcal{GB}[k-1,1]. It is slightly more general then the functor of linearisation lin:𝒢[k]𝒢[k1,1]{\operatorname{lin}}:\mathcal{GB}[k]\rightarrow\mathcal{GB}[k-1,1] introduced in [BGG16]. Actually, lin=λT{\operatorname{lin}}=\lambda\circ\mathrm{T} is the composition of the tangent functor T:𝒢[k]𝒢[k,1]\mathrm{T}:\mathcal{GB}[k]\rightarrow\mathcal{GB}[k,1] with the functor λ\lambda. The construction of the functor λ\lambda is given in two steps. In the first step, we set to zero all coordinates for Fk𝒢[k,1]F^{k}\in\mathcal{GB}[k,1] of weight (0,1)(0,1). After shifting in weight by (1,0)(-1,0), the target λ(Fk)\lambda(F^{k}) is obtained from the latter by removing coordinates of weight (k,0)(k,0).

Definition 2.1 (Functor λ\lambda).

Let (Fk;Δ1k,Δ21)(F^{k};\Delta^{k}_{1},\Delta^{1}_{2}) be a weighted vector bundle as in (2.1), where (Fk,Δ1k)𝒢[k](F^{k},\Delta^{k}_{1})\in\mathcal{GB}[k] and (Fk,Δ21)(F^{k},\Delta^{1}_{2}) is a vector bundle. Let λv\lambda^{\mathrm{v}} denote the kernel of the VB morphism σk:FkF0\sigma^{k}:F^{k}\to F^{0}. Although Δ:=Δ1kΔ21\Delta:=\Delta_{1}^{k}-\Delta_{2}^{1} is not a combination with non-negative coefficients, it is a weight vector field on the submanifold λvFk\lambda^{\mathrm{v}}\subseteq F^{k} . We define the graded bundle λ(Fk)\lambda(F^{k}) as the reduction of the graded bundle (λv,Δ|λv)(\lambda^{\mathrm{v}},\Delta|_{\lambda^{\mathrm{v}}}) from order kk to k1k-1,

λ(Fk):=λv[Δ|λvk1], where λv=kerσk.{\lambda(F^{k}):=\lambda^{\mathrm{v}}[\Delta|_{\lambda^{\mathrm{v}}}\leq k-1],\text{ where }\lambda^{\mathrm{v}}=\ker\sigma^{k}.} (2.4)

In other words, we set to zero the coordinates of weight (0,1)(0,1) and then we remove the coordinates of weight (k,0)(k,0). Consider the following diagrams:

Fkkerσkλ(Fk)F¯kFk¯=σ¯k1kF¯k1,λ(Fk)λ(F1)F1widehatF¯k1M\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 7.91pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&\\&&\crcr}}}\ignorespaces{\hbox{\kern-7.91pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{F^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 0.0pt\raise-28.3611pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 31.91pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ker\sigma^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 31.91pt\raise-1.93748pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@hook{-1}}}}}}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 7.91pt\raise-1.93748pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 44.88512pt\raise-30.1111pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 81.86023pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 81.86023pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\lambda(F^{k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 95.89523pt\raise-28.3611pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-6.89125pt\raise-38.78886pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\underline{F}^{k}}$}}}}}}}{\hbox{\kern 39.38512pt\raise-38.78886pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\underline{F^{k}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 16.99257pt\raise-34.63321pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.15565pt\hbox{$\scriptstyle{=}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 6.89127pt\raise-38.78886pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 63.89641pt\raise-31.69998pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.83888pt\hbox{$\scriptstyle{\underline{\sigma}^{k}_{k-1}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 86.90398pt\raise-38.78886pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 86.90398pt\raise-38.78886pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\underline{F}^{k-1}}$}}}}}}}\ignorespaces}}}}\ignorespaces,\quad\quad\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 14.035pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\\&\crcr}}}\ignorespaces{\hbox{\kern-14.035pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\lambda(F^{k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 38.035pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 0.0pt\raise-28.3611pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 38.035pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\lambda(F^{1})\simeq\widehat{F^{1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 56.68877pt\raise-29.63889pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-8.99126pt\raise-38.78886pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\underline{F}^{k-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 48.83252pt\raise-38.78886pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 48.83252pt\raise-38.78886pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{M}$}}}}}}}\ignorespaces}}}}\ignorespaces

In the diagram on the left, the projection kerσkλ(Fk)\ker\sigma^{k}\to\lambda(F^{k}) is a fiber-wise linear isomorphism, so kerσk\ker\sigma^{k} is the pullback of the vector bundle λ(Fk)\lambda(F^{k}) with respect to the projection σ¯k1k\underline{\sigma}^{k}_{k-1}.

In the diagram on the right, λ(Fk)𝒢[k1,1]\lambda(F^{k})\in\mathcal{GB}[k-1,1] is recognized as a weighted vector bundle whose weight vector fields are inherited from Δ\Delta and Δ21\Delta_{2}^{1}. The base of the graded bundle  λ(Fk)\lambda(F^{k}) is identified as the core of the DVB F1𝒢[1,1]F^{1}\in\mathcal{GB}[1,1]. If (xa,y(α,β)i)(x^{a},y^{i}_{(\alpha,\beta)}) are graded coordinates on the weighted VB FkF^{k}, then the adapted coordinates on λ(Fk)\lambda(F^{k}) are obtained by omitting those y(α,β)iy^{i}_{(\alpha,\beta)} with (α,β){(k,0),(0,1)}(\alpha,\beta)\in\{(k,0),(0,1)\}, and the coordinates of weight (w,1)(w,1) are assigned a new weight (w1,1)(w-1,1).

Lemma 2.2.

Let σk:EkM\sigma^{k}:E^{k}\rightarrow M be a graded bundle of order kk. There are canonical isomorphism of weighted vector bundles:

  1. (i)

    If σk:EkM\sigma^{k}:E^{k}\rightarrow M is a graded bundle of order kk, then λ(TEk)lin(Ek)\lambda(\mathrm{T}E^{k})\simeq{\operatorname{lin}}(E^{k}).

  2. (ii)

    If σ:EM\sigma:E\rightarrow M is a vector bundle, then λ(TkE)Tk1E\lambda(\mathrm{T}^{k}E)\simeq\mathrm{T}^{k-1}E.

Proof.

Only (ii) needs a proof, as (i) follows directly from the construction of λ\lambda and the linearisation functor.

For the proof of (ii), observe that the inclusion λv(TkE)TkE\lambda^{\mathrm{v}}(\mathrm{T}^{k}E)\subset\mathrm{T}^{k}E is realized by the mapping

Tk1E×Tk1MTkMTkE,([a]k1,[γ]k)[tta(t)]k,\mathrm{T}^{k-1}E\times_{\mathrm{T}^{k-1}M}\mathrm{T}^{k}M\hookrightarrow\mathrm{T}^{k}E,\quad([a]_{k-1},[\gamma]_{k})\mapsto[t\mapsto ta(t)]_{k}, (2.5)

where curves a:Ea:\mathbb{R}\rightarrow E and γ:M\gamma:\mathbb{R}\rightarrow M are such that σa=γ\sigma\circ a=\gamma. Indeed, the image of the mapping (2.5) is the subbundle TMkETkE\mathrm{T}^{k}_{M}E\subset\mathrm{T}^{k}E. In the standard local coordinates (xa,yi)(x^{a},y^{i}) on EE, it is given by the vanishing coordinates of weight (0,1)(0,1), i.e., λv(TkE)=TMkE={(xa,(α),yi,(β)):yi,(0)=0}\lambda^{\mathrm{v}}(\mathrm{T}^{k}E)=\mathrm{T}^{k}_{M}E=\{(x^{a,({\alpha})},y^{i,({\beta})}):y^{i,(0)}=0\}.

Finally, we realize that the canonical projection λv(TkE)λ(TkE)\lambda^{\mathrm{v}}(\mathrm{T}^{k}E)\rightarrow\lambda(\mathrm{T}^{k}E) defined locally by removing coordinates of weight (k,0)(k,0), i.e., the coordinates xa,(k)x^{a,(k)}, coincides with the projection

Tk1E×Tk1MTkMTk1E,([a]k1,[γ]k)[a]k1.\mathrm{T}^{k-1}E\times_{\mathrm{T}^{k-1}M}\mathrm{T}^{k}M\rightarrow\mathrm{T}^{k-1}E,\quad([a]_{k-1},[\gamma]_{k})\mapsto[a]_{k-1}.

From (2.5) we easily find that the obtained isomorphism Tk1Eλ(TkE)\mathrm{T}^{k-1}E\rightarrow\lambda(\mathrm{T}^{k}E), denoted by IEkI^{k}_{E}, has the formula

(IEk)(xa,(α))=xa,(α),(IEk)(yi,(β))=βy˙i,(β1){{(I^{k}_{E})}^{*}(x^{a,({\alpha})})=x^{a,({\alpha})},\quad{(I^{k}_{E})}^{*}(y^{i,({\beta})})={\beta}\dot{y}^{i,({\beta}-1)}} (2.6)

where (xa,(α),yi,(β))(x^{a,({\alpha})},y^{i,({\beta})}), 0αk10\leq{\alpha}\leq k-1, 1βk1\leq{\beta}\leq k, are the coordinates for λ(TkE)\lambda(\mathrm{T}^{k}E) inherited from TkE\mathrm{T}^{k}E. ∎

2.4 Natural inclusions and isomorphisms I

For later use, we shall fix the natural inclusions:

jMk:TMTkMwidehatTkM,[γ]1[tγ(tk/k!)]k,j^{k}_{M}:\mathrm{T}M\rightarrow\widehat{\mathrm{T}^{k}M}\subset\mathrm{T}^{k}M,[\gamma]_{1}\mapsto[t\mapsto\gamma(t^{k}/k!)]_{k}, (2.7)

and

iMk,l:Tk+lMTkTlM,[γ]k+l[t[sγ(t+s)]l]k,i^{k,l}_{M}:\mathrm{T}^{k+l}M\rightarrow\mathrm{T}^{k}\mathrm{T}^{l}M,[\gamma]_{k+l}\longmapsto[t\mapsto[s\mapsto\gamma(t+s)]_{l}]_{k}, (2.8)

so iMk,l([γ]k+l)=[𝐭lγ]k=(𝐭k𝐭lγ)(0)i^{k,l}_{M}([\gamma]_{k+l})=[\bm{\mathrm{t}}^{l}\gamma]_{k}={\left(\bm{\mathrm{t}}^{k}\bm{\mathrm{t}}^{l}\gamma\right)(0)} where 𝐭lγ:TlM\bm{\mathrm{t}}^{l}\gamma:\mathbb{R}\rightarrow\mathrm{T}^{l}M is the ll-th tangent lift of the curve γ\gamma. In coordinates,

jMk(xa,x˙a)=(xa,0,0,x˙a),j^{k}_{M}(x^{a},\dot{x}^{a})=(x^{a},0,\ldots 0,\dot{x}^{a}), (2.9)
(iMk,l)(xa,(α,β))=xa,(α+β).(i^{k,l}_{M})^{\ast}(x^{a,{({\alpha},{\beta})}})=x^{a,({\alpha}+{\beta})}. (2.10)

where xa,(α,β)=(xa,(α))(β)x^{a,({\alpha},{\beta})}=\left(x^{a,({\alpha})}\right)^{({\beta})}. In addition to jMkj^{k}_{M}, given a vector bundle σ:EM\sigma:E\rightarrow M, there is a canonical VB isomorphism of the core bundle of (Tk1E,dTk1ΔE+ΔTk1E)(\mathrm{T}^{k-1}E,\mathrm{d}_{\mathrm{T}^{k-1}}\Delta_{E}+\Delta_{\mathrm{T}^{k-1}E}) and the vector bundle (E,ΔE)(E,\Delta_{E}) which is defined by

ȷEk:ETk1EwidehatTk1E,v[ttk1(k1)!v]k1\jmath^{k}_{E}:E\xrightarrow{\simeq}\widehat{\mathrm{T}^{k-1}E}\subset\mathrm{T}^{k-1}E,\quad v\mapsto[t\mapsto\frac{t^{k-1}}{(k-1)!}v]_{k-1} (2.11)

The compatibility with the map jMkj^{k}_{M} is expressed by the commutative diagram

TM\textstyle{\mathrm{T}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}jMk\scriptstyle{j^{k}_{M}}ȷTMk1\scriptstyle{\jmath^{k-1}_{\mathrm{T}M}}TkM\textstyle{\mathrm{T}^{k}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}iMk1,1\scriptstyle{i^{k-1,1}_{M}}Tk1TM\textstyle{T^{k-1}\mathrm{T}M} (2.12)

A graded bundle (Ek,Δ)(E^{k},\Delta) embeds naturally into its linearisation via the digitalisation map

diagk:EklinEk,(diagk)(y˙wi)=wywi, 1wk\mathrm{diag}^{k}:E^{k}\hookrightarrow{\operatorname{lin}}{E^{k}},\quad(\mathrm{diag}^{k})^{*}(\dot{y}^{i}_{w})=wy^{i}_{w},\,1\leq w\leq k (2.13)

in the adapted coordinates (xa,ywi;y˙wi)(x^{a},y^{i^{\prime}}_{w^{\prime}};\dot{y}^{i}_{w}), where 1wk1\leq w\leq k, 1wk11\leq w^{\prime}\leq k-1, on lin(Ek)=λ(TEk){\operatorname{lin}}(E^{k})=\lambda(\mathrm{T}E^{k}) induced from TEk\mathrm{T}E^{k}, as mentioned earlier. Moreover, diagk\mathrm{diag}^{k} covers the identity over Ek1E^{k-1}. This map is induced from the weight vector field considered as a map Δ:Ek𝐕Ek\Delta:E^{k}\rightarrow\mathbf{V}E^{k}, where 𝐕Ek=kerTσk\mathbf{V}E^{k}=\ker\mathrm{T}\sigma^{k} denotes the vertical subbundle of TEk\mathrm{T}E^{k}. In other words, the weight vector field Δ\Delta is projectable with respect to the canonical projection 𝐕Eklin(Ek)\mathbf{V}E^{k}\rightarrow{\operatorname{lin}}(E^{k}). Moreover, in the special case Ek=TkME^{k}=\mathrm{T}^{k}M, the map diagk\mathrm{diag}^{k} coincides with iM1,k1:TkMTTk1Mi^{1,k-1}_{M}:\mathrm{T}^{k}M\rightarrow\mathrm{T}\mathrm{T}^{k-1}M composed with the inverse of the isomorphism I:lin(TkM)=λ(TTkM)λ(κMk)λ(TkTM)ITMkTk1TMTTk1MI:{\operatorname{lin}}(\mathrm{T}^{k}M)=\lambda(\mathrm{T}\mathrm{T}^{k}M)\xrightarrow{\lambda(\kappa^{k}_{M})}\lambda(\mathrm{T}^{k}\mathrm{T}M)\xrightarrow{I^{k}_{\mathrm{T}M}}\mathrm{T}^{k-1}\mathrm{T}M\simeq\mathrm{T}\mathrm{T}^{k-1}M, where ITMkI^{k}_{\mathrm{T}M} is the isomorphism established in Lemma 2.2 (ii). The isomorphism I:lin(TkM)TTk1MI:{\operatorname{lin}}(\mathrm{T}^{k}M)\rightarrow\mathrm{T}\mathrm{T}^{k-1}M coincides with the isomorphism found in ([BGG15b, Example 2.2.3], [BGR16]) and is given by

I(dxa,(α))=1α+1dxa,(α+1)I^{\ast}(\mathrm{d}x^{a,({\alpha})})=\frac{1}{{\alpha}+1}\mathrm{d}x^{a,({\alpha}+1)}

for α=0,1,,k1{\alpha}=0,1,\ldots,k-1.

Lemma 2.3.

Let σi:EikMi\sigma_{i}:E^{k}_{i}\rightarrow M_{i}, for i=1,2i=1,2, be graded bundles of order kk and let ϕ:E1kE2k\phi:E^{k}_{1}\rightarrow E^{k}_{2} be a 𝒢[k]\mathcal{GB}[k]-morphisms. Then the linearisation of ϕ\phi intertwines the canonical inclusions diagi:Eiklin(Eik)\mathrm{diag}_{i}:E_{i}^{k}\hookrightarrow{\operatorname{lin}}(E^{k}_{i}):

lin(E1k)\textstyle{{\operatorname{lin}}(E^{k}_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}lin(ϕ)\scriptstyle{{\operatorname{lin}}(\phi)}lin(E2k)\textstyle{{\operatorname{lin}}(E^{k}_{2})}E1k\textstyle{E^{k}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}diag1\scriptstyle{\mathrm{diag}_{1}}ϕ\scriptstyle{\phi}E2k\textstyle{E^{k}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}diag2\scriptstyle{\mathrm{diag}_{2}}
Proof.

The map lin(ϕ){\operatorname{lin}}(\phi) is the unique map which makes the following diagram commutative:

𝐕E1k\textstyle{\mathbf{V}E^{k}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tϕ\scriptstyle{\mathrm{T}\phi}𝐕E2k\textstyle{\mathbf{V}E^{k}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}lin(E1k)\textstyle{{\operatorname{lin}}(E^{k}_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}lin(ϕ)\scriptstyle{{\operatorname{lin}}(\phi)}lin(E2k)\textstyle{{\operatorname{lin}}(E^{k}_{2})}

where 𝐕Eik=kerTσiTEik\mathbf{V}E^{k}_{i}=\ker\mathrm{T}\sigma_{i}\subseteq\mathrm{T}E^{k}_{i}. The weight vector fields Δ1\Delta_{1}, Δ2\Delta_{2} are ϕ\phi-related as ϕ:E1kE2k\phi:E^{k}_{1}\rightarrow E^{k}_{2} is a 𝒢[k]\mathcal{GB}[k]-morphism ( [GR11, Theorem 2.3]). Hence TϕΔ1=Δ2ϕ\mathrm{T}\phi\circ\Delta_{1}=\Delta_{2}\circ\phi, and the thesis follows directly from the definition of the diagonalisation map. ∎

2.5 Vector bundle comorphisms

We shall recall the definition of a comorphism between vector bundles from [JR18] where one can also find more information and references on the origins and generalizations of this concept.

Definition 2.4.

A vector bundle comorphism (VB comorphism, for short), from a vector bundle σ1:E1M1\sigma_{1}:E_{1}\rightarrow M_{1} to a vector bundle σ2:E2M2\sigma_{2}:E_{2}\rightarrow M_{2}, is a relation rE1×E2r\subset E_{1}\times E_{2}, for which there exist a base map r¯:M2M1\underline{r}:M_{2}\rightarrow M_{1} and a VB morphism r!:r¯E1E2r^{!}:{\underline{r}}^{*}E_{1}\rightarrow E_{2} covering the identity on M2M_{2} such that

r={(v,r!(v,y)):vE1,yM2,σ1(v)=r¯(y)}r=\{(v,r^{!}(v,y)):v\in E_{1},y\in M_{2},\sigma_{1}(v)=\underline{r}(y)\}

where r¯E1E1×M2{\underline{r}}^{*}E_{1}\subset E_{1}\times M_{2} is the pullback of the vector bundle σ1\sigma_{1} with respect to the map r¯\underline{r}. We say that the base map r¯:M2M1\underline{r}:M_{2}\rightarrow M_{1} (which is uniquely defined) covers rr, and we depict this in the following diagram:

E1σ1rE2σ2M1M2r¯.\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 9.11624pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\\&\crcr}}}\ignorespaces{\hbox{\kern-7.84125pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{E_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 0.0pt\raise-18.88698pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-0.77625pt\hbox{$\scriptstyle{\sigma_{1}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 0.0pt\raise-28.62398pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 16.60759pt\raise 4.35625pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.35625pt\hbox{$\scriptstyle{r}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 32.14124pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 4.5pt\hbox{{\hbox{\kern-4.5pt\hbox{\ignorespaces\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@stopper}}}}}{\hbox{\kern 0.0pt\hbox{\ignorespaces\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}}}}}}{\hbox{\kern 0.0pt\hbox{\ignorespaces\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{-1}}}}}}}}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 34.39124pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{E_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 42.23248pt\raise-18.88698pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-0.77625pt\hbox{$\scriptstyle{\sigma_{2}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 42.23248pt\raise-28.62398pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-9.11624pt\raise-37.77396pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{M_{1}}$}}}}}}}{\hbox{\kern 33.11624pt\raise-37.77396pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{M_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 15.61624pt\raise-31.93507pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.83888pt\hbox{$\scriptstyle{\underline{r}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 9.11626pt\raise-37.77396pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}\ignorespaces}}}}\ignorespaces.

Thus, rr is the union of graphs of linear maps ry:(E1)r¯(y)(E2)yr_{y}:(E_{1})_{\underline{r}(y)}\rightarrow(E_{2})_{y} between the corresponding fibers, where yy varies in M2M_{2}. There is a one-to-one correspondence between VB comorphisms σ1σ2\sigma_{1}{-\!\!\!-\!\!\rhd}\sigma_{2} and VB morphisms σ2σ1\sigma_{2}^{\ast}\rightarrow\sigma_{1}^{\ast} between the dual bundles. A VB comorphism r:σ1σ2r:\sigma_{1}{-\!\!\!-\!\!\rhd}\sigma_{2} gives rise to a mapping between the spaces of sections,

r:Γ(σ1)Γ(σ2),r(s)(y)=ry(s(r¯(y))).{\overrightarrow{r}}:\operatorname{\Gamma}(\sigma_{1})\rightarrow\operatorname{\Gamma}(\sigma_{2}),\quad{\overrightarrow{r}}(s)(y)=r_{y}(s(\underline{r}(y))).

The map r{\overrightarrow{r}} satisfies

r(s+s)=r(s)+r(s),r(fs)=r¯(f)s{\overrightarrow{r}}(s+s^{\prime})={\overrightarrow{r}}(s)+{\overrightarrow{r}}(s^{\prime}),\quad\quad{\overrightarrow{r}}(f\cdot s)=\underline{r}^{\ast}(f)\cdot{\overrightarrow{s}} (2.14)

and any such map gives rise to a VB comorphism r:σ1σ2r:\sigma_{1}{-\!\!\!-\!\!\rhd}\sigma_{2}.

VB comorphisms form a category denoted by 𝒱𝒞\mathcal{VBC}. A morphism from r𝒱𝒞r\in\mathcal{VBC} to r𝒱𝒞r^{\prime}\in\mathcal{VBC}, where r:σ1σ2r:\sigma_{1}{-\!\!\!-\!\!\rhd}\sigma_{2} and r:σ1σ2r^{\prime}:\sigma_{1}^{\prime}{-\!\!\!-\!\!\rhd}\sigma_{2}^{\prime} are VB comorphism and σi:EiMi\sigma_{i}:E_{i}\rightarrow M_{i}, σi:EiMi\sigma_{i}^{\prime}:E_{i}^{\prime}\rightarrow M_{i}^{\prime} are vector bundles, is given by a pair (ϕ1,ϕ2)(\phi_{1},\phi_{2}) of VB morphisms ϕi:EiEi\phi_{i}:E_{i}\rightarrow E_{i}^{\prime} such that (ϕ1×ϕ2)(r)r(\phi_{1}\times\phi_{2})(r)\subset r^{\prime} ([JR13, Definition 2.3 and Proposition 2.6]). It is denoted by (ϕ1,ϕ2):rr(\phi_{1},\phi_{2}):r\Rightarrow r^{\prime}.

A VB comorphism r:σ1σ2r:\sigma_{1}{-\!\!\!-\!\!\rhd}\sigma_{2} is weighted of order kk if the total spaces E1,E2E_{1},E_{2} are given a structure of a graded bundle of order kk with respect to which rr is a graded subbundle of the product E1×E2E_{1}\times E_{2}.

We shall need the following result in Section 3. Roughly speaking, it states that λ\lambda is also a functor on the category of weighted vector bundle comorphisms.

Lemma 2.5.

Let F1k,F2k𝒢[k,1]F_{1}^{k},F_{2}^{k}\in\mathcal{GB}[k,1] be weighted, order kk, vector bundles and let πi:FikF¯ik\pi_{i}:F_{i}^{k}\rightarrow\underline{F}_{i}^{k} denotes the corresponding VB projections. Let rk:π1π2r^{k}:\pi_{1}{\rightarrow\!\!\vartriangleright}\pi_{2} be a weighted, order kk, VB comorphism covering r¯k:F¯2kF¯1k\underline{r}^{k}:\underline{F}_{2}^{k}\rightarrow\underline{F}_{1}^{k}. Then λ(rk):λ(F1k)λ(F2k)\lambda(r^{k}):\lambda(F_{1}^{k}){\rightarrow\!\!\vartriangleright}\lambda(F_{2}^{k}) is a VB comorphism covering r¯k1:F¯2k1F¯1k1\underline{r}^{k-1}:\underline{F}_{2}^{k-1}\rightarrow\underline{F}_{1}^{k-1}:

λ(F1k)\textstyle{\lambda(F_{1}^{k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}λ(rk)\scriptstyle{\lambda(r^{k})}λ(F2k)\textstyle{\lambda(F_{2}^{k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F¯1k1\textstyle{\underline{F}_{1}^{k-1}}F¯2k1\textstyle{\underline{F}_{2}^{k-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}r¯k1\scriptstyle{\underline{r}^{k-1}}

Moreover, if (ϕ1,ϕ2):rr(\phi_{1},\phi_{2}):r\Rightarrow r^{\prime} is a morphism between weighted VB comorphisms r:F1F2r:F_{1}{\rightarrow\!\!\vartriangleright}F_{2} and r:F1F2r^{\prime}:{F_{1}^{\prime}}{\rightarrow\!\!\vartriangleright}{F_{2}^{\prime}} then (λ(ϕ1),λ(ϕ2)):λ(r)λ(r)(\lambda(\phi_{1}),\lambda(\phi_{2})):\lambda(r)\Rightarrow\lambda(r^{\prime}) is the same.

Proof.

Note that rkr^{k} is a weighted vector subbundle of F1k×F2kF_{1}^{k}\times F_{2}^{k}, hence λ(rk)\lambda(r^{k}) is a weighted vector subbundle of λ(F1k×F2k)=λ(F1k)×λ(Fk2)\lambda(F_{1}^{k}\times F_{2}^{k})=\lambda(F_{1}^{k})\times\lambda(F^{2}_{k}). Let us trace the subsequent steps of the construction of the weighted vector bundle λ(rk)\lambda(r^{k}), as in Definition 2.1. We have λv(rk)=rk(kerσ1k×kerσ2k)\lambda^{\mathrm{v}}(r^{k})=r^{k}\cap(\ker\sigma^{k}_{1}\times\ker\sigma^{k}_{2}) where σik:FikFi0\sigma_{i}^{k}:F_{i}^{k}\rightarrow F^{0}_{i} are as in (2.1) for i=1,2i=1,2. Hence, λv(rk)\lambda^{\mathrm{v}}(r^{k}) is a VB comorphism kerσ1kkerσ2k\ker\sigma^{k}_{1}{\rightarrow\!\!\vartriangleright}\ker\sigma^{k}_{2} covering r¯k:F¯2kF¯1k\underline{r}^{k}:\underline{F}_{2}^{k}\rightarrow\underline{F}_{1}^{k}. The goal λ(rk)\lambda(r^{k}) is obtained from λv(rk)\lambda^{\mathrm{v}}(r^{k}) by the reduction to order k1k-1 of the base map r¯k\underline{r}^{k}. Since the projections λv(Fik)λ(Fik)\lambda^{\mathrm{v}}(F^{k}_{i})\rightarrow\lambda(F^{k}_{i}) are fiber-wise linear isomorphisms, λ(rk)\lambda(r^{k}) remains a VB comorphism.

For the last part of Lemma, we have already noticed that the functor λ\lambda preserves the products and inclusions. By [JR18, Proposition 2.6] (ϕ1×ϕ2)(r)r(\phi_{1}\times\phi_{2})(r)\subseteq r^{\prime}, hence (λ(ϕ1)×λ(ϕ2))(λ(r))r(\lambda(\phi_{1})\times\lambda(\phi_{2}))(\lambda(r))\subseteq r^{\prime}, so (λ(ϕ1),λ(ϕ2))(\lambda(\phi_{1}),\lambda(\phi_{2})) is a morphism in the category 𝒱𝒞\mathcal{VBC}. ∎

The core Ekwidehat\widehat{E^{k}} acts naturally on the graded bundle EkE^{k}. This action Ek×MEkwidehatEkE^{k}\times_{M}\widehat{E^{k}}\to E^{k} is denoted by (ak,v)ak+vEk(a^{k},v)\mapsto a^{k}\bm{\boldsymbol{+}}v\in E^{k} and gives rise to a VB comorphism,

vEkwidehat\textstyle{v\in\widehat{E^{k}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}TEkv(ak)\textstyle{\mathrm{T}E^{k}\ni{v^{\uparrow}}(a^{k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M\textstyle{M}Ekak\textstyle{E^{k}\ni a^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σk\scriptstyle{\sigma^{k}} (2.15)

where v(ak)TakEk{v^{\uparrow}}(a^{k})\in\mathrm{T}_{a^{k}}E^{k} is the vector represented by the curve tak+(tv)t\mapsto a^{k}\bm{\boldsymbol{+}}(tv). In coordinates (xa,ywi,zkμ)(x^{a},y^{i}_{w},z^{\mu}_{k}) on EkE^{k}, where ywiy^{i}_{w}’s have weights 1wk11\leq w\leq k-1, and w(zkμ)=k\mathrm{w}(z^{\mu}_{k})=k, the associated map on sections is given by vv=vμ(x)zkμv\mapsto{v^{\uparrow}}=v^{\mu}(x)\partial_{z^{\mu}_{k}}, where v=μvμ(x)cμv=\sum_{\mu}v^{\mu}(x)c_{\mu} and (cμ)(c_{\mu}) is a local frame of Γ(Ekwidehat)\operatorname{\Gamma}(\widehat{E^{k}}). Since EkEk1E^{k}\rightarrow E^{k-1} is an affine bundle modelled on the pullback of the core EkwidehatM\widehat{E^{k}}\rightarrow M, there is a map

Ek×Ek1EkEkwidehat,(a,a)aa,E^{k}\times_{E^{k-1}}E^{k}\rightarrow\widehat{E^{k}},\quad(a^{\prime},a)\mapsto a^{\prime}\bm{\boldsymbol{-}}a, (2.16)

where aaa^{\prime}-a is the unique vector vΓ(Ekwidehat)v\in\operatorname{\Gamma}(\widehat{E^{k}}) such that a+v=aa\bm{\boldsymbol{+}}v=a^{\prime}.

Lemma 2.6.

The mapping associated with the VB comorphism (2.15),

Γ(Ekwidehat)𝔛k(Ek),vv,\operatorname{\Gamma}(\widehat{E^{k}})\rightarrow\mathfrak{X}_{-k}(E^{k}),v\mapsto{v^{\uparrow}}, (2.17)

is a 𝒞(M)\mathcal{C}^{\infty}(M)-module isomorphism. Moreover, if σi:EikMi\sigma_{i}:E^{k}_{i}\rightarrow M_{i}, for i=1,2i=1,2, are graded bundles of order kk and ϕ:E1kE2k\phi:E^{k}_{1}\rightarrow E^{k}_{2} is a 𝒢[k]\mathcal{GB}[k]-morphism then weight k-k vector fields Xi𝔛k(E1k)X_{i}\in\mathfrak{X}_{-k}(E^{k}_{1}) are ϕ\phi-related if and only if the corresponding sections viΓ(Eik𝑤𝑖𝑑𝑒ℎ𝑎𝑡)v_{i}\in\operatorname{\Gamma}(\widehat{E^{k}_{i}}) are ϕ𝑤𝑖𝑑𝑒ℎ𝑎𝑡\widehat{\phi}-related. If M1=M2M_{1}=M_{2} and ϕ\phi covers the identity, then the last condition means that v2=ϕ𝑤𝑖𝑑𝑒ℎ𝑎𝑡v1v_{2}=\widehat{\phi}\circ v_{1}.

Proof.

Let viΓ(Eikwidehat)v_{i}\in\operatorname{\Gamma}(\widehat{E^{k}_{i}}), Xi=viX_{i}={v_{i}^{\uparrow}} for i=1,2i=1,2. The vector fields XiX_{i} are represented by the families of curves tai+tvi(ai¯)t\mapsto a_{i}\bm{\boldsymbol{+}}tv_{i}(\underline{a_{i}}) where aiEika_{i}\in E_{i}^{k} and ai¯=σik(ai)Mi\underline{a_{i}}=\sigma^{k}_{i}(a_{i})\in M_{i}. The sections v1,v2v_{1},v_{2} are ϕwidehat\widehat{\phi}-related if and only if v2(m2)=ϕwidehat(v1(m1))v_{2}(m_{2})=\widehat{\phi}(v_{1}(m_{1})) for any pair (m1,m2)(m_{1},m_{2}) such that ϕ¯(m1)=m2\underline{\phi}(m_{1})=m_{2}. Note that ϕ(a1+tv1(a1¯))=ϕ(a1)+tϕwidehat(v1(a1¯))\phi(a_{1}\bm{\boldsymbol{+}}tv_{1}(\underline{a_{1}}))=\phi(a_{1})\bm{\boldsymbol{+}}t\widehat{\phi}(v_{1}(\underline{a_{1}})), hence if v1,v2v_{1},v_{2} are ϕwidehat\widehat{\phi}-related then (Tϕ)X1(a1)=X2(ϕ(a2))(\mathrm{T}\phi)X_{1}(a_{1})=X_{2}(\phi(a_{2})). Thus, X1,X2X_{1},X_{2} are ϕ\phi-related. The proof in the converse direction is very similar and is left to the reader. ∎

2.6 Higher algebroids

It is well-known that a Lie algebroid (σ:AM,[,],)(\sigma:A\to M,[\cdot,\cdot],\sharp) can be represented as a linear Poisson tensor on AA^{\ast}, the total space of the dual vector bundle. This, in turn, gives rise to a VB morphism ε:TATA\varepsilon:\mathrm{T}^{\ast}A\to\mathrm{T}A^{\ast} which is a Poisson map and retains all the information about the algebroid structure on AA. The dual of ε\varepsilon is a VB comorphism κ:TστA\kappa:\mathrm{T}\sigma{\rightarrow\!\!\vartriangleright}\tau_{A} which was a starting point in the concept of HAs originated in [JR13].

A general algebroid structure on a vector bundle σ:EM\sigma:E\rightarrow M can be encoded as a VB comorphism κ:TστE\kappa:\mathrm{T}\sigma{-\!\!\!-\!\!\rhd}\tau_{E} of a special kind, see [JR18], Proposition 2.15. In this correspondence κ\kappa should be also a vector subbundle of τE×Tσ\tau_{E}\times\mathrm{T}\sigma, and the induced VB morphism between the core bundles should be the identity,

κwidehat=idTEwidehat\widehat{\kappa}=\operatorname{id}_{\widehat{\mathrm{T}E}} (2.18)

Let us recall that the core of the DVB TE\mathrm{T}E is the subbundle 𝐕ME\mathbf{V}_{M}E of the vertical bundle 𝐕E\mathbf{V}E of EE, and it is naturally identified with the vector bundle EE itself. Moreover, algebroid morphisms ϕ:(E1,κ1)(E2,κ2)\phi:(E_{1},\kappa_{1})\rightarrow(E_{2},\kappa_{2}) are in a one-to-one correspondence with 𝒱𝒞\mathcal{VBC}-morphisms (Tϕ,Tϕ):κ1κ2(\mathrm{T}\phi,\mathrm{T}\phi):\kappa_{1}\Rightarrow\kappa_{2}. The above concept of an algebroid has a direct analogue in higher-order, which we shall recall now.

Definition 2.7.

[JR18] A general (kthk^{\mathrm{th}}-order) higher algebroid (HA, in short) is a graded bundle σk:EkM\sigma^{k}:E^{k}\rightarrow M of order kk together with a weighted VB comorphism κkTkE1×TEk\kappa^{k}\subset\mathrm{T}^{k}E^{1}\times\mathrm{T}E^{k} from Tkσ1\mathrm{T}^{k}\sigma^{1} to τEk\tau_{E^{k}} (covering a mapping k:EkTkM\sharp^{k}:E^{k}\rightarrow\mathrm{T}^{k}M) such that the relation κ1:Tσ1TτE1\kappa^{1}:\mathrm{T}\sigma^{1}{\rightarrow\!\!\vartriangleright}\mathrm{T}\tau_{E^{1}}, being the reduction to order one of κk\kappa^{k}, equips σ1:E1M\sigma^{1}:E^{1}\rightarrow M with an algebroid structure:

TkE1\textstyle{\mathrm{T}^{k}E^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tkσ1\scriptstyle{\mathrm{T}^{k}\sigma^{1}}κk\scriptstyle{\kappa^{k}}TEk\textstyle{\mathrm{T}E^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τEk\scriptstyle{\tau_{E^{k}}}TkM\textstyle{\mathrm{T}^{k}M}Ek\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces E^{k}}k\scriptstyle{\sharp^{k}} (2.19)

In addition:

  1. (i)

    If κ1\kappa^{1} is a symmetric relation, then the HA (Ek,κk)(E^{k},\kappa^{k}) is called skew.111111This is equivalent to saying that the bracket [,][\cdot,\cdot] on Γ(E1)\operatorname{\Gamma}(E^{1}) is skew-symmetric, see [JR18].

  2. (ii)

    If (σ1,κ1)(\sigma^{1},\kappa^{1}) is skew and, in addition, the diagram

    TkE1\textstyle{\mathrm{T}^{k}E^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tk1\scriptstyle{\mathrm{T}^{k}\sharp^{1}}κk\scriptstyle{\kappa^{k}}TEk\textstyle{\mathrm{T}E^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tk\scriptstyle{\mathrm{T}\sharp^{k}}TkTM\textstyle{\mathrm{T}^{k}\mathrm{T}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κMk\scriptstyle{\kappa^{k}_{M}}TTkM\textstyle{\mathrm{T}\mathrm{T}^{k}M} (2.20)

    is commutative, i.e., (Tk1,Tk):κkκMk(\mathrm{T}^{k}\sharp^{1},\mathrm{T}\sharp^{k}):\kappa^{k}\Rightarrow\kappa^{k}_{M} is a morphism in 𝒱𝒞\mathcal{VBC}, then we call (σk,κk)(\sigma^{k},\kappa^{k}) an almost Lie higher algebroid;

  3. (iii)

    Both vector bundles, Tkσ1\mathrm{T}^{k}\sigma^{1} and τEk\tau_{E^{k}} in the diagram (2.19), carry a canonical algebroid structure.121212 The kk-tangent lift of (σ1,κ1)(\sigma^{1},\kappa^{1}) gives an algebroid structure on Tkσ1\mathrm{T}^{k}\sigma^{1}. If (Ek,κk)(E^{k},\kappa^{k}) is a skew HA and κk\kappa^{k} is a subalgebroid of the product of these algebroids then (σk,κk)(\sigma^{k},\kappa^{k}) is called a Lie HA.131313This condition can be restated as the dual VB morphism εk:TEkTk(E1)\varepsilon^{k}:\mathrm{T}^{\ast}E^{k}\rightarrow\mathrm{T}^{k}(E^{1})^{\ast} is a Poisson map, e.g. [Gra12]. Moreover, a Lie HA has to be AL, i.e., the condition (iii) implies (ii), see [JR18].

A morphism between higher algebroids (σEk:EkM,κk,E)(\sigma^{k}_{E}:E^{k}\rightarrow M,\kappa^{k,E}) and (σFk:FkN,κk,F)(\sigma^{k}_{F}:F^{k}\rightarrow N,\kappa^{k,F}) is a morphism of graded bundles ϕk:EkFk\phi^{k}:E^{k}\rightarrow F^{k} such that (Tkϕ1,Tϕk):κk,Eκk,F(\mathrm{T}^{k}\phi^{1},\mathrm{T}\phi^{k}):\kappa^{k,E}\Rightarrow\kappa^{k,F} is a 𝒱𝒞\mathcal{VBC}-morphism. Higher algebroids with 𝒱𝒞\mathcal{VBC}-morphisms form a category. The reduction of a HA (Ek,κk)(E^{k},\kappa^{k}) to a lower order jj, 1j<k1\leq j<k gives a HA denoted by (Ej,κj)(E^{j},\kappa^{j}) which is skew (resp. AL, Lie) if (Ek,κk)(E^{k},\kappa^{k}) was so.

Example 2.8 ([JR18] HAs of order 22, in coordinates).

Let (xa,yi,zμ)(x^{a},y^{i},z^{\mu}) be local graded coordinates on a graded bundle σ2:E2M\sigma^{2}:E^{2}\rightarrow M of order 22. Taking into account only the graded bundle structure of κ2\kappa^{2}, we obtain the following system of equations for κ2T2E1×MTE2𝒢[1,2]\kappa^{2}\subset\mathrm{T}^{2}E^{1}\times_{M}\mathrm{T}E^{2}\in\mathcal{GB}[1,2]. (We have underlined the coordinates on TE2\mathrm{T}E^{2} in order to distinguish them from the coordinates on T2E1\mathrm{T}^{2}E^{1}.)

κ2:{x˙a=Qiay¯ix¨a=12Qijay¯iy¯j+Qμaz¯μ, where Qija=Qjia,x¯˙a=Q~iayiy¯˙i=Qjiy˙j+Qjkiy¯jyk, where Qji=δji,z¯˙μ=Qiμy¨i+Qijμy¯iy˙j+Qνiμz¯νyi+12Qij,kμy¯iy¯jyk, where Qij,kμ=Qji,kμ,\kappa^{2}:\begin{cases}\dot{x}^{a}=&Q^{a}_{i}\,\underline{y}^{i}\\ \ddot{x}^{a}=&\frac{1}{2}\,Q^{a}_{ij}\,\underline{y}^{i}\underline{y}^{j}+Q^{a}_{\mu}\,\underline{z}^{\mu},\quad\text{ where }Q^{a}_{ij}=Q^{a}_{ji},\\ \dot{\underline{x}}^{a}=&\tilde{Q}^{a}_{i}\,y^{i}\\ \dot{\underline{y}}^{i}=&Q^{i}_{j}\dot{y}^{j}+Q^{i}_{jk}\,\underline{y}^{j}y^{k},\quad\text{ where }Q^{i}_{j}=\delta^{i}_{j},\\ \dot{\underline{z}}^{\mu}=&Q^{\mu}_{i}\,\ddot{y}^{i}+Q^{\mu}_{ij}\,\underline{y}^{i}\dot{y}^{j}+Q^{\mu}_{\nu i}\,\underline{z}^{\nu}y^{i}+\frac{1}{2}Q^{\mu}_{ij,k}\underline{y}^{i}\underline{y}^{j}y^{k},\quad\text{ where }Q^{\mu}_{ij,k}=Q^{\mu}_{ji,k},\end{cases} (2.21)

for some structure functions QQ^{\cdots}_{\cdots}. The condition Qji=δjiQ^{i}_{j}=\delta^{i}_{j} corresponds to (2.18) and it ensures that the order-one reduction of κ2\kappa^{2} gives a (general) algebroid structure on A=E1A=E^{1}. If (E2,κ2)(E^{2},\kappa^{2}) is a skew HA then Q~ia=Qia\tilde{Q}^{a}_{i}=Q^{a}_{i} and Qjki=QkjiQ^{i}_{jk}=-Q^{i}_{kj} since κ1\kappa^{1} is a symmetric relation. The structure functions satisfy certain equations reflecting the axioms of a higher-order algebroid. These equations are derived in Appendix, Subsection 4.3.

Example 2.9.

The natural diffeomorphism κMk:TkTMTTkM\kappa^{k}_{M}:\mathrm{T}^{k}\mathrm{T}M\rightarrow\mathrm{T}\mathrm{T}^{k}M defines a Lie, order kk HA on τMk:TkMM\tau^{k}_{M}:\mathrm{T}^{k}M\rightarrow M. Indeed, (TkM,κMk)(\mathrm{T}^{k}M,\kappa^{k}_{M}) satisfies the Lie condition (Definition 2.7(iii)) because εMk:=(κMk):TTkMTkTM\varepsilon^{k}_{M}:=(\kappa^{k}_{M})^{\ast}:\mathrm{T}^{\ast}\mathrm{T}^{k}M\rightarrow\mathrm{T}^{k}\mathrm{T}^{\ast}M is a Poisson map. It also comes from a more general result, see [JR18, Proposition 4.13].

2.7 Reformulation of the definition of a HA in terms of algebroid lifts

We shall review the construction of higher lifts s(α)s^{({\alpha})} of sections of a vector bundle. This notion is used in various parts of this work, such as in the definition of algebroid lifts sαk𝔛(Ek){s}^{\langle{{\alpha}-k}\rangle}\in\mathfrak{X}(E^{k}) (see (2.29)), which facilitate the convenient description of the axioms of HAs (see Theorem 2.11).

Fix kk\in\mathbb{N} and let sΓ(σ)s\in\operatorname{\Gamma}(\sigma) be a section of a vector bundle σ:EM\sigma:E\to M. We can interpret ss as a linear function ι(s)\iota(s) on EE^{\ast}, the linear dual of EE. Let 0αk0\leq{\alpha}\leq k. Then the (α)({\alpha})-lift of ι(s)\iota(s) is a function on TkE\mathrm{T}^{k}E^{\ast}, commuting with htTkE=Tk(htE)h^{\mathrm{T}^{k}E^{\ast}}_{t}=\mathrm{T}^{k}(h^{E^{\ast}}_{t}), the homogeneity structure on TkE\mathrm{T}^{k}E^{\ast}. Therefore, ι(s)(α)\iota(s)^{({\alpha})} can be interpreted as a section of the linear dual of the vector bundle Tkσ:TkETkM\mathrm{T}^{k}\sigma^{\ast}:\mathrm{T}^{k}E^{\ast}\rightarrow\mathrm{T}^{k}M, which is identified with the vector bundle Tkσ:TkETkM\mathrm{T}^{k}\sigma:\mathrm{T}^{k}E\rightarrow\mathrm{T}^{k}M via the non-degenerate pairing

,Tkσ:TkE×TkMTkETk(E×ME),σ(k),\langle{\cdot,\cdot}\rangle_{\mathrm{T}^{k}\sigma}:\mathrm{T}^{k}E^{\ast}\times_{\mathrm{T}^{k}M}\mathrm{T}^{k}E\simeq\mathrm{T}^{k}(E^{\ast}\times_{M}E)\xrightarrow{\langle{\cdot,\cdot}\rangle_{\sigma}^{(k)}}\mathbb{R}, (2.22)

obtained as (k)(k)-lift of the pairing ,σ:E×ME\langle{\cdot,\cdot}\rangle_{\sigma}:E^{\ast}\times_{M}E\rightarrow\mathbb{R}. The section of Tkσ\mathrm{T}^{k}\sigma obtained this way is denoted by s(α)s^{(\alpha)} and called the (α)({\alpha})-lift of the section ss. In standard coordinates (xa,yi)(x^{a},y^{i}) on EE, and (xa,ξi)(x^{a},\xi_{i}) on EE^{\ast}, where ξi=ι(ei)\xi_{i}=\iota(e_{i}), the (k)(k)-lift of the function ,σ=yiξi\langle{\cdot,\cdot}\rangle_{\sigma}=y^{i}\xi_{i} is obtained using the general Leibniz rule, and has the form

(xa,(α),ξi(β)),(xa,(α),yi,(β))Tkσ=α=0k(kα)ξi(α)yi,(kα).\langle{(x^{a,({\alpha})},\xi_{i}^{({\beta})}),(x^{a,({\alpha})},y^{i,({\beta})})}\rangle_{\mathrm{T}^{k}\sigma}=\sum_{{\alpha}=0}^{k}\binom{k}{{\alpha}}\xi_{i}^{({\alpha})}y^{i,(k-{\alpha})}.

It follows that the family (ei(α))(e_{i}^{({\alpha})}), where 0αk0\leq\alpha\leq k and ι(ei(α))=ξi(α)\iota(e_{i}^{({\alpha})})=\xi_{i}^{({\alpha})}, forms a local frame of sections of the vector bundle Tkσ\mathrm{T}^{k}\sigma. Moreover, yj,(β),ei(kα)Tkσ=δjiδβα(kα)\langle{y^{j,(\beta)},e_{i}^{(k-\alpha)}}\rangle_{\mathrm{T}^{k}\sigma}=\delta^{i}_{j}\delta^{\alpha}_{\beta}\binom{k}{\alpha}, hence

yi,(α)ei(kα)=(kα)1,y^{i,({\alpha})}\circ e_{i}^{(k-{\alpha})}=\binom{k}{{\alpha}}^{-1}, (2.23)

as the composition of functions ei(kα):TkMTkEe_{i}^{(k-{\alpha})}:\mathrm{T}^{k}M\rightarrow\mathrm{T}^{k}E and yi,(α):TkEy^{i,({\alpha})}:T^{k}E\rightarrow\mathbb{R}. From this it is straightforward to verify that this construction of s(α)s^{({\alpha})} is equivalent to the one presented in [JR18]. We have

(fs)(β)=α=0β(βα)f(α)s(βα)(f\cdot s)^{(\beta)}=\sum_{\alpha=0}^{\beta}\binom{\beta}{\alpha}f^{(\alpha)}s^{(\beta-\alpha)} (2.24)

for β=0,1,,k\beta=0,1,\ldots,k, f𝒞(M)f\in\mathcal{C}^{\infty}(M) and sΓ(E)s\in\operatorname{\Gamma}(E). This is simply the Leibniz rule for the iterated derivative.

Definition 2.10 (Vertical lifts).

Let 0αk0\leq\alpha\leq k. We define a VB comorphism 𝐕αk\mathbf{V}^{k}_{\alpha},

TαE\textstyle{\mathrm{T}^{\alpha}E\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐕αk\scriptstyle{\mathbf{V}^{k}_{\alpha}}TkE\textstyle{\mathrm{T}^{k}E\ignorespaces\ignorespaces\ignorespaces\ignorespaces}TαM\textstyle{\mathrm{T}^{\alpha}M}TkM\textstyle{\mathrm{T}^{k}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ταk\scriptstyle{\tau^{k}_{\alpha}} (2.25)

covering the natural projection ταk:TkMTαM\tau^{k}_{\alpha}:\mathrm{T}^{k}M\rightarrow\mathrm{T}^{\alpha}M by

(𝐕αk)[γ]k([a]α)=[tα!k!tkαa(t)]k\left(\mathbf{V}^{k}_{\alpha}\right)_{[\gamma]_{k}}([a]_{\alpha})=[t\mapsto\frac{\alpha!}{k!}t^{k-\alpha}a(t)]_{k} (2.26)

where γ\gamma is a curve in MM and aa is a curve in EE such that [a¯]k=[γ]k[\underline{a}]_{k}=[\gamma]_{k} where a¯=σa\underline{a}=\sigma\circ a.

Note that for α=k1\alpha=k-1 we recover the map (2.5). It is clear that (2.26) does not depend on the choice of representatives γ\gamma and aa.

The (α)(\alpha)-lift s(α)s^{(\alpha)} of the section sΓ(E)s\in\operatorname{\Gamma}(E) can be presented as the composition of the complete lift Tαs\mathrm{T}^{\alpha}s with the vertical lift 𝐕αk{\overrightarrow{\mathbf{V}}}^{k}_{\alpha}:

s(α)=𝐕αk(Tαs).s^{(\alpha)}={\overrightarrow{\mathbf{V}}}^{k}_{\alpha}(\mathrm{T}^{\alpha}s). (2.27)

A simple coordinate-based proof is left to the reader.

A (general) algebroid structure κ\kappa on the vector bundle σ:EM\sigma:E\rightarrow M can be lifted by means of kthk^{\mathrm{th}}-tangent functor to the vector bundle Tkσ:TkETkM\mathrm{T}^{k}\sigma:\mathrm{T}^{k}E\rightarrow\mathrm{T}^{k}M (see [JR18]). The lifted structure is called kthk^{\mathrm{th}}-order tangent lift of (σ,κ)(\sigma,\kappa) and denoted as (Tkσ,dTkκ)(\mathrm{T}^{k}\sigma,\mathrm{d}_{\mathrm{T}^{k}}\kappa). The algebroid bracket [,]Tkσ[\cdot,\cdot]_{\mathrm{T}^{k}\sigma} on Tkσ\mathrm{T}^{k}\sigma satisfies

[k!(kα)!s1(kα),k!(kβ)!s2(kβ)]Tkσ=k!(kαβ)!([s1,s2]σ)(kαβ),[\frac{k!}{(k-\alpha)!}s_{1}^{(k-\alpha)},\frac{k!}{(k-\beta)!}s_{2}^{(k-\beta)}]_{\mathrm{T}^{k}\sigma}=\frac{k!}{(k-\alpha-\beta)!}([s_{1},s_{2}]_{\sigma})^{(k-\alpha-\beta)}\ , (2.28)

for any integers α,β=0,1,,k\alpha,\beta=0,1,\ldots,k such that α+βk\alpha+\beta\leq k, and any sections s1,s2Γ(E)s_{1},s_{2}\in\operatorname{\Gamma}(E). Additionally, [s1(kα),s2(kβ)]Tkσ=0[s_{1}^{(k-\alpha)},s_{2}^{(k-\beta)}]_{\mathrm{T}^{k}\sigma}=0 if α+β>k\alpha+\beta>k. Moreover, if (σ,κ)(\sigma,\kappa) is a skew/AL/Lie algebroid, then so is (Tkσ,dTkκ)(\mathrm{T}^{k}\sigma,\mathrm{d}_{\mathrm{T}^{k}}\kappa).

Assume that (σ,κ)(\sigma,\kappa) is Lie. From (2.28), we observe that assigning the weight αβ\alpha-\beta to a section of the form f(α)s(kβ)f^{(\alpha)}s^{(k-\beta)}, where f𝒞(M)f\in\mathcal{C}^{\infty}(M) and sΓ(E)s\in\operatorname{\Gamma}(E), turns the Lie subalgebra of Γ(Tkσ)\operatorname{\Gamma}(\mathrm{T}^{k}\sigma) generated by homogeneous sections into a graded Lie algebra concentrated in weights k\geq-k. This Lie algebra has a Lie subalgebra Γ0(Tkσ)\operatorname{\Gamma}_{\leq 0}(\mathrm{T}^{k}\sigma) generated by homogeneous sections of non-positive weights. It is of finite rank over 𝒞(M)\mathcal{C}^{\infty}(M).

Using the structure of a higher algebroid on a graded bundle σk:EkM\sigma^{k}:E^{k}\rightarrow M one can define algebroid lifts of a section sΓM(E1)s\in\operatorname{\Gamma}_{M}(E^{1}) as follows:

sα:=k!(kα)!κk(s(kα))𝔛α(Ek),kα0.{s}^{\langle{-\alpha}\rangle}:=\frac{k!}{(k-\alpha)!}{\overrightarrow{\kappa^{k}}}(s^{(k-\alpha)})\in\mathfrak{X}_{-\alpha}(E^{k}),\quad{-k\leq-\alpha\leq 0.} (2.29)

The notation is slightly different from that in [JR18] where the algebroid (kα)(k-\alpha)-lift of a section ss was denoted by s[kα]{s}^{[k-\alpha]} and it is related as sα=k!(kα)!s[kα]{s}^{\langle{-\alpha}\rangle}=\frac{k!}{(k-\alpha)!}{s}^{[k-\alpha]}. Thanks to this correction, the vector field sα{s}^{\langle{-\alpha}\rangle} has weight α-\alpha and the equation (4.6) in [JR18] simplifies to

[s1α,s2β]=[s1,s2]α+β[{s_{1}}^{\langle{{\alpha}}\rangle},{s_{2}}^{\langle{{\beta}}\rangle}]={[s_{1},s_{2}]}^{\langle{{\alpha}+{\beta}}\rangle} (2.30)

for any s1,s2Γ(E1)s_{1},s_{2}\in\operatorname{\Gamma}(E^{1}) and α,β0{\alpha},{\beta}\leq 0 such that kα+β-k\leq{\alpha}+{\beta}.

Using (2.24) we get

(fs)α=β=0kα1β!(k)f(β)sαβ.{(fs)}^{\langle{-\alpha}\rangle}=\sum_{\beta=0}^{k-\alpha}\frac{1}{\beta!}(\sharp^{k})^{\ast}f^{(\beta)}\,{s}^{\langle{-\alpha-\beta}\rangle}. (2.31)

In particular,

(fs)k=fsk.{(fs)}^{\langle{-k}\rangle}=f\,{s}^{\langle{-k}\rangle}. (2.32)

Any vector field X𝔛0(Ek)X\in\mathfrak{X}_{0}(E^{k}) of weight 0 has a form

X=Xa(x)xa+iXi(x,y)ywi,X=X^{a}(x)\partial_{x^{a}}+\sum_{i}X^{i}(x,y)\partial_{y^{i}_{w}},

and has a well defined projection on MM, denoted by X0k=Xa(x)xa𝔛(M)X\!\!\!\!\downarrow^{k}_{0}=X^{a}(x)\partial_{x^{a}}\in\mathfrak{X}(M). Similarly, a vector field Y𝔛1(Ek)Y\in\mathfrak{X}_{-1}(E^{k}) of weight 1-1 is projectable onto E1E^{1}, the projection is denoted by Y1k𝔛1(E1)Γ(E1)Y\!\!\!\!\downarrow^{k}_{1}\in\mathfrak{X}_{-1}(E^{1})\simeq\operatorname{\Gamma}(E^{1}), see Lemma 4.1. Below is a reformulation of axioms of higher-order algebroids in terms of algebroid lifts.

Theorem 2.11.

Let σk:EkM\sigma^{k}:E^{k}\rightarrow M be a graded bundle of order kk.

  1. (i)

    Assume that the order-one reduction of σk\sigma^{k} is a trivial VB of rank nn, i.e., it admits a trivialization E1M×nE^{1}\simeq M\times\mathbb{R}^{n}, and let (ei)i=1,,n(e_{i})_{i=1,\ldots,n} be the corresponding frame of σ1:E1M\sigma^{1}:E^{1}\rightarrow M. A general HA is provided by a graded bundle morphism k:EkTkM\sharp^{k}:E^{k}\rightarrow\mathrm{T}^{k}M and a collection of homogeneous vector fields Xi,α𝔛α(Ek)X_{i,{{\alpha}}}\in\mathfrak{X}_{{\alpha}}(E^{k}), where kα0-k\leq{\alpha}\leq 0, 1in1\leq i\leq n, such that the projection of each vector field Xi,1𝔛1(Ek)X_{i,-1}\in\mathfrak{X}_{-1}(E^{k}) onto E1E^{1} coincides with eiΓ(E1)𝔛1(E1)e_{i}\in\operatorname{\Gamma}(E^{1})\simeq\mathfrak{X}_{-1}(E^{1}). Moreover, the vector fields which define κ1\kappa^{1} – the order-one reduction of κk\kappa^{k}, are the projections of Xi,0X_{i,0} and Xi,1X_{i,-1} onto E1E^{1}.

  2. (ii)

    A skew HA (Ek,κk)(E^{k},\kappa^{k}) is almost Lie if and only if for any section sΓ(E1)s\in\operatorname{\Gamma}(E^{1}) and kα0-k\leq{\alpha}\leq 0 the vector fields sα𝔛α(Ek){s}^{\langle{{\alpha}}\rangle}\in\mathfrak{X}_{{\alpha}}(E^{k}) and (1s)α𝔛α(TkM){\left(\sharp^{1}s\right)}^{\langle{{\alpha}}\rangle}\in\mathfrak{X}_{{\alpha}}(\mathrm{T}^{k}M) are k\sharp^{k}-related.

  3. (iii)

    [JR18, Proposition 4.9] An almost Lie HA (σk,κk)(\sigma^{k},\kappa^{k}) is Lie if and only if

    κk|Γ0(Tkσ1):Γ0(Tkσ1)𝔛0(Ek){\overrightarrow{\kappa^{k}}}|_{\operatorname{\Gamma}_{\leq 0}(\mathrm{T}^{k}\sigma^{1})}:\operatorname{\Gamma}_{\leq 0}(\mathrm{T}^{k}\sigma^{1})\rightarrow\mathfrak{X}_{\leq 0}(E^{k})

    is a Lie algebra homomorphism.

Proof.
  1. (i)

    The sections (ei(α))(e_{i}^{({\alpha})}) form a frame for TkE1TkM\mathrm{T}^{k}E^{1}\rightarrow\mathrm{T}^{k}M, hence their pullbacks ((k)ei(α))\left((\sharp^{k})^{\ast}e_{i}^{({\alpha})}\right) form a frame for the pullback vector bundle (k)Tkσ:TkE1×(Tkσ,k)EkEk(\sharp^{k})^{\ast}\mathrm{T}^{k}\sigma:\mathrm{T}^{k}E^{1}\times_{(\mathrm{T}^{k}\sigma,\sharp^{k})}E^{k}\rightarrow E^{k}. To set a comorphism κk:TkστEk\kappa^{k}:\mathrm{T}^{k}\sigma{\rightarrow\!\!\vartriangleright}\tau_{E^{k}}, this amounts to defining a VB morphism from the VB (k)Tkσ(\sharp^{k})^{\ast}\mathrm{T}^{k}\sigma to the tangent bundle of EkE^{k}, covering the identity idEk\operatorname{id}_{E^{k}}. This is done by assigning vector fields to the sections from the local frame. We send (k)ei(α)(\sharp^{k})^{\ast}e_{i}^{({\alpha})} to Xi,αX_{i,{\alpha}}. In other words, κk(ei(α))=Xi,α{\overrightarrow{\kappa^{k}}}(e_{i}^{({\alpha})})=X_{i,{\alpha}}. Then the obtained comorphism κk\kappa^{k} is weighted, as the vector fields Xi,αX_{i,{\alpha}} are homogeneous and k\sharp^{k} preserves the weight.

    The condition (2.18) corresponds to the fact that eiΓ(E1)𝔛1(E1)e_{i}\in\operatorname{\Gamma}(E^{1})\simeq\mathfrak{X}_{-1}(E^{1}) coincides with the projection of Xi,1X_{i,1} onto E1E^{1}.

  2. (ii)

    The commutativity of the diagram (2.20), corresponding to the almost Lie axiom, can be reformulated as follows: For any section sΓ(Tkσ1)s\in\operatorname{\Gamma}(\mathrm{T}^{k}\sigma^{1}), the vector fields κk(s)𝔛(Ek){\overrightarrow{\kappa^{k}}}(s)\in\mathfrak{X}(E^{k}) and κMkTkσ1(s)𝔛(TkM)\kappa^{k}_{M}\circ\mathrm{T}^{k}\sigma^{1}(s)\in\mathfrak{X}(\mathrm{T}^{k}M) are k\sharp^{k}-related (see the proof of [JR18, Proposition 4.9]). In particular, in any AL HA (Ek,κk)(E^{k},\kappa^{k}), for any section sΓ(E1)s\in\operatorname{\Gamma}(E^{1}) and αk{\alpha}\geq-k, the vector fields sα𝔛(Ek){s}^{\langle{{\alpha}}\rangle}\in\mathfrak{X}(E^{k}) and (s)α𝔛(TkM){(\sharp s)}^{\langle{{\alpha}}\rangle}\in\mathfrak{X}(\mathrm{T}^{k}M) are k\sharp^{k}-related. (The latter are algebroid lifts with respect to the HA structure on τMk:TkMM\tau^{k}_{M}:\mathrm{T}^{k}M\rightarrow M.) On the other hand, if f𝒞(TkM)f\in\mathcal{C}^{\infty}(\mathrm{T}^{k}M) and sΓ(Tkσ1)s\in\operatorname{\Gamma}(\mathrm{T}^{k}\sigma^{1}) then

    κk(fs)=(k)(f)κk(s),κMkTk1(fs)=fκMkTk1(s).{\overrightarrow{\kappa^{k}}}(f\cdot s)=(\sharp^{k})^{\ast}(f)\cdot{\overrightarrow{\kappa^{k}}}(s),\quad\kappa^{k}_{M}\circ\mathrm{T}^{k}\sharp^{1}(f\cdot s)=f\cdot\kappa^{k}_{M}\circ\mathrm{T}^{k}\sharp^{1}(s).

    Therefore, if the vector fields κk(s){\overrightarrow{\kappa^{k}}}(s) and κMkTk1(s)\kappa^{k}_{M}\circ\mathrm{T}^{k}\sharp^{1}(s) are k\sharp^{k}-related, then the same is true if we replace ss with fsf\cdot s. Hence, the thesis (ii) holds since sections of the form s(kα)s^{(k-{\alpha})}, where 0αk0\leq{\alpha}\leq k, span Γ(Tkσ1)\operatorname{\Gamma}(\mathrm{T}^{k}\sigma^{1}) as 𝒞(TkM)\mathcal{C}^{\infty}(\mathrm{T}^{k}M)-module. The proof of (iii) is presented in [JR18].

Remark 2.12.

It suffices to verify the conditions given in Theorem 2.11 locally. Moreover, it is sufficient to take the sections of the vector bundle σ1:E1M\sigma^{1}:E^{1}\rightarrow M to be the elements of a frame (ei)(e_{i}) of local sections. In this way, the almost Lie axiom and Lie axiom can be reduced (locally) to a finite number of equations:

  1. (AL axiom)

    The vector fields eiα{e_{i}}^{\langle{-{\alpha}}\rangle} and (ei)α{(\sharp e_{i})}^{\langle{-{\alpha}}\rangle} are k\sharp^{k}-related for any 0αk0\leq{\alpha}\leq k.

  2. (Lie axiom)

    [eiα,ejβ]τEk=[ei,ej]σ1αβ[{e_{i}}^{\langle{-{\alpha}}\rangle},{e_{j}}^{\langle{-\beta}\rangle}]_{\tau_{E^{k}}}={[e_{i},e_{j}]_{\sigma^{1}}}^{\langle{-{\alpha}-\beta}\rangle} for any 0α,β0\leq{\alpha},\beta such that α+βk\alpha+\beta\leq k.

Remark 2.13.

There is also a dual construction of the algebroid lifts sα{s}^{\langle{-{\alpha}}\rangle} associated with a HA (Ek,κk)(E^{k},\kappa^{k}), which coincides with the construction presented in [GU99] for Lie algebroids, i.e., when k=1k=1. Given a section sΓ(E)s\in\operatorname{\Gamma}(E) considered as a linear function ι(s)\iota(s) on EE^{\ast}, we have (α)({\alpha})-lifts ι(s)(α)𝒞(TkE)\iota(s)^{({\alpha})}\in\mathcal{C}^{\infty}(\mathrm{T}^{k}E^{\ast}) for α=0,1,,k{\alpha}=0,1,\ldots,k. As we mentioned (see (2.22)), the vector bundles Tkσ:TkETkM\mathrm{T}^{k}\sigma:\mathrm{T}^{k}E\rightarrow\mathrm{T}^{k}M and Tkσ:TkETkM\mathrm{T}^{k}\sigma^{\ast}:\mathrm{T}^{k}E^{\ast}\rightarrow\mathrm{T}^{k}M are in natural duality, hence the dual of κk\kappa^{k} is a weighted vector bundle morphism εk\varepsilon^{k} of the form

TEk\textstyle{\mathrm{T}^{\ast}E^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τEk\scriptstyle{\tau^{\ast}_{E^{k}}}εk\scriptstyle{\varepsilon^{k}}TkE\textstyle{\mathrm{T}^{k}E^{\ast}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tkσ\scriptstyle{\mathrm{T}^{k}\sigma^{\ast}}Ek\textstyle{E^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}k\scriptstyle{\sharp^{k}}TkM\textstyle{\mathrm{T}^{k}M}

By pulling back ι(s)(α)\iota(s)^{({\alpha})} via εk\varepsilon^{k} we obtain linear functions on TEk\mathrm{T}^{\ast}E^{k}, thus vector fields on EkE^{k}. It is evident (by working fiberwise) that this way we recover our algebroid lifts, i.e,

sα,τEk=(εk)ι(s)(kα)\langle{s}^{\langle{-{\alpha}}\rangle},\cdot\rangle_{\tau_{E^{k}}}=(\varepsilon^{k})^{\ast}\iota(s)^{(k-{\alpha})}

Let (Ek,κk)(E^{k},\kappa^{k}) be a HA and (Ej,κj)(E^{j},\kappa^{j}) be its reduction to order jj, where 1j<k1\leq j<k. The following lemma states that algebroid lifts sακk{s}^{\langle\alpha\rangle_{\kappa^{k}}} and sακj{s}^{\langle\alpha\rangle_{\kappa^{j}}} obtained using κk\kappa^{k} and κj\kappa^{j}, respectively, are compatible in some natural sense.

Lemma 2.14.

Let sΓ(E1)s\in\operatorname{\Gamma}(E^{1}) and 0αj<k0\leq\alpha\leq j<k. Then the vector field sακk𝔛(Ek){s}^{\langle{-}\alpha\rangle_{\kappa^{k}}}\in\mathfrak{X}(E^{k}) is projectable onto EjE^{j} and its projection is sακj{s}^{\langle-\alpha\rangle_{\kappa^{j}}}.

Proof.

We shall use the construction of (α)({\alpha})-lifts of a section ss, as defined in (2.27). We have

sακk=k!(kα)!κk(𝐕kαk(Tkαs))=κk(ξαk),{s}^{\langle{-\alpha}\rangle_{\kappa^{k}}}=\frac{k!}{(k-\alpha)!}\,{\overrightarrow{\kappa^{k}}}\left({\overrightarrow{\mathbf{V}}}^{k}_{k-\alpha}(\mathrm{T}^{k-\alpha}s)\right)={\overrightarrow{\kappa^{k}}}(\xi^{k}_{\alpha}),

where the section ξαk:TkMTkE1\xi^{k}_{\alpha}:\mathrm{T}^{k}M\rightarrow\mathrm{T}^{k}E^{1} is defined by [γ]k[ttαs(γ(t))]k[\gamma]_{k}\mapsto[t\mapsto t^{\alpha}\cdot s(\gamma(t))]_{k} where γ\gamma is a curve in MM. Hence, the vector field sακk{s}^{\langle{-\alpha}\rangle_{\kappa^{k}}} is the composition of maps idEk×(ξαkk):EkEk×TkMTkE1\operatorname{id}_{E^{k}}\times(\xi^{k}_{\alpha}\circ\sharp^{k}):E^{k}\rightarrow E^{k}\times_{\mathrm{T}^{k}M}\mathrm{T}^{k}E^{1} with the VB morphism (κk)!:Ek×TkMTkE1TEk(\kappa^{k})^{!}:E^{k}\times_{\mathrm{T}^{k}M}\mathrm{T}^{k}E^{1}\rightarrow\mathrm{T}E^{k} induced by κk\kappa^{k}. The thesis follows from the commutativity of the diagram

Ek\textstyle{E^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σjk\scriptstyle{\sigma^{k}_{j}}idEk×(ξαkk)\scriptstyle{\operatorname{id}_{E^{k}}\times(\xi^{k}_{\alpha}\circ\sharp^{k})}Ek×TkMTkE1\textstyle{E^{k}\times_{\mathrm{T}^{k}M}\mathrm{T}^{k}E^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(κk)!\scriptstyle{(\kappa^{k})^{!}}TEk\textstyle{\mathrm{T}E^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tσjk\scriptstyle{\mathrm{T}\sigma^{k}_{j}}Ej\textstyle{E^{j}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idEj×(ξαjj)\scriptstyle{\operatorname{id}_{E^{j}}\times(\xi^{j}_{\alpha}\circ\sharp^{j})}Ej×TjMTjE1\textstyle{E^{j}\times_{\mathrm{T}^{j}M}\mathrm{T}^{j}E^{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(κj)!\scriptstyle{(\kappa^{j})^{!}}TEj\textstyle{\mathrm{T}E^{j}}

2.8 Prolongations of an almost Lie algebroid

Let 𝒢\mathcal{G} be a Lie groupoid with source and target maps denoted by α,β:𝒢M\alpha,\beta:\mathcal{G}\rightarrow M, respectively. We consider a foliation 𝒢α\mathcal{G}^{\alpha} on 𝒢\mathcal{G} defined by α\alpha-fibers 𝒢xα={g𝒢:α(g)=x}\mathcal{G}_{x}^{\alpha}=\{g\in\mathcal{G}:\alpha(g)=x\}, the distribution T𝒢αT𝒢\mathrm{T}\mathcal{G}^{\alpha}\subset\mathrm{T}\mathcal{G} tangent to the leaves of 𝒢α\mathcal{G}^{\alpha}, related objects like Tk𝒢α\mathrm{T}^{k}\mathcal{G}^{\alpha} and the right action of 𝒢\mathcal{G} on itself, Rg:hhgR_{g}:h\mapsto hg where h𝒢β(g)αh\in\mathcal{G}^{\alpha}_{\beta(g)}. The Lie algebroid of 𝒢\mathcal{G} is usually defined as the vector bundle σ:𝒜(𝒢):=TM𝒢αM\sigma:\mathcal{A}(\mathcal{G}):=\mathrm{T}_{M}\mathcal{G}^{\alpha}\rightarrow M equipped with a map :𝒜(𝒢)TM\sharp:\mathcal{A}(\mathcal{G})\rightarrow\mathrm{T}M called the anchor, defined as =Tβ|𝒜(𝒢)\sharp=\mathrm{T}\beta|_{\mathcal{A}(\mathcal{G})} and the Lie bracket on Γ(𝒜(𝒢))\operatorname{\Gamma}(\mathcal{A}(\mathcal{G})) inherited from the Lie bracket of right-invariant vector fields on 𝒢\mathcal{G}. (Such vector fields are in a one-to-one correspondence with sections of σ\sigma.) Another, yet equivalent construction of the Lie algebroid structure on 𝒜(𝒢)\mathcal{A}(\mathcal{G}), is provided by the reduction map 1\mathcal{R}^{1},

T𝒢α\textstyle{\mathrm{T}\mathcal{G}^{\alpha}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}1\scriptstyle{\mathcal{R}^{1}}τ𝒢\scriptstyle{\tau_{\mathcal{G}}}𝒜(G)\textstyle{\mathcal{A}(G)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ\scriptstyle{\tau}𝒢\textstyle{\mathcal{G}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}β\scriptstyle{\beta}M,\textstyle{M,} (2.33)

which is a fiber-wise VB isomorphism obtained from the collection of maps TRg1:Tg𝒢αTβ(g)𝒢α\mathrm{T}R_{g^{-1}}:\mathrm{T}_{g}\mathcal{G}^{\alpha}\rightarrow\mathrm{T}_{\beta(g)}\mathcal{G}^{\alpha}. The Lie algebroid structure on 𝒜(𝒢)\mathcal{A}(\mathcal{G}) is defined by means of the VB comorphism κ:T𝒜(𝒢)T𝒜(𝒢)\kappa:\mathrm{T}\mathcal{A}(\mathcal{G}){\rightarrow\!\!\vartriangleright}\mathrm{T}\mathcal{A}(\mathcal{G}) which is obtained as the reduction of κ𝒢:TT𝒢TT𝒢\kappa_{\mathcal{G}}:\mathrm{T}\mathrm{T}\mathcal{G}\rightarrow\mathrm{T}\mathrm{T}\mathcal{G}. The advantage of the latter over the standard construction of the Lie functor is that it can be easily generalized to higher orders. This is obtained by means of the higher-order reduction map k:Tk𝒢αTMk𝒢α\mathcal{R}^{k}:\mathrm{T}^{k}\mathcal{G}^{\alpha}\rightarrow\mathrm{T}^{k}_{M}\mathcal{G}^{\alpha}, defined analogously to 1\mathcal{R}^{1}, by the collection of maps TkRg1:Tgk𝒢αTβ(g)k𝒢α\mathrm{T}^{k}R_{g^{-1}}:\mathrm{T}^{k}_{g}\mathcal{G}^{\alpha}\rightarrow\mathrm{T}^{k}_{\beta(g)}\mathcal{G}^{\alpha}.

Definition 2.15.

[JR15, Definition 3.3, Lemma 3.4] The kthk^{\mathrm{th}}-order Lie algebroid of a Lie groupoid 𝒢\mathcal{G} is the graded bundle 𝒜k(𝒢):=TMk𝒢α\mathcal{A}^{k}(\mathcal{G}):=\mathrm{T}^{k}_{M}\mathcal{G}^{\alpha} together with a VB comorphism κk:=(Tk1,Tk)(κ~𝒢k)\kappa^{k}:=(\mathrm{T}^{k}\mathcal{R}^{1},\mathrm{T}\mathcal{R}^{k})(\widetilde{\kappa}^{k}_{\mathcal{G}}) where κ~𝒢k\widetilde{\kappa}^{k}_{\mathcal{G}} is the restriction of κ𝒢k\kappa^{k}_{\mathcal{G}} to (TkT)𝒢α×(TTk)𝒢α(\mathrm{T}^{k}\mathrm{T})\mathcal{G}^{\alpha}\times(\mathrm{T}\mathrm{T}^{k})\mathcal{G}^{\alpha} subject to the natural inclusions (TkT)𝒢αTk(T𝒢α)(\mathrm{T}^{k}\mathrm{T})\mathcal{G}^{\alpha}\subset\mathrm{T}^{k}(\mathrm{T}\mathcal{G}^{\alpha}) and (TTk)𝒢αT(Tk𝒢α)(\mathrm{T}\mathrm{T}^{k})\mathcal{G}^{\alpha}\subset\mathrm{T}(\mathrm{T}^{k}\mathcal{G}^{\alpha}).

Actually, (𝒜k(𝒢),κk)(\mathcal{A}^{k}(\mathcal{G}),\kappa^{k}) is a Lie HA in the sense of Definition 2.7 ([JR18, Proposition 4.13] and [JR15, Section 5]).

In [JR15] we introduced a slightly bigger class of examples of HAs obtained by means of the construction called the prolongation of an almost Lie algebroid (A,κ)(A,\kappa). We will outline this construction, highlighting a possible more general context for certain constructions.

A pair of a vector bundle σ:AM\sigma:A\rightarrow M and a VB morphism :ATM\sharp:A\to\mathrm{T}M covering the identity idM\operatorname{id}_{M} is called an anchored vector bundle. A curve a:Aa:\mathbb{R}\rightarrow A is called admissible if the tangent lift of the curve a¯=σa:M\underline{a}=\sigma\circ a:\mathbb{R}\rightarrow M coincides with the curve a\sharp\circ a, i.e., 𝐭a¯=a\bm{\mathrm{t}}\underline{a}=\sharp\circ a. The subset A[k]A^{[k]} of Tk1A\mathrm{T}^{k-1}A, defined as

A[k]={[a]k1a:A is an admissible curve },A^{[k]}=\{[a]_{k-1}\mid a:\mathbb{R}\to A\text{ is an admissible curve }\}, (2.34)

is called the kthk^{\mathrm{th}}-order prolongation of the anchored vector bundle AA (see [Pop04]). According to [BGG15b, Theorem 2.2.7], we have

A[2]={XTA:(Tσ)X=τA(X)},A^{[2]}=\{X\in\mathrm{T}A:(\mathrm{T}\sigma)X=\sharp\,\tau_{A}(X)\}, (2.35)

and

A[k]=(Tk1)1(TkM)A^{[k]}=(\mathrm{T}^{k-1}\sharp)^{-1}(\mathrm{T}^{k}M)

where TkM\mathrm{T}^{k}M is considered as a subset of Tk1TM\mathrm{T}^{k-1}\mathrm{T}M via iMk1,1i^{k-1,1}_{M}. It follows that the constructions of the sets A[k]A^{[k]} here and EkE^{k} in [JR15, Definition 4.1] are equivalent, see also [BGG15b] or [JR15, Theorem 4.5 (viii)], or [Mar15]. In particular, A[k+1]=TA[k]TkAA^{[k+1]}=\mathrm{T}A^{[k]}\cap\mathrm{T}^{k}A, considered as subsets of TTk1A\mathrm{T}\mathrm{T}^{k-1}A.

Define kthk^{\mathrm{th}}-order anchor map [k]:A[k]TkM\sharp^{[k]}:A^{[k]}\rightarrow\mathrm{T}^{k}M as [k]=(Tk1)|A[k]\sharp^{[k]}=(\mathrm{T}^{k-1}\sharp)|_{A^{[k]}}. It is a graded bundle morphism.

An AL algebroid structure on the vector bundle AA can be prolonged to a HA structure on A[k]A^{[k]} by means of the comorphisms κ[k]:TkATA[k]\kappa^{[k]}:\mathrm{T}^{k}A{\rightarrow\!\!\vartriangleright}\mathrm{T}A^{[k]}, covering [k]\sharp^{[k]}, defined as

κ[k]=(κAk1Tk1κ)(TkATA[k]),\kappa^{[k]}=(\kappa_{A}^{k-1}\circ\mathrm{T}^{k-1}\kappa)\cap(\mathrm{T}^{k}A\cap\mathrm{T}A^{[k]}), (2.36)

see [JR15, Proposition 4.6]. The comorphism κ[k]\kappa^{[k]} can also be defined inductively as it is presented in [JR15, Definition 4.2].

2.9 Canonical inclusions II

Here, we highlight some natural embeddings induced by the anchored bundle structure on a vector bundle AMA\rightarrow M.

In addition to the inclusion A[k]Tk1AA^{[k]}\subseteq\mathrm{T}^{k-1}A from the definition of A[k]A^{[k]} (2.34), there are inclusions ıAk,l:A[k+l]TkA[l]\imath^{k,l}_{A}:A^{[k+l]}\to\mathrm{T}^{k}A^{[l]} defined by the restriction of iAk,l1i^{k,l-1}_{A} to A[k+l]A^{[k+l]}:

Tk+l1A\textstyle{\mathrm{T}^{k+l-1}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}iAk,l1\scriptstyle{i^{k,l-1}_{A}}TkTl1A\textstyle{\mathrm{T}^{k}\mathrm{T}^{l-1}A}A[k+l]\textstyle{A^{[k+l]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ıAk,l\scriptstyle{\imath^{k,l}_{A}}TkA[l]\textstyle{\mathrm{T}^{k}A^{[l]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

We should prove that the image ıAk,l(A[k+l])\imath^{k,l}_{A}(A^{[k+l]}) is in TkA[l]\mathrm{T}^{k}A^{[l]}, considered as a subset of TkTl1A\mathrm{T}^{k}\mathrm{T}^{l-1}A. Let [a]k+l1A[k+l][a]_{k+l-1}\in A^{[k+l]} where aa is an admissible path in AA. Then ıAk,l([a]k+l1)=𝐭t=0k𝐭s=0l1a(t+s)\imath^{k,l}_{A}([a]_{k+l-1})=\bm{\mathrm{t}}^{k}_{t=0}\bm{\mathrm{t}}^{l-1}_{s=0}a(t+s). For any tt\in\mathbb{R}, the path sa(t+s)s\mapsto a(t+s) is admissible, hence the curve t𝐭s=0l1a(t+s)t\mapsto\bm{\mathrm{t}}^{l-1}_{s=0}a(t+s) lies in A[l1]A^{[l-1]}, so ıAk,l([a]k+l1)TkA[l1]\imath^{k,l}_{A}([a]_{k+l-1})\in\mathrm{T}^{k}A^{[l-1]} as we claimed. Using (2.10) we find that

(ıAk,l)(yi,(α,β))=yi,(α+β)(\imath^{k,l}_{A})^{\ast}(y^{i,({\alpha},{\beta})})=y^{i,({\alpha}+{\beta})} (2.37)

where αk\alpha\leq k, βl1{\beta}\leq l-1. Recall, (xa,yi,(α))(x^{a},y^{i,({\alpha})}), 0αk10\leq{\alpha}\leq k-1 is a coordinate chart for A[k]A^{[k]} induced from Tk1A\mathrm{T}^{k-1}A.

The rank of the graded bundle A[k]A^{[k]} is (r,r,r,,r)(r,r,r,\ldots,r) where r=rankAr=\operatorname{rank}A. Since ıAk1,1\imath^{k-1,1}_{A} is an inclusion and the ranks of the VBs A[k]widehat\widehat{A^{[k]}} and Tk1Awidehat\widehat{\mathrm{T}^{k-1}A} are the same, it induces an isomorphism ıAk1,1widehat:A[k]widehatTk1Awidehat\widehat{\imath^{k-1,1}_{A}}:\widehat{A^{[k]}}\rightarrow\widehat{\mathrm{T}^{k-1}A} of the core bundles. We define an isomorphism ȷA[k]:AA[k]widehatA[k]\jmath^{[k]}_{A}:A\xrightarrow{\simeq}\widehat{A^{[k]}}\subset A^{[k]} using the diagram

A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ȷA[k]\scriptstyle{\jmath^{[k]}_{A}}ȷAk\scriptstyle{\jmath^{k}_{A}}A[k]widehatA[k]\textstyle{\widehat{A^{[k]}}\subset A^{[k]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ıAk1,1\scriptstyle{\imath^{k-1,1}_{A}}Tk1AwidehatTk1A.\textstyle{\widehat{\mathrm{T}^{k-1}A}\subset\mathrm{T}^{k-1}A.} (2.38)

i.e., ıAk1,1ȷA[k]:ATk1A\imath^{k-1,1}_{A}\circ\jmath^{[k]}_{A}:A\rightarrow\mathrm{T}^{k-1}A coincides with ȷAk:ATk1AwidehatTk1A\jmath^{k}_{A}:A\xrightarrow{\simeq}\widehat{\mathrm{T}^{k-1}A}\subset\mathrm{T}^{k-1}A. In the special case A=TMA=\mathrm{T}M, the map ȷTM[k]\jmath^{[k]}_{\mathrm{T}M} coincides with jMk:TMTkMwidehatj^{k}_{M}:\mathrm{T}M\xrightarrow{\simeq}\widehat{\mathrm{T}^{k}M}, due to (2.12).

The following statement concerns the structure of the prolongation of an AL algebroid:

Lemma 2.16.

Let (A,κ)(A,\kappa) be an AL algebroid. The following diagram of isomorphisms is commutative

Γ(A)\textstyle{\operatorname{\Gamma}(A)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ȷA[k]\scriptstyle{\jmath^{[k]}_{A}}s1k!sk\scriptstyle{s\mapsto\frac{1}{k!}{s}^{\langle{-k}\rangle}}Γ(A[k]widehat)\textstyle{\operatorname{\Gamma}(\widehat{A^{[k]}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}vv\scriptstyle{v\mapsto{v^{\uparrow}}}𝔛k(A[k]).\textstyle{\mathfrak{X}_{-k}(A^{[k]}).}

In particular, for X𝔛(M)X\in\mathfrak{X}(M) we have a commutative diagram

𝔛(M)\textstyle{\mathfrak{X}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}jMk\scriptstyle{j^{k}_{M}}X1k!Xk\scriptstyle{X\mapsto{\frac{1}{k!}}{X}^{\langle{-k}\rangle}}Γ(TkMwidehat)\textstyle{\operatorname{\Gamma}(\widehat{\mathrm{T}^{k}M})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝔛k(TkM),\textstyle{\mathfrak{X}_{-k}(\mathrm{T}^{k}M),}

where Xk{X}^{\langle{-k}\rangle} is the algebroid (k)(-k)-lift of the vector field XX, associated with the HA (TkM,κMk)(\mathrm{T}^{k}M,\kappa^{k}_{M}). Moreover, the core of the anchor map [k]:A[k]TkM\sharp^{[k]}:A^{[k]}\rightarrow\mathrm{T}^{k}M can be identified with \sharp under the isomorphisms ȷA[k]:AA[k]𝑤𝑖𝑑𝑒ℎ𝑎𝑡\jmath^{[k]}_{A}:A\rightarrow\widehat{A^{[k]}} and jMk:TMTkM𝑤𝑖𝑑𝑒ℎ𝑎𝑡j^{k}_{M}:\mathrm{T}M\rightarrow\widehat{\mathrm{T}^{k}M}.

Proof.

In view of (2.32), it suffices to check that the first diagram is commutative for sections from the local frame (ei)(e_{i}) of Γ(A)\operatorname{\Gamma}(A). Due to the definition of ȷA[k]\jmath^{[k]}_{A}, this problem reduces to verifying that the vector fields 1k!sk𝔛k(A[k])\frac{1}{k!}{s}^{\langle{-k}\rangle}\in\mathfrak{X}_{-k}(A^{[k]}) and (ȷAk(s))𝔛k(Tk1A){\left(\jmath^{k}_{A}(s)\right)^{\uparrow}}\in\mathfrak{X}_{-k}(\mathrm{T}^{k-1}A) are ıAk1,1\imath^{k-1,1}_{A}-related, where we can take s=eiΓ(A)s=e_{i}\in\operatorname{\Gamma}(A). From (2.11), we see that (ȷAk(ei))=yi,(k1){\left(\jmath^{k}_{A}(e_{i})\right)^{\uparrow}}=\partial_{y^{i,(k-1)}}. The vector field sk{s}^{\langle{-k}\rangle} denotes the algebroid lift of ss with respect to (A[k],κ[k])(A^{[k]},\kappa^{[k]}), the kthk^{\mathrm{th}}-order prolongation of the algebroid (A,κ)(A,\kappa). From the definition of algebroid lifts and (2.23) we obtain

1k!eik=κ[k](ei(0))=κ[k](𝐕0kei)=yi,(k1),\frac{1}{k!}{e_{i}}^{\langle{-k}\rangle}={\overrightarrow{\kappa^{[k]}}}(e_{i}^{(0)})={\overrightarrow{\kappa^{[k]}}}({\overrightarrow{\mathbf{V}^{k}_{0}}}e_{i})=\partial_{y^{i,(k-1)}},

The last equality follows from the fact that κ\kappa is the identity on the core bundle; hence, the same holds for Tk1κ\mathrm{T}^{k-1}\kappa, as well as for κAk1:Tk1TATTk1A\kappa_{A}^{k-1}:\mathrm{T}^{k-1}\mathrm{T}A\to\mathrm{T}\mathrm{T}^{k-1}A and κ[k]\kappa^{[k]}.

For the last statement, concerning the case A=TMA=\mathrm{T}M, note that the inclusions A[k]Tk1AA^{[k]}\to\mathrm{T}^{k-1}A and TkMTk1TM\mathrm{T}^{k}M\rightarrow\mathrm{T}^{k-1}\mathrm{T}M induce the identity on the cores. Hence, [k]widehat\widehat{\sharp^{[k]}} coincides with Tk1widehat\widehat{\mathrm{T}^{k-1}\sharp}, which can be identified with :ATM\sharp:A\rightarrow\mathrm{T}M, as claimed. ∎

3 Structure of higher algebroids

In this section (Ek,κk)(E^{k},\kappa^{k}) is a HA of order kk and (A=E1,κ=κ1)(A=E^{1},\kappa=\kappa^{1}) is its reduction to order one.

3.1 Morphism Θk:A[k]Ek\mathrm{\Theta}^{k}:A^{[k]}\rightarrow E^{k}.

We shall construct a canonical VB morphism from kthk^{\mathrm{th}}-order prolongation A[k]A^{[k]} of an AL algebroid AA (see Preliminaries) to a given kthk^{\mathrm{th}}-order HA (Ek,κk)(E^{k},\kappa^{k}) whose order-one reduction coincides with AA.

Definition 3.1.

Let (Ek,κk)(E^{k},\kappa^{k}) be a HA of order kk. We apply the functor λ\lambda to the relation κk𝒢[k,1]\kappa^{k}\in\mathcal{GB}[k,1] and define the relation Θk\mathrm{\Theta}^{k} to be the intersection of A[k]×EkA^{[k]}\times E^{k} with λ(κk)\lambda(\kappa^{k}) subject to the natural inclusions and isomorphisms: ıAk1,1:A[k]Tk1Aλ(TkA)\imath^{k-1,1}_{A}:A^{[k]}\hookrightarrow\mathrm{T}^{k-1}A\simeq\lambda(\mathrm{T}^{k}A), and diagk:Eklin(Ek)=λ(TEk)\mathrm{diag}^{k}:E^{k}\hookrightarrow{\operatorname{lin}}(E^{k})=\lambda(\mathrm{T}E^{k}) (defined in Preliminaries):

Tk1Aλ(TkA)\textstyle{\mathrm{T}^{k-1}A\simeq\lambda(\mathrm{T}^{k}A)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}λ(κk)\scriptstyle{\lambda(\kappa^{k})}λ(TEk)=lin(Ek)\textstyle{\lambda(\mathrm{T}E^{k}){=}{\operatorname{lin}}(E^{k})}A[k]\textstyle{A^{[k]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Θk\scriptstyle{\mathrm{\Theta}^{k}}Ek\textstyle{E^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}
Theorem 3.2.

Let (σk:EkM,κk)(\sigma^{k}:E^{k}\rightarrow M,\kappa^{k}) be an AL HA and let (A,κ)(A,\kappa) be its order-one reduction. Then

  1. (a)

    Θk\mathrm{\Theta}^{k} is (the graph of) a 𝒢[k]\mathcal{GB}[k]-morphism, Θk:A[k]Ek\mathrm{\Theta}^{k}:A^{[k]}\rightarrow E^{k},

  2. (b)

    Θk\mathrm{\Theta}^{k} intertwines the anchor morphisms: kΘk=[k]\sharp^{k}\circ\mathrm{\Theta}^{k}=\sharp^{[k]}.

Proof.

First, we shall prove that if (U,V)Θk(U,V)\in\mathrm{\Theta}^{k} then [k](U)=k(V)\sharp^{[k]}(U)=\sharp^{k}(V). Then we shall show that Θk\mathrm{\Theta}^{k} is a mapping, and this will complete the proof of (a) and (b). Indeed, in vie of the characterisation of graded bundle morphisms [GR11], we only need to add that the relation Θk\mathrm{\Theta}^{k} is invariant with respect to the homogeneity structure on A[k]×EkA^{[k]}\times E^{k}. This is because the inclusions A[k]λ(TkA)A^{[k]}\hookrightarrow\lambda(\mathrm{T}^{k}A), Ekλ(TEk)E^{k}\hookrightarrow\lambda(\mathrm{T}E^{k}) are graded bundle morphisms.

Take (U,V)Θk(U,V)\in\mathrm{\Theta}^{k}, let us denote by U~\tilde{U} (resp., V~\tilde{V}) the images of UU (resp. VV) in Tk1A\mathrm{T}^{k-1}A (resp., lin(Ek){\operatorname{lin}}(E^{k})) and consider the diagram

U~Tk1A\textstyle{\tilde{U}\in\mathrm{T}^{k-1}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tk11\scriptstyle{\mathrm{T}^{k-1}\sharp^{1}}λ(κk)\scriptstyle{\lambda(\kappa^{k})}lin(Ek)V~\textstyle{{\operatorname{lin}}(E^{k})\ni\tilde{V}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}lin(k)\scriptstyle{{\operatorname{lin}}(\sharp^{k})}UA[k]\textstyle{U\in A^{[k]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[k]\scriptstyle{\sharp^{[k]}}Θk\scriptstyle{\mathrm{\Theta}^{k}}EkV\textstyle{E^{k}\ni V\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}k\scriptstyle{\sharp^{k}}Tk1TM\textstyle{\mathrm{T}^{k-1}\mathrm{T}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κMk1\scriptstyle{\kappa_{M}^{k-1}}TTk1M\textstyle{\mathrm{T}\mathrm{T}^{k-1}M}TkM\textstyle{\mathrm{T}^{k}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}iMk1,1\scriptstyle{i^{k-1,1}_{M}}=\scriptstyle{=}TkM\textstyle{\mathrm{T}^{k}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}iM1,k1\scriptstyle{i^{1,k-1}_{M}}
  • The top trapezoid in the middle commutes (i.e., (Tk11,lin(k)):λ(κk)κMk1(\mathrm{T}^{k-1}\sharp^{1},{\operatorname{lin}}(\sharp^{k})):\lambda(\kappa^{k})\Rightarrow\kappa_{M}^{k-1} is a morphism in the category 𝒱𝒞\mathcal{VBC}) because it is obtained by applying the functor λ\lambda to the diagram (2.20), which is commutative since (Ek,κk)(E^{k},\kappa^{k}) is almost Lie. Here we used Lemmas 2.2 and 2.5.

  • The parallelogram on the left also commutes as [k]=Tk11|A[k]\sharp^{[k]}=\mathrm{T}^{k-1}\sharp^{1}|_{A^{[k]}}, see [JR15, Theorem 4.5 (ix)].

  • The parallelogram on the right also commutes. This follows from a more general Lemma 2.3.

As (U,V)Θk(U,V)\in\mathrm{\Theta}^{k}, so (U~,V~)λ(κk)(\tilde{U},\tilde{V})\in\lambda(\kappa^{k}), hence Tk11(U~)Tk1TM\mathrm{T}^{k-1}\sharp^{1}(\tilde{U})\in\mathrm{T}^{k-1}\mathrm{T}M and lin(k)(V~)TTk1M{\operatorname{lin}}(\sharp^{k})(\tilde{V})\in\mathrm{T}\mathrm{T}^{k-1}M are related by means of κMk1\kappa_{M}^{k-1}. However, due to the commutativity of the left and right parallelograms, both Tk11(U~)\mathrm{T}^{k-1}\sharp^{1}(\tilde{U}) and lin(k)(V~){\operatorname{lin}}(\sharp^{k})(\tilde{V}) are images of [k](U)\sharp^{[k]}(U) and k(V)\sharp^{k}(V), respectively, under the canonical inclusions of TkM\mathrm{T}^{k}M into Tk1TM\mathrm{T}^{k-1}\mathrm{T}M and TTk1M\mathrm{T}\mathrm{T}^{k-1}M, respectively. Moreover, these images are κMk1\kappa_{M}^{k-1}-related, due to the commutativity of the top trapezoid. Since κMk1\kappa^{k-1}_{M} intertwines the canonical inclusions, κMk1iMk1,1=iM1,k1\kappa_{M}^{k-1}\circ i^{k-1,1}_{M}=i^{1,k-1}_{M}, we get [k](U)=k(V)\sharp^{[k]}(U)=\sharp^{k}(V), as was claimed.

Now we shall prove (a), i.e., that the relation Θk\mathrm{\Theta}^{k} is a mapping. We shall proceed by induction on kk.

Obviously, Θ1:AA\mathrm{\Theta}^{1}:A\rightarrow A is the identity mapping. Let k>1k>1 and assume that Θk1:A[k1]Ek1\mathrm{\Theta}^{k-1}:A^{[k-1]}\rightarrow E^{k-1} is a mapping. The graph of Θk1\mathrm{\Theta}^{k-1} is invariant with respect to the homogeneity structure of A[k1]×Ek1A^{[k-1]}\times E^{k-1}, hence Θk1\mathrm{\Theta}^{k-1} is a morphism of graded bundles.

Step A. We shall fist prove that for any UkA[k]U^{k}\in A^{[k]} there is at least one VkEkV^{k}\in E^{k} such that (Uk,Vk)Θk(U^{k},V^{k})\in\mathrm{\Theta}^{k}.

We know from Lemma 2.5 that λ(κk)\lambda(\kappa^{k}) is a VB comorphism covering k1:Ek1Tk1M\sharp^{k-1}:E^{k-1}\rightarrow\mathrm{T}^{k-1}M. Set V~k=λ(κk)v(U~k)lin(Ek)\tilde{V}^{k}=\lambda(\kappa^{k})_{v}(\tilde{U}^{k})\in{\operatorname{lin}}(E^{k}), where v:=Θk1(Uk1)Ek1v:=\mathrm{\Theta}^{k-1}(U^{k-1})\in E^{k-1} and Uk1=σk1k(Uk)A[k1]U^{k-1}=\sigma^{k}_{k-1}(U^{k})\in A^{[k-1]}. Consider the diagram

U~kTk1A\textstyle{\tilde{U}^{k}\in\mathrm{T}^{k-1}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tk1σ1\scriptstyle{\mathrm{T}^{k-1}\sigma^{1}}λ(κk)\scriptstyle{\lambda(\kappa^{k})}lin(Ek)V~k\textstyle{{\operatorname{lin}}(E^{k})\ni\tilde{V}^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}UkA[k]\textstyle{U^{k}\in A^{[k]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\circlearrowleft}Θk\scriptstyle{\mathrm{\Theta}^{k}}Ek\textstyle{E^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tk1M\textstyle{\mathrm{T}^{k-1}M}A[k1]\textstyle{A^{[k-1]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[k1]\scriptstyle{\sharp^{[k-1]}}Θk1\scriptstyle{\mathrm{\Theta}^{k-1}}Ek1\textstyle{E^{k-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}=\scriptstyle{=}Ek1v\textstyle{E^{k-1}\ni v\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}k1\scriptstyle{\sharp^{k-1}}

We shall check first that the definition of V~k\tilde{V}^{k} is correct, i.e.,

k1(v)=Tk1σ1(U~k).\sharp^{k-1}(v)=\mathrm{T}^{k-1}\sigma^{1}(\tilde{U}^{k}). (3.1)

This amounts to show that the compositions A[k]A[k1]Θk1Ek1k1Tk1MA^{[k]}\to A^{[k-1]}\xrightarrow{\mathrm{\Theta}^{k-1}}E^{k-1}\xrightarrow{\sharp^{k-1}}\mathrm{T}^{k-1}M and A[k]Tk1ATk1MA^{[k]}\xhookrightarrow{}\mathrm{T}^{k-1}A\to\mathrm{T}^{k-1}M coincide. According to our inductive hypothesis, k1Θk1\sharp^{k-1}\circ\mathrm{\Theta}^{k-1} is equal to [k1]\sharp^{[k-1]}, hence (3.1) reduces to the commutativity of the square diagram on the left (pointed by the circular arrow \circlearrowleft). The map [k1]\sharp^{[k-1]} is the restriction of Tk2\mathrm{T}^{k-2}\sharp to A[k1]Tk2AA^{[k-1]}\subset\mathrm{T}^{k-2}A, see [JR15, Theorem 4.5 (ix)], hence it suffices to prove that the following diagram is commutative.

A[k]\textstyle{A^{[k]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}A[k1]\textstyle{A^{[k-1]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tk2A\textstyle{\mathrm{T}^{k-2}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tk2\scriptstyle{\mathrm{T}^{k-2}\sharp}Tk1A\textstyle{\mathrm{T}^{k-1}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tk1σ\scriptstyle{\mathrm{T}^{k-1}\sigma}Tk1M\textstyle{\mathrm{T}^{k-1}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tk2TM\textstyle{\mathrm{T}^{k-2}\mathrm{T}M}

The inclusions A[k]Tk1AA^{[k]}\xhookrightarrow{}\mathrm{T}^{k-1}A are compatible with projections Tk+1ATkA\mathrm{T}^{k+1}A\rightarrow\mathrm{T}^{k}A, i.e., the following diagram on the left is commutative:

A[k]\textstyle{A^{[k]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}A[k1]\textstyle{A^{[k-1]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tk1A\textstyle{\mathrm{T}^{k-1}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tk2A,\textstyle{\mathrm{T}^{k-2}A,}         Tk1A\textstyle{\mathrm{T}^{k-1}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tk1σ\scriptstyle{\mathrm{T}^{k-1}\sigma}Tk2A\textstyle{\mathrm{T}^{k-2}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tk2\scriptstyle{\mathrm{T}^{k-2}\sharp}Tk1M\textstyle{\mathrm{T}^{k-1}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Tk2TM\textstyle{\mathrm{T}^{k-2}\mathrm{T}M}

The diagram on the right is not commutative in general, however for XA[k]Tk1AX\in A^{[k]}\subset\mathrm{T}^{k-1}A we have (by [JR15, Theorem 4.5 (viii)]): iMk2,1(Tk1σ)(X)=(Tk2)τk2k1(X)i^{k-2,1}_{M}\circ(\mathrm{T}^{k-1}\sigma)(X)=(\mathrm{T}^{k-2}\sharp)\circ\tau^{k-1}_{k-2}(X). This is enough for our claim (3.1).

Now we prove that V~klin(Ek)\tilde{V}^{k}\in{\operatorname{lin}}(E^{k}) is in Eklin(Ek)E^{k}\subset{\operatorname{lin}}(E^{k}). Consider the diagram

Ek\textstyle{E^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}diagk\scriptstyle{\mathrm{diag}^{k}}lin(Ek)V~k\textstyle{{\operatorname{lin}}(E^{k})\ni\tilde{V}^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π\scriptstyle{\pi}lin(σk1k)\scriptstyle{{\operatorname{lin}}(\sigma^{k}_{k-1})}vEk1\textstyle{v\in E^{k-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}diagk1\scriptstyle{\mathrm{diag}^{k-1}}lin(Ek1)\textstyle{{\operatorname{lin}}(E^{k-1})}

The subset of lin(Ek){\operatorname{lin}}(E^{k}) of those elements XX for which lin(σk1k)(X)=diagk1(π(X)){\operatorname{lin}}(\sigma^{k}_{k-1})(X)=\mathrm{diag}^{k-1}(\pi(X)) coincides with EkE^{k}. We are given v=Θk1(Uk1)Ek1v=\mathrm{\Theta}^{k-1}(U^{k-1})\in E^{k-1} and V~klin(Ek)\tilde{V}^{k}\in{\operatorname{lin}}(E^{k}) such that π(V~k)=v\pi(\tilde{V}^{k})=v and diagk1(v)=lin(σk1k)(V~k)\mathrm{diag}^{k-1}(v)={\operatorname{lin}}(\sigma^{k}_{k-1})(\tilde{V}^{k}). It follows that V~kEk\tilde{V}^{k}\in E^{k}.

Step B. We shall prove that for a given UkA[k]U^{k}\in A^{[k]} there is at most one VkEkV^{k}\in E^{k} such that (Uk,Vk)Θk(U^{k},V^{k})\in\mathrm{\Theta}^{k}. This will finish the proof of (a) and (b).

Assume (Uk,Vik)(U^{k},V^{k}_{i}) are in Θk\mathrm{\Theta}^{k} for i=1,2i=1,2. Using the inductive hypothesis, we know that V1k1=V2k1Ek1V^{k-1}_{1}=V^{k-1}_{2}\in E^{k-1}, hence VikV^{k}_{i}, considered as elements of lin(Ek){\operatorname{lin}}(E^{k}), are in the same fiber of the vector bundle π:lin(Ek)Ek1\pi:{\operatorname{lin}}(E^{k})\rightarrow E^{k-1}. As (Uk,Vik)λ(κk)(U^{k},V^{k}_{i})\in\lambda(\kappa^{k}) and λ(κk)\lambda(\kappa^{k}) is a VB comorphism over k1:Ek1Tk1M\sharp^{k-1}:E^{k-1}\rightarrow\mathrm{T}^{k-1}M, we must have V1k=λ(κk)v(Uk)=V2kV^{k}_{1}=\lambda(\kappa^{k})_{v}(U^{k})=V^{k}_{2} as we claimed. ∎

Example 3.3 (Θ2\mathrm{\Theta}^{2} and Θ3\mathrm{\Theta}^{3} in coordinates).

We shall provide an explicit coordinate expression for Θ2:A[2]E2\mathrm{\Theta}^{2}:A^{[2]}\rightarrow E^{2}, assuming κ2\kappa^{2} is given a general local form as in (2.21). According to the procedure given in the Definition 2.1, to get λ(κ2)\lambda(\kappa^{2}), in the first step we set yi=0y^{i}=0 and x¯a=0\underline{x}^{a}=0. Then we eliminate coordinates of weight (2,0)(2,0), i.e., the coordinates x¨a\ddot{x}^{a} and z¯μ\underline{z}^{\mu}. In this way we arrive at the VB comorphism λ(κ2):λ(T2A)lin(E2)\lambda(\kappa^{2}):\lambda(\mathrm{T}^{2}A){\rightarrow\!\!\vartriangleright}{\operatorname{lin}}(E^{2}) over :ATM\sharp:A\rightarrow\mathrm{T}M given by

λ(κ2):{x˙a=Qiay¯i,y¯˙i=y˙i,z¯˙μ=Qiμy¨i+Qijμy¯iy˙j.\lambda(\kappa^{2}):\begin{cases}\dot{x}^{a}=&Q^{a}_{i}\,\underline{y}^{i},\\ \dot{\underline{y}}^{i}=&\dot{y}^{i},\\ \dot{\underline{z}}^{\mu}=&Q^{\mu}_{i}\,\ddot{y}^{i}+Q^{\mu}_{ij}\,\underline{y}^{i}\dot{y}^{j}.\end{cases} (3.2)

Let (U2,V2)A[2]×E2(U^{2},V^{2})\in A^{[2]}\times E^{2}, and let (U~2,V~2)λ(T2A)×lin(E2)(\tilde{U}^{2},\tilde{V}^{2})\in\lambda(\mathrm{T}^{2}A)\times{\operatorname{lin}}(E^{2}) be the image of (U2,V2)(U^{2},V^{2}) under the canonical inclusions (see (2.37), (2.6), (2.13)):

IE2ıA1,1:A[2]TAλ(T2A),(xa,x˙a,y˙i,y¨i)(U~2)=(xa,Qiayi,yi,2y˙i)(U2)I^{2}_{E}\circ\imath^{1,1}_{A}:A^{[2]}\xhookrightarrow{}\mathrm{T}A\simeq\lambda(\mathrm{T}^{2}A),\quad(x^{a},\dot{x}^{a},\dot{y}^{i},\ddot{y}^{i})(\tilde{U}^{2})=(x^{a},Q^{a}_{i}y^{i},y^{i},2\dot{y}^{i})(U^{2})

and

diag2:E2lin(E2),(x¯a,y¯i,y¯˙i,z¯˙μ)(V~2)=(x¯a,y¯i,y¯i,2z¯μ)(V2).\mathrm{diag}^{2}:E^{2}\hookrightarrow{\operatorname{lin}}(E^{2}),\quad(\underline{x}^{a},\underline{y}^{i},\underline{\dot{y}}^{i},\underline{\dot{z}}^{\mu})(\tilde{V}^{2})=(\underline{x}^{a},\underline{y}^{i},\underline{y}^{i},2\underline{z}^{\mu})(V^{2}).

(Recall that (xa,(α),yi,(β))(x^{a,({\alpha})},y^{i,({\beta})}), 0αk10\leq{\alpha}\leq k-1, 1βk1\leq{\beta}\leq k, are coordinates for λ(TkA)\lambda(\mathrm{T}^{k}A) inherited from the adapted coordinates on TkA\mathrm{T}^{k}A. The coordinate system for lin(Ek)=λ(TEk){\operatorname{lin}}(E^{k})=\lambda(\mathrm{T}E^{k}) is inherited from the adapted coordinate system on the tangent bundle of EkE^{k}.) By plugging these expressions to (3.2) we find that (U~2,V~2)λ(κ2)(\tilde{U}^{2},\tilde{V}^{2})\in\lambda(\kappa^{2}) if and only if

{Qiayi=Qiay¯i,y¯i=yi,2z¯μ=Qiμ 2y˙i+Qijμy¯iyj,\begin{cases}Q^{a}_{i}y^{i}&=Q^{a}_{i}\,\underline{y}^{i},\\ \underline{y}^{i}&=y^{i},\\ 2\underline{z}^{\mu}&=Q^{\mu}_{i}\,2\dot{y}^{i}+Q^{\mu}_{ij}\,\underline{y}^{i}y^{j},\end{cases}

hence Θ2\mathrm{\Theta}^{2} is an affine bundle morphism Θ2:A[2]E2\mathrm{\Theta}^{2}:A^{[2]}\rightarrow E^{2} covering the identity idA\operatorname{id}_{A} given by

Θ2(xa,yi,y˙i)=(x¯a=xa,y¯i=yi,z¯μ=Qiμy˙i+12Q(ij)μyiyj),\mathrm{\Theta}^{2}(x^{a},y^{i},\dot{y}^{i})=(\underline{x}^{a}=x^{a},\underline{y}^{i}=y^{i},\underline{z}^{\mu}=Q^{\mu}_{i}\dot{y}^{i}+\frac{1}{2}Q^{\mu}_{(ij)}y^{i}y^{j}), (3.3)

where Qijμ=Q(ij)μ+Q[ij]μQ^{\mu}_{ij}=Q^{\mu}_{(ij)}+Q^{\mu}_{[ij]} is the decomposition into symmetric and anty-symmetric part, namely

Q(ij)μ=12Qijμ+12Qjiμ,Q[ij]μ=12Qijμ12Qjiμ.Q^{\mu}_{(ij)}=\frac{1}{2}Q^{\mu}_{ij}+\frac{1}{2}Q^{\mu}_{ji},\quad Q^{\mu}_{[ij]}=\frac{1}{2}Q^{\mu}_{ij}-\frac{1}{2}Q^{\mu}_{ji}. (3.4)

In order three, additional equations for κ3\kappa^{3} appear. Let (xa,yi,zμ,tα)(x^{a},y^{i},z^{\mu},t^{\alpha}) be graded coordinates on a graded bundle σ3:E3M\sigma^{3}:E^{3}\rightarrow M where the coordinates tαt^{\alpha} have order 3. The additional equations for κ3\kappa^{3}, extending those for κ2\kappa^{2}, are of the form (we have omitted expressions that do not account for λ(κ3)\lambda(\kappa^{3})): x˙˙˙a=x˙˙˙a(x¯,y¯,z¯,t¯)\dddot{x}^{a}=\dddot{x}^{a}(\underline{x},\underline{y},\underline{z},\underline{t}) and

t¯˙α=Qiαy˙˙˙i+Qijαy¯iy¨j+Qνiαz¯νy˙i+12Qij,kαy¯iy¯jy˙k+qiα(x¯,y¯,z¯,t¯)yi\dot{\underline{t}}^{\alpha}=Q^{\alpha}_{i}\dddot{y}^{i}+Q^{\alpha}_{ij}\underline{y}^{i}\ddot{y}^{j}+Q^{\alpha}_{\nu i}\underline{z}^{\nu}\dot{y}^{i}+\frac{1}{2}Q^{\alpha}_{ij,k}\underline{y}^{i}\underline{y}^{j}\dot{y}^{k}+{q^{\alpha}_{i}(\underline{x},\underline{y},\underline{z},\underline{t})y^{i}}

for some functions qiαq^{\alpha}_{i} on E3E^{3} of weight 33. Now we set to zero the coordinates of weight (0,1)(0,1), i.e., yi=0y^{i}=0, x¯˙a=0\underline{\dot{x}}^{a}=0 and eliminate the coordinates of weight (3,0)(3,0), x˙˙˙a\dddot{x}^{a} and t¯α\underline{t}^{\alpha}. This way we get, in addition to (3.2), the following equations defining λ(κ3)\lambda(\kappa^{3}):

λ(κ3):{x¨a=Qμaz¯μ+12Qijay¯iy¯j,t¯˙α=Qiαy˙˙˙i+Qijαy¯iy¨j+Qνiαz¯νy˙i+12Qij,kαy¯iy¯jy˙k.\lambda(\kappa^{3}):\begin{cases}\ddot{x}^{a}&=Q^{a}_{\mu}\underline{z}^{\mu}+\frac{1}{2}Q^{a}_{ij}\underline{y}^{i}\underline{y}^{j},\\ \dot{\underline{t}}^{\alpha}&=Q^{\alpha}_{i}\dddot{y}^{i}+Q^{\alpha}_{ij}\underline{y}^{i}\ddot{y}^{j}+Q^{\alpha}_{\nu i}\underline{z}^{\nu}\dot{y}^{i}+\frac{1}{2}Q^{\alpha}_{ij,k}\underline{y}^{i}\underline{y}^{j}\dot{y}^{k}.\end{cases} (3.5)

which is a VB comorphism over 2\sharp^{2}:

A[3]T2Aλ(T3A)\textstyle{A^{[3]}\subset\mathrm{T}^{2}A\simeq\lambda(\mathrm{T}^{3}A)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}λ(κ3)\scriptstyle{\lambda(\kappa^{3})}lin(E3)E3\textstyle{{\operatorname{lin}}(E^{3})\supset E^{3}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}T2M\textstyle{\mathrm{T}^{2}M}E2\textstyle{E^{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}2\scriptstyle{\sharp^{2}}

Let U3A[3]U^{3}\in A^{[3]}, V3E3V^{3}\in E^{3} and let U~3\tilde{U}^{3} in λ(T3A)\lambda(\mathrm{T}^{3}A) and V~3\tilde{V}^{3} in lin(E3){\operatorname{lin}}(E^{3}) denote their images subject to the inclusions A[3]λ(T2A)A^{[3]}\xhookrightarrow{}\lambda(\mathrm{T}^{2}A) and diag:E3lin(E3)\mathrm{diag}:E^{3}\xhookrightarrow{}{\operatorname{lin}}(E^{3}). We have

(xa,x˙a,x¨a;y˙i,y¨i,y˙˙˙i)(U~3)=(xa,Qiayi,Xa;yi,2y˙i,3y¨i)(U3)(xa,y¯i,z¯μ;y¯˙i,z¯˙μ,t¯˙α)(V~3)=(xa,y¯i,z¯μ,y¯i,2z¯μ,3t¯α)(V3).\begin{split}(x^{a},\dot{x}^{a},\ddot{x}^{a};\dot{y}^{i},\ddot{y}^{i},\dddot{y}^{i})(\tilde{U}^{3})=(x^{a},Q^{a}_{i}y^{i},X^{a};y^{i},2\dot{y}^{i},3\ddot{y}^{i})(U^{3})\\ (x^{a},\underline{y}^{i},\underline{z}^{\mu};\underline{\dot{y}}^{i},\underline{\dot{z}}^{\mu},\underline{\dot{t}}^{\alpha})(\tilde{V}^{3})=(x^{a},\underline{y}^{i},\underline{z}^{\mu},\underline{y}^{i},2\underline{z}^{\mu},3\underline{t}^{\alpha})(V^{3}).\end{split} (3.6)

where, due to the definition of A[k]A^{[k]},

Xa=x¨a(U~3)=(Qiayi)=Qiaxbx˙byi+Qiay˙i=12Qwidehatijayiyj+Qiay˙i,X^{a}=\ddot{x}^{a}(\tilde{U}^{3})=(Q^{a}_{i}y^{i})^{\cdot}=\frac{\partial Q^{a}_{i}}{\partial x^{b}}\dot{x}^{b}y^{i}+Q^{a}_{i}\dot{y}^{i}=\frac{1}{2}\widehat{Q}^{a}_{ij}y^{i}y^{j}+Q^{a}_{i}\dot{y}^{i},

where

Qwidehatija=QiaxbQjb+QjaxbQib.\widehat{Q}^{a}_{ij}=\frac{\partial Q^{a}_{i}}{\partial x^{b}}Q^{b}_{j}+\frac{\partial Q^{a}_{j}}{\partial x^{b}}Q^{b}_{i}. (3.7)

Now assume that U~3\tilde{U}^{3}, V~3\tilde{V}^{3} are λ(κ3)\lambda(\kappa^{3})-related. We shall show that the first equation for λ(κ3)\lambda(\kappa^{3}) in (3.5) is satisfied automatically. It amounts to show that x¨a(U~3)\ddot{x}^{a}(\tilde{U}^{3}) given above coincides with

Qμaz¯μ+12Qijay¯iy¯j(on V~3)=Qμa(Qiμy˙i+12Q(ij)μyiyj)+12Qijayiyj(on U~3).Q^{a}_{\mu}\underline{z}^{\mu}+\frac{1}{2}Q^{a}_{ij}\underline{y}^{i}\underline{y}^{j}(\text{on }\tilde{V}^{3})=Q^{a}_{\mu}\left(Q^{\mu}_{i}\dot{y}^{i}+\frac{1}{2}Q^{\mu}_{(ij)}y^{i}y^{j}\right)+\frac{1}{2}Q^{a}_{ij}y^{i}y^{j}(\text{on }\tilde{U}^{3}).

The last equality is due to (3.3) as the reductions of U~3\tilde{U}^{3}, V~3\tilde{V}^{3} to order 2 are λ(κ2)\lambda(\kappa^{2})-related. By comparing the coefficients at y˙i\dot{y}^{i} and at yiyjy^{i}y^{j} we see that the first equation in (3.5) is equivalent to the equations (4.13a) and (4.16b) considered in Appendix, which are true in any order-two AL HA. The second equation for λ(κ3)\lambda(\kappa^{3}) gives

t¯α=t¯α(V3)=13t¯˙α(V~3)=13Qiαy˙˙˙i(U~3)+13Qijαy¯i(V~3)y¨j(U~3)+13Qνiαz¯ν(V~3)y˙i(U~3)+16Qij,kαy¯iy¯j(V~3)y˙k(U~3)\underline{t}^{\alpha}=\underline{t}^{\alpha}(V^{3})=\frac{1}{3}\underline{\dot{t}}^{\alpha}(\tilde{V}^{3})=\frac{1}{3}Q^{\alpha}_{i}\dddot{y}^{i}(\tilde{U}^{3})+\frac{1}{3}Q^{\alpha}_{ij}\underline{y}^{i}(\tilde{V}^{3})\ddot{y}^{j}(\tilde{U}^{3})+\frac{1}{3}Q^{\alpha}_{\nu i}\underline{z}^{\nu}(\tilde{V}^{3})\dot{y}^{i}(\tilde{U}^{3})+\frac{1}{6}Q^{\alpha}_{ij,k}\underline{y}^{i}\underline{y}^{j}(\tilde{V}^{3})\dot{y}^{k}(\tilde{U}^{3})

from which, using (3.6), we find a complete formula for Θ3\mathrm{\Theta}^{3},

Θ3:(xa,yi,y˙i,y¨i)(x¯a=xa,y¯i=yi,z¯μ=12Q(ij)μyiyj,t¯α=Qiαy¨i+(23Qijα+13QνiαQjν)yiy˙j+16(QνiαQjkν+Qij,kα)yiyjyk).\begin{split}\mathrm{\Theta}^{3}:(x^{a},y^{i},\dot{y}^{i},\ddot{y}^{i})\mapsto(\underline{x}^{a}=x^{a},\underline{y}^{i}=y^{i},\underline{z}^{\mu}=\frac{1}{2}Q^{\mu}_{(ij)}y^{i}y^{j},\\ \underline{t}^{\alpha}=Q^{\alpha}_{i}\ddot{y}^{i}+(\frac{2}{3}Q^{\alpha}_{ij}+\frac{1}{3}Q^{\alpha}_{\nu i}Q^{\nu}_{j})y^{i}\dot{y}^{j}+\frac{1}{6}(Q^{\alpha}_{\nu i}Q^{\nu}_{jk}+Q^{\alpha}_{ij,k})y^{i}y^{j}y^{k}).\end{split} (3.8)
Remark 3.4.

In deriving the formula for the mapping Θ2\mathrm{\Theta}^{2} we did not use the assumption from Theorem 3.2 that (E2,κ2)(E^{2},\kappa^{2}) is AL. Actually, Θ2\mathrm{\Theta}^{2} is a well defined mapping for any skew HA (E2,κ2)(E^{2},\kappa^{2}).

Conjecture 3.5.

Let (σk:EkM,κk)(\sigma^{k}:E^{k}\rightarrow M,\kappa^{k}) be a Lie HA. Then Θk:A[k]Ek\mathrm{\Theta}^{k}:A^{[k]}\rightarrow E^{k} is a HA morphism.

It suffices to prove that (TkΘ1,TΘk):κ[k]κk(\mathrm{T}^{k}\mathrm{\Theta}^{1},\mathrm{T}\mathrm{\Theta}^{k}):\kappa^{[k]}\Rightarrow\kappa^{k} is a 𝒱𝒞\mathcal{VBC}-morphism. Recall, Θ1\mathrm{\Theta}^{1} is the identity on E1E^{1}, hence it remains to verify that the diagram

TkA\textstyle{\mathrm{T}^{k}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}κk\scriptstyle{\kappa^{k}}κ[k]\scriptstyle{\kappa^{[k]}}TEk\textstyle{\mathrm{T}E^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}TA[k]\textstyle{\mathrm{T}A^{[k]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}TΘk\scriptstyle{\mathrm{T}\mathrm{\Theta}^{k}}TkM\textstyle{\mathrm{T}^{k}M}Ek\textstyle{E^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}k\scriptstyle{\sharp^{k}}A[k]\textstyle{A^{[k]}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}[k]\scriptstyle{\sharp^{[k]}}Θk\scriptstyle{\mathrm{\Theta}^{k}} (3.9)

is commutative. We already know that the bottom triangle is commutative (see Theorem 3.2). Therefore, we now need to prove that

(κk)Z¯(X)=TΘk((κk)Y¯(X))\left(\kappa^{k}\right)_{\underline{Z}}(X)=\mathrm{T}\mathrm{\Theta}^{k}(\left(\kappa^{k}\right)_{\underline{Y}}(X)) (3.10)

for any Y¯A[k]\underline{Y}\in A^{[k]} and XTkAX\in\mathrm{T}^{k}A such that Tkσ1(X)=[k](Y¯)\mathrm{T}^{k}\sigma^{1}(X)=\sharp^{[k]}(\underline{Y}), where Z¯=Θk(Y¯)Ek\underline{Z}=\mathrm{\Theta}^{k}(\underline{Y})\in E^{k}. In other words, this means that the algebroid α\langle{\alpha}\rangle-lifts with respect to (Ek,κk)(E^{k},\kappa^{k}) and (A[k],κ[k])(A^{[k]},\kappa^{[k]}) are Θk\mathrm{\Theta}^{k} related for all kα0-k\leq\alpha\leq 0. We shall prove Conjecture 3.5 for k=2k=2 by direct computations. See Appendix, Subsection 4.3.

3.2 Higher algebroids in order two

In this subsection, we shall look closer at higher algebroids (E2,κ2)(E^{2},\kappa^{2}) of order two. First, we shall describe the structure of the graded bundle E2E^{2}, see Lemma 3.6. Then, we shall derive a number of structure maps which fully determine (E2,κ2)(E^{2},\kappa^{2}) and reformulate the definition of a skew HA in terms of these structure maps and relations between them, see Theorem 3.13. We shall examine skew and Lie HAs in which the base MM is a point, see Theorem 3.15. We shall also find the relations between the structure maps and the conditions under which (E2,κ2)(E^{2},\kappa^{2}) becomes an almost Lie (Theorem 3.16) and a Lie HA (Theorem 3.20). Finally, we will describe the relation between order-two HAs and Lie algebroid representations up to homotopy, see Theorem 3.26. We assume that (E2,κ2)(E^{2},\kappa^{2}) is a skew HA.

Throughout this subsection, we denote by AA the order-one reduction of E2E^{2}, i.e., A=E1A=E^{1}, and C=E2widehatC=\widehat{E^{2}} – the core of E2E^{2}. We set (xa,yi,zμ)(x^{a},y^{i},z^{\mu}) as a system of graded coordinates on E2E^{2} and fix local frames (ei)(e_{i}) of Γ(A)\operatorname{\Gamma}(A) and (cμ)(c_{\mu}) of Γ(C)\operatorname{\Gamma}(C) such that

yi(ej)=δji,zμ|C(cν)=δνμ.y^{i}(e_{j})=\delta^{i}_{j},\quad z^{\mu}|_{C}(c_{\nu})=\delta^{\mu}_{\nu}. (3.11)

3.2.1 The structure of the graded bundle of a skew HA of order two

Define

:=Θ2widehat:AC\partial:=\widehat{\mathrm{\Theta}^{2}}:A\rightarrow C (3.12)

as the core of the map Θ2:A[2]E2\mathrm{\Theta}^{2}:A^{[2]}\rightarrow E^{2}, see Definition 3.1, where the core of A[2]A^{[2]} is identified with AA under the isomorphism ȷA[2]:AA[2]widehat\jmath^{[2]}_{A}:A\simeq\widehat{A^{[2]}}, see (2.38). . Consider the map Θ~2\tilde{\mathrm{\Theta}}^{2} defined on the product (over MM) of graded bundles A[2]A^{[2]} and C[2]C_{[2]} by

Θ~2:A[2]×MC[2]E2,(a,c)Θ2(a)+c\tilde{\mathrm{\Theta}}^{2}:A^{[2]}\times_{M}C_{[2]}\to E^{2},\quad(a,c)\mapsto\mathrm{\Theta}^{2}(a)\bm{\boldsymbol{+}}c (3.13)

(We recall that F[k]F_{[k]} stands for the graded bundle defined on the total space of the VB FMF\rightarrow M by assigning weight kk to linear functions on FF.) The map Θ~2\tilde{\mathrm{\Theta}}^{2} is a surjective morphism of graded bundles, hence E2E^{2} can be identified as the quotient of A[2]×MC[2]A^{[2]}\times_{M}C_{[2]} by the equivalence relations \sim, where

(a,c)(a,c)Θ2(a)+c=Θ2(a)+c.(a,c)\sim(a^{\prime},c^{\prime})\iff\mathrm{\Theta}^{2}(a)\bm{\boldsymbol{+}}c=\mathrm{\Theta}^{2}(a^{\prime})\bm{\boldsymbol{+}}c^{\prime}.

Since Θ2\mathrm{\Theta}^{2} covers the identity idA\operatorname{id}_{A}, the elements aa and aA[2]a^{\prime}\in A^{[2]} project to the same element in AA. As E2E1E^{2}\rightarrow E^{1} is an affine bundle, we can write Θ2(a)Θ2(a)=Θ2widehat(aa)=ccC\mathrm{\Theta}^{2}(a^{\prime})\bm{\boldsymbol{-}}\mathrm{\Theta}^{2}(a)=\widehat{\mathrm{\Theta}^{2}}(a^{\prime}-a)=c^{\prime}-c\in C. In other words, (a,c)(a,c)=(aa,cc)(a^{\prime},c^{\prime})\bm{\boldsymbol{-}}(a,c)=(a^{\prime}-a,c^{\prime}-c) is in the graph of the map -\partial, which is a subset of A×MCA\times_{M}C, the core of the graded bundle  A[2]×MC[2]A^{[2]}\times_{M}C_{[2]}. Therefore, what we need to define E2E^{2} is only the map \partial.

Lemma 3.6.

Let (E2,κ2)(E^{2},\kappa^{2}) be a skew, order-two HA.

  1. (i)

    There is a canonical isomorphism of graded bundles

    E2(A[2]×MC[2])/graph()E^{2}\simeq{\raisebox{1.79997pt}{$(A^{[2]}\times_{M}C_{[2]})$}\left/\raisebox{-1.79997pt}{$\operatorname{graph}(-\partial)$}\right.}

    where :AC\partial:A\rightarrow C is given in (3.12).

  2. (ii)

    A choice of local frames (ei)(e_{i}) and (cμ)(c_{\mu}) of Γ(A)\operatorname{\Gamma}(A) and Γ(C)\operatorname{\Gamma}(C), respectively, gives rise to a graded coordinate system (xa,yi,wμ)(x^{a},y^{i},w^{\mu}) for E2E^{2} (considered as the quotient of A[2]×MC[2]A^{[2]}\times_{M}C_{[2]}), defined by wμ=cμ+Qiμy˙iw^{\mu}=c_{\mu}^{\ast}+Q^{\mu}_{i}\dot{y}^{i}. The composition A[2]A[2]×MC[2](A[2]×MC[2])/graph()A^{[2]}\hookrightarrow A^{[2]}\times_{M}C_{[2]}\rightarrow{\raisebox{1.79997pt}{$(A^{[2]}\times_{M}C_{[2]})$}\left/\raisebox{-1.79997pt}{$\operatorname{graph}(-\partial)$}\right.} coincides with the map Θ2\mathrm{\Theta}^{2} which, in the introduced coordinates (xa,yi,wμ)(x^{a},y^{i},w^{\mu}), read as

    (Θ2)(xa)=xa,(Θ2)(yi)=yi,(Θ2)(wμ)=Qiμy˙i.(\mathrm{\Theta}^{2})^{\ast}(x^{a})=x^{a},\quad(\mathrm{\Theta}^{2})^{\ast}(y^{i})=y^{i},\quad(\mathrm{\Theta}^{2})^{\ast}(w^{\mu})=Q^{\mu}_{i}\dot{y}^{i}.
Definition 3.7.

We call (xa,yi,wμ)(x^{a},y^{i},w^{\mu}) an adapted system of graded coordinated on a HA (E2,κ2)(E^{2},\kappa^{2}). It is uniquely defined once we set a system of local frames (ei)(e_{i}), (cμ)(c_{\mu}), and is characterised by the equality Q(ij)μ=0Q^{\mu}_{(ij)}=0.

Proof.

Set E~2=A[2]×MC[2]\tilde{E}^{2}=A^{[2]}\times_{M}C_{[2]}, V=graph()V=\operatorname{graph}(-\partial). We have already shown that there is a well defined bijection between E2E^{2} and the quotient E~2/{\raisebox{1.79997pt}{$\tilde{E}^{2}$}\left/\raisebox{-1.79997pt}{$\sim$}\right.} defined as the set of equivalence classes of the following equivalence relation: eee\sim e^{\prime} if and only if there exists vVv\in V such that e=e+ve^{\prime}=e\bm{\boldsymbol{+}}v. It is also evident that this bijection is an isomorphism of graded bundles since it is a special case of the following more general construction: given a graded bundle EkE^{k} of order kk and a vector subbundle VMV\rightarrow M of the core Ekwidehat\widehat{E^{k}}, the quotient Ek/V{\raisebox{1.79997pt}{$E^{k}$}\left/\raisebox{-1.79997pt}{$V$}\right.} which is the orbit space of the action on EkE^{k} of the subbundle VEkwidehatV\subset\widehat{E^{k}} of the core, inherits a graded bundle structure from EkE^{k}.

Recall that (yi)(y_{i}) and (cμ)(c_{\mu}^{\ast}) denote the dual frames to (ei)(e_{i}) and (cμ)(c_{\mu}), respectively. Then (xa,yi,y˙i,cμ)(x^{a},y^{i},\dot{y}^{i},c_{\mu}^{\ast}) forms a graded coordinate system on A[2]×MC[2]A^{[2]}\times_{M}C_{[2]}. The introduced equivalence relation on this space reads as: (xa,yi,y˙i,cμ)(x¯a,y¯i,y¯˙i,c¯μ)(x^{a},y^{i},\dot{y}^{i},c_{\mu}^{\ast})\sim(\underline{x}^{a},\underline{y}^{i},\dot{\underline{y}}^{i},\underline{c}_{\mu}^{\ast}) if and only if xa=x¯ax^{a}=\underline{x}^{a}, yi=y¯iy^{i}=\underline{y}^{i}, c¯μcμ=Qiμ(y˙iy¯˙i)\underline{c}_{\mu}^{\ast}-c_{\mu}^{\ast}=Q^{\mu}_{i}(\dot{y}^{i}-\dot{\underline{y}}^{i}). Therefore, the functions wμ:=cμ+Qiμy˙iw^{\mu}:=c_{\mu}^{\ast}+Q^{\mu}_{i}\dot{y}^{i} are well-defined on the quotient E~2/{\raisebox{1.79997pt}{$\tilde{E}^{2}$}\left/\raisebox{-1.79997pt}{${\sim}$}\right.}, and (xa,yi,wμ)(x^{a},y^{i},w^{\mu}) is a graded coordinate system on this quotient.

Let (xa,yi,zμ)(x^{a},y^{i},z^{\mu}) be as in (3.11). The map Θ~2:E~2E2\tilde{\mathrm{\Theta}}^{2}:\tilde{E}^{2}\rightarrow E^{2}, defined in (3.13), is given by

(Θ~2)(xa)=xa,(Θ~2)(yi)=yi,(Θ~2)(zμ)=cμ+Qiμy˙i+12Q(ij)μyiyj=wμ+12Q(ij)μyiyj.\left(\tilde{\mathrm{\Theta}}^{2}\right)^{\ast}(x^{a})=x^{a},\left(\tilde{\mathrm{\Theta}}^{2}\right)^{\ast}(y^{i})=y^{i},\left(\tilde{\mathrm{\Theta}}^{2}\right)^{\ast}(z^{\mu})=c_{\mu}^{\ast}+Q^{\mu}_{i}\dot{y}^{i}+\frac{1}{2}Q^{\mu}_{(ij)}y^{i}y^{j}=w^{\mu}+\frac{1}{2}Q^{\mu}_{(ij)}y^{i}y^{j}.

Hence, the isomorphism from point (i), denoted by I:E~2/E2I:{\raisebox{1.79997pt}{$\tilde{E}^{2}$}\left/\raisebox{-1.79997pt}{${\sim}$}\right.}\to E^{2}, is given by I(zμ)=wμ+12Q(ij)μyiyjI^{\ast}(z^{\mu})=w^{\mu}+\frac{1}{2}Q^{\mu}_{(ij)}y^{i}y^{j}, and the composition of the inclusion A[2]A[2]×MC[2]A^{[2]}\hookrightarrow A^{[2]}\times_{M}C_{[2]} with Θ~2\tilde{\mathrm{\Theta}}^{2} coincides with (3.3), i.e., with the formula for Θ2\mathrm{\Theta}^{2}, which proves the claim from point (ii) and completes the proof. ∎

3.2.2 The structure maps of a skew HA of order two

The subspace 𝔛2(E2)𝔛(E2)\mathfrak{X}_{-2}(E^{2})\subset\mathfrak{X}(E^{2}) of vector fields of weight 2-2 is a locally free 𝒞(M)\mathcal{C}^{\infty}(M)-module canonically isomorphic to the space of sections of the core bundle C=E2widehatC=\widehat{E^{2}}. We shall often identify these spaces without further comment. In coordinates as in (3.11), the isomorphism takes cμc_{\mu} to cμ=zμ{c_{\mu}^{\uparrow}}=\partial_{z^{\mu}}, see Lemma 2.6.

The reduction of κ2\kappa^{2} to order one yields a skew algebroid, whose structure maps will be denoted by [,][\cdot,\cdot] and :=1:ATM\sharp:=\sharp^{1}:A\rightarrow\mathrm{T}M. We assume that κ2\kappa^{2} has a local form introduced in (2.21). Then

ei=Qiaxa,[ei,ej]=Qijkek,\sharp e_{i}=Q^{a}_{i}\partial_{x^{a}},\quad[e_{i},e_{j}]=Q^{k}_{ij}e_{k},

where (ei)(e_{i}) is a local frame of sections of AMA\to M which is dual to the frame (yi)(y^{i}). The core of the anchor map 2:E2T2M\sharp^{2}:E^{2}\rightarrow\mathrm{T}^{2}M provides a VB morphism

C:CTM,C(cμ)=Qμaxa.\sharp^{C}:C\rightarrow\mathrm{T}M,\quad\sharp^{C}(c_{\mu})=Q^{a}_{\mu}\partial_{x^{a}}. (3.14)

In more detail, C\sharp^{C} is the composition of 2widehat:E2widehatT2Mwidehat\widehat{\sharp^{2}}:\widehat{E^{2}}\to\widehat{\mathrm{T}^{2}M} with the isomorphism T2MwidehatTM\widehat{T^{2}M}\simeq\mathrm{T}M, see (2.9).

The next mapping is a VB morphism ~:AC\tilde{\partial}:A\rightarrow C defined by 𝒞(M)\mathcal{C}^{\infty}(M)-linear map

s12s2𝔛2(E2)Γ(C),s\mapsto{\frac{1}{2}{s}^{\langle{-2}\rangle}}\in\mathfrak{X}_{-2}(E^{2})\simeq\operatorname{\Gamma}(C), (3.15)

where sΓ(A)s\in\operatorname{\Gamma}(A), see (2.29) for algebroid lifts. (It will turn out soon that ~=\tilde{\partial}=\partial, the core of the map Θ2\mathrm{\Theta}^{2}.) In a similar manner we define

β:Γ(A)×Γ(A)Γ(C),β(s1,s2)=12[s11,s21]𝔛2(E2)Γ(C),\beta:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(A)\rightarrow\operatorname{\Gamma}(C),\quad\beta(s_{1},s_{2})={\frac{1}{2}\,[{s_{1}}^{\langle{-1}\rangle},{s_{2}}^{\langle{-1}\rangle}]}\in\mathfrak{X}_{-2}(E^{2})\simeq\operatorname{\Gamma}(C), (3.16)

which is a skew-symmetric mapping and

:Γ(A)×Γ(C)Γ(C),sv=[s0,v]𝔛2(E2)Γ(C)\Box:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(C)\rightarrow\operatorname{\Gamma}(C),\quad\Box_{s}v=[{s}^{\langle{0}\rangle},v]\in\mathfrak{X}_{-2}(E^{2})\simeq\Gamma(C) (3.17)

called the action of AA on CC. The system of equations (2.21) describing κ2\kappa^{2} results in the following formulas for the algebroid lifts ekα{e_{k}}^{\langle{{\alpha}}\rangle}:

{ek0=Qkaxa+Qjkiyjyi+(Qνkμzν+12Qij,kμyiyj)zμ,ek1=yk+Qikμyizμ,ek2=2Qkμzμ.\begin{cases}{e_{k}}^{\langle{0}\rangle}&=Q^{a}_{k}\partial_{x^{a}}+Q_{jk}^{i}y^{j}\partial_{y^{i}}+(Q^{\mu}_{\nu k}\,z^{\nu}+\frac{1}{2}\,Q_{ij,k}^{\mu}\,y^{i}y^{j})\,\partial_{z^{\mu}},\\ {e_{k}}^{\langle{-1}\rangle}&=\partial_{y^{k}}+Q^{\mu}_{ik}\,y^{i}\,\partial_{z^{\mu}},\\ {e_{k}}^{\langle{-2}\rangle}&={2}Q^{\mu}_{k}\,\partial_{z^{\mu}}.\end{cases} (3.18)

From this, we can easily derive the coordinate expressions for the introduced mappings ~,β,\tilde{\partial},\beta,\Box:

{~(ei)=Qiμcμ,β(ei,ej)=Q[ij]μcμeicν=Qνiμcμ.\begin{cases}\tilde{\partial}(e_{i})&=Q^{\mu}_{i}c_{\mu},\\ \beta(e_{i},e_{j})&=Q^{\mu}_{[ij]}c_{\mu}\\ \Box_{e_{i}}c_{\nu}&=-Q^{\mu}_{\nu i}c_{\mu}.\end{cases} (3.19)

where Q[ij]μQ^{\mu}_{[ij]} is given in (3.4), and the minus sign in the last line arises from our preference for working with left actions. Note that ~\tilde{\partial} coincides with \partial as Θ2widehat(ei)=Qiμcμ\widehat{\mathrm{\Theta}^{2}}(e_{i})=Q^{\mu}_{i}c_{\mu}, see (3.3).

The symmetric part Q(ij)μQ^{\mu}_{(ij)} of QijμQ^{\mu}_{ij} is involved141414The assignment (ei,ej)Q(ij)μcμ(e_{i},e_{j})\mapsto Q^{\mu}_{(ij)}c_{\mu} does not give rise to a globally defined map. Change (xa,yi,zμ)(x^{a},y^{i},z^{\mu}) to (xa,yi,zμ+12uijμ)(x^{a},y^{i},z^{\mu}+\frac{1}{2}u^{\mu}_{ij}) gives another assignment. in the canonical map Θ2:A[2]E2\mathrm{\Theta}^{2}:A^{[2]}\rightarrow E^{2} (see equations (3.3)). It turns out that the remaining structure functions Qij,kμQ^{\mu}_{ij,k} alone do not define any geometric mapping. Instead, it is the functions

Q~ijkμ=Qij,kμQjklQliμQiklQljμ+QjiνQνkμQkaQjiμxa\tilde{Q}^{\mu}_{ijk}=Q^{\mu}_{ij,k}-Q^{l}_{jk}Q^{\mu}_{li}-Q^{l}_{ik}Q^{\mu}_{lj}+Q^{\nu}_{ji}Q^{\mu}_{\nu k}-Q^{a}_{k}\frac{\partial Q^{\mu}_{ji}}{\partial x^{a}} (3.20)

that give a mapping

δ:Γ(A)×Γ(A)×Γ(A)Γ(C),δ(ei,ej,ek)=12Q~ijkμcμ.\delta:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(A)\rightarrow\operatorname{\Gamma}(C),\delta(e_{i},e_{j},e_{k})=\frac{1}{2}\tilde{Q}^{\mu}_{ijk}c_{\mu}. (3.21)

A coordinate-free definition of δ\delta is

δ(s1,s2,s)=12[s11,[s21,s0]]𝔛2(E2)Γ(C).\delta(s_{1},s_{2},s)=\frac{1}{2}\,[{s_{1}}^{\langle{-1}\rangle},[{s_{2}}^{\langle{-1}\rangle},{s}^{\langle{0}\rangle}]]\in\mathfrak{X}_{-2}(E^{2})\simeq\operatorname{\Gamma}(C). (3.22)

It is just a matter of direct computation of Lie brackets to show (3.21). Introduce

δssym(s1,s2):=12δ(s1,s2,s)+12δ(s2,s1,s) and δsalt(s1,s2):=12δ(s1,s2,s)12δ(s2,s1,s),{\delta}^{\mathrm{sym}}_{s}(s_{1},s_{2}):=\frac{1}{2}\delta(s_{1},s_{2},s)+\frac{1}{2}\delta(s_{2},s_{1},s)\,\text{ and }\quad{\delta}^{\mathrm{alt}}_{s}(s_{1},s_{2}):=\frac{1}{2}\delta(s_{1},s_{2},s)-\frac{1}{2}\delta(s_{2},s_{1},s), (3.23)

so δ=δalt+δsym\delta={\delta}^{\mathrm{alt}}+{\delta}^{\mathrm{sym}}. The skew-symmetric part δsalt{\delta}^{\mathrm{alt}}_{s} of δ(,,s)\delta(\cdot,\cdot,s) satisfies

δsalt(s1,s2):=12δ(s1,s2,s)12δ(s2,s1,s)=14[[s11,s21],s0]=12sβ(s1,s2).{\delta}^{\mathrm{alt}}_{s}(s_{1},s_{2}):=\frac{1}{2}\delta(s_{1},s_{2},s)-\frac{1}{2}\delta(s_{2},s_{1},s)={\frac{1}{4}}[[{s_{1}}^{\langle{-1}\rangle},{s_{2}}^{\langle{-1}\rangle}],{s}^{\langle{0}\rangle}]={-\frac{1}{2}}\Box_{s}\beta(s_{1},s_{2}). (3.24)

Further decomposition of δsym{\delta}^{\mathrm{sym}} by means of the Schur decomposition VSym2V=Sym3VWV\otimes\operatorname{Sym}^{2}V=\operatorname{Sym}^{3}V\oplus W (where WW is the kernel of the total symmetrization map) yields no additional information as g𝕊3δ(sg(1),sg(2),sg(3))=0\sum_{g\in\mathbb{S}_{3}}\delta(s_{g(1)},s_{g(2)},s_{g(3)})=0 due to the Jacobi identity for vector fields. It will be convenient to work with

ω:Γ(A)×Γ(A)×Γ(A)Γ(C),ω(s1,s2,s)=δ(s1,s2,s)β(s1,[s2,s])\omega:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(A)\rightarrow\operatorname{\Gamma}(C),\quad\omega(s_{1},s_{2},s)=\delta(s_{1},s_{2},s)-\beta(s_{1},[s_{2},s]) (3.25)

and its symmetric part

ωssym(s1,s2):=12ω(s1,s2,s)+12ω(s2,s1,s){{\omega}^{\mathrm{sym}}_{s}(s_{1},s_{2}):=\frac{1}{2}\omega(s_{1},s_{2},s)+\frac{1}{2}\omega(s_{2},s_{1},s)} (3.26)

instead of δ\delta and δsym{\delta}^{\mathrm{sym}}. In local coordinates we have ωeksym(ei,ej):=12ω¯ij,kμcμ{\omega}^{\mathrm{sym}}_{e_{k}}(e_{i},e_{j}):={\frac{1}{2}}\bar{\omega}_{ij,k}^{\mu}c_{\mu} where

ω¯ij,kμ=Qij,kμ+QkjlQ(il)μ+QkilQ(jl)μ+QjaQiklxaQlμ+QiaQjklxa+QkaQ(ij)μxa.\bar{\omega}_{ij,k}^{\mu}=Q^{\mu}_{ij,k}+Q^{l}_{kj}Q^{\mu}_{(il)}+Q^{l}_{ki}Q^{\mu}_{(jl)}+Q^{a}_{j}\frac{\partial Q^{l}_{ik}}{\partial x^{a}}Q^{\mu}_{l}+Q^{a}_{i}\frac{\partial Q^{l}_{jk}}{\partial x^{a}}+Q^{a}_{k}\frac{Q^{\mu}_{(ij)}}{\partial x^{a}}. (3.27)

Note that ωalt:=ωωsym{\omega}^{\mathrm{alt}}:=\omega-{\omega}^{\mathrm{sym}} satisfies

ωsalt(s1,s2)=12sβ(s1,s2)12β(s1,[s2,s])+12β(s2,[s1,s]),{\omega}^{\mathrm{alt}}_{s}(s_{1},s_{2})=-\frac{1}{2}\Box_{s}\beta(s_{1},s_{2})-\frac{1}{2}\beta(s_{1},[s_{2},s])+\frac{1}{2}\beta(s_{2},[s_{1},s]), (3.28)

due to (3.24).

Example 3.8.

We shall describe the structure maps of (A[2],κ[2])(A^{[2]},\kappa^{[2]}) – the second order prolongation A[2]A^{[2]} of an AL algebroid (A,κ)(A,\kappa). In standard coordinates (xa,yi,dxa,dyi)(x^{a},y^{i},dx^{a},dy^{i}) on TA\mathrm{T}A it is given locally by the equations dxa=Qiayidx^{a}=Q^{a}_{i}y^{i}, hence (xa,yi,dyi)(x^{a},y^{i},dy^{i}) form a coordinate chart for A[2]A^{[2]}. The coordinate description of κ[2]T2A×TA[2]\kappa^{[2]}\subseteq\mathrm{T}^{2}A\times\mathrm{T}A^{[2]} is

κ[2]:{x˙a=Qiay¯ix¨a=12Qwidehatijay¯iy¯j+Qiady¯i,x¯˙a=Qiayiy¯˙i=y˙i+Qjkiy¯jyk(dy¯l)=y¨l+Qijldy¯iyj+Qijly¯iy˙j+12Qwidehatij,kly¯iy¯jyk,\kappa^{[2]}:\begin{cases}\dot{x}^{a}=&Q^{a}_{i}\,\underline{y}^{i}\\ \ddot{x}^{a}=&\frac{1}{2}\widehat{Q}^{a}_{ij}\,\underline{y}^{i}\underline{y}^{j}+Q^{a}_{i}\,\underline{dy}^{i},\\ \dot{\underline{x}}^{a}=&Q^{a}_{i}\,y^{i}\\ \dot{\underline{y}}^{i}=&\dot{y}^{i}+Q^{i}_{jk}\,\underline{y}^{j}y^{k}\\ \left(\underline{dy}^{l}\right)^{\cdot}=&\ddot{y}^{l}+Q^{l}_{ij}\,\underline{dy}^{i}\,y^{j}+Q^{l}_{ij}\,\underline{y}^{i}\dot{y}^{j}+\frac{1}{2}\widehat{Q}^{l}_{ij,k}\underline{y}^{i}\underline{y}^{j}y^{k},\end{cases} (3.29)

where Qwidehatija\widehat{Q}^{a}_{ij} are defined in (3.7) and

Qwidehatij,kl=QiklxaQja+QjklxaQia.\widehat{Q}^{l}_{ij,k}=\frac{\partial Q^{l}_{ik}}{\partial x^{a}}Q^{a}_{j}+\frac{\partial Q^{l}_{jk}}{\partial x^{a}}Q^{a}_{i}. (3.30)

We find that :AA[2]widehat\partial:A\rightarrow\widehat{A^{[2]}} defined in (3.15), coincides with the identity on AA, with respect to the isomorphism given in Lemma 2.16,

=idA:AA[2]widehatA.\partial=\operatorname{id}_{A}:A\rightarrow\widehat{A^{[2]}}\simeq A.

Moreover, (A[2],κ[2])(A^{[2]},\kappa^{[2]}) is Lie, so κ[2]{\overrightarrow{\kappa^{[2]}}} is a Lie algebra morphism. Hence, β(s1,s2)=12[s11,s21]=12[s1,s2]A2\beta(s_{1},s_{2})=\frac{1}{2}[{s_{1}}^{\langle{-1}\rangle},{s_{2}}^{\langle{-1}\rangle}]=\frac{1}{2}{[s_{1},s_{2}]_{A}}^{\langle{-2}\rangle}, see Theorem 2.11. Thus we may write β(s1,s2)=[s1,s2]A\beta(s_{1},s_{2})=[s_{1},s_{2}]_{A} up to the isomorphism A[2]widehatA\widehat{A^{[2]}}\simeq A. Similarly, sv=[s,v]A\Box_{s}v=[s,v]_{A}, δ(s1,s2,s)=[[s1,s2]A,s]A\delta(s_{1},s_{2},s)=[[s_{1},s_{2}]_{A},s]_{A} and ωalt=0{\omega}^{\mathrm{alt}}=0 from (3.28) and the Jacobi identity. Moreover, C=\sharp^{C}=\sharp by Lemma 2.16.

Remark 3.9.

We can analogously define the following structure maps for any order kk HA (Ek,κk)(E^{k},\kappa^{k}):

ϕα1,,αn(s1,,sn)=1k![[[s1α1,s2α2],s3α3],,snαn]Γ(Ekwidehat),\phi_{{\alpha}_{1},\ldots,{\alpha}_{n}}(s_{1},\ldots,s_{n})=\frac{1}{k!}[\ldots[[{s_{1}}^{\langle{-{\alpha}_{1}}\rangle},{s_{2}}^{\langle{-{\alpha}_{2}}\rangle}],{s_{3}}^{\langle{-{\alpha}_{3}}\rangle}],\ldots,{s_{n}}^{\langle{-{\alpha}_{n}}\rangle}]\in\operatorname{\Gamma}(\widehat{E^{k}}),

where siΓ(A)s_{i}\in\operatorname{\Gamma}(A) and 0αi0\leq{\alpha}_{i} are such that i=1nαi=k\sum_{i=1}^{n}{\alpha}_{i}=k. In particular, s1k!sks\mapsto\frac{1}{k!}{s}^{\langle{-k}\rangle} defines a VB morphism

k:AEkwidehat,Γ(Ekwidehat)𝔛k(Ek).\partial^{k}:A\to\widehat{E^{k}},\quad\operatorname{\Gamma}(\widehat{E^{k}})\simeq\mathfrak{X}_{-k}(E^{k}).

Moreover, we have the structure maps

:Γ(A)×Γ(Ekwidehat)Γ(Ekwidehat),sv=[s[0],v]Γ(Ekwidehat)\Box:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(\widehat{E^{k}})\rightarrow\operatorname{\Gamma}(\widehat{E^{k}}),\quad\Box_{s}v=[{s}^{[0]},v]\in\operatorname{\Gamma}(\widehat{E^{k}})\\

and

Ek:EkwidehatTM\sharp^{E^{k}}:\widehat{E^{k}}\rightarrow\mathrm{T}M (3.31)

defined as the core of the anchor map k:EkTkM\sharp^{k}:E^{k}\rightarrow\mathrm{T}^{k}M composed with the isomorphism TkMwidehatTM\widehat{\mathrm{T}^{k}M}\simeq\mathrm{T}M given in (2.7). If (Ek,κk)(E^{k},\kappa^{k}) is Lie then, due to Theorem 2.11, ϕα1,,αn(s1,,sn)=1k!([[[s1,s2]A,s3]A,,sn]A)k\phi_{{\alpha}_{1},\ldots,{\alpha}_{n}}(s_{1},\ldots,s_{n})=\frac{1}{k!}{\left([\ldots[[s_{1},s_{2}]_{A},s_{3}]_{A},\ldots,s_{n}]_{A}\right)}^{\langle{-k}\rangle}, so all the structure maps ϕα1,,αn\phi_{{\alpha}_{1},\ldots,{\alpha}_{n}} with fixed nn coincide with ϕk,0,,0\phi_{k,0,\ldots,0}.

Let us assume that (A,,[,]A)(A,\sharp,[\cdot,\cdot]_{A}) is a Lie algebroid. Then (A[k],κ[k])(A^{[k]},\kappa^{[k]}) is a Lie HA. The map ϕk:Γ(A)Γ(A[k]widehat)\phi_{k}:\operatorname{\Gamma}(A)\rightarrow\operatorname{\Gamma}(\widehat{A^{[k]}}) gives the identification AA[k]widehatTk1AA\simeq\widehat{A^{[k]}}\subset\mathrm{T}^{k-1}A which coincides with ȷA[k1]:AA[k]widehatTk1A\jmath^{[k-1]}_{A}:A\rightarrow\widehat{A^{[k]}}\subset\mathrm{T}^{k-1}A, and

ϕα1,,αn(s1,,sn)=[[[s1,s2]A,s3]A,,sn]A,\phi_{{\alpha}_{1},\ldots,{\alpha}_{n}}(s_{1},\ldots,s_{n})=[\ldots[[s_{1},s_{2}]_{A},s_{3}]_{A},\ldots,s_{n}]_{A},

while sv=[s,v]A\Box_{s}v=[s,v]_{A}.

We introduce a few additional maps, denoted by ξ\xi, ψ\psi, ε\varepsilon, ε0\varepsilon_{0}, ε1\varepsilon_{1}, associated with a skew HA of order two. It will turn out that if (E2,κ2)(E^{2},\kappa^{2}) is AL then all these maps, except for ε1\varepsilon_{1}, vanish. If (E2,κ2)(E^{2},\kappa^{2}) is Lie then also ε1\varepsilon_{1} is zero. These maps will be used in formulation of tensor-like properties of the structure maps we have already introduced.

Definition 3.10.

Let (E2,κ2)(E^{2},\kappa^{2}) be a skew HA, s1,s2Γ(A)s_{1},s_{2}\in\operatorname{\Gamma}(A), f𝒞(M)f\in\mathcal{C}^{\infty}(M). We define

ξ:Γ(A)×Γ(A)𝔛(M),ξ(s1,s2)=[s1,s2][s1,s2],\displaystyle\xi:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(A)\rightarrow\mathfrak{X}(M),\quad\xi(s_{1},s_{2})=\sharp[s_{1},s_{2}]-[\sharp s_{1},\sharp s_{2}], (3.32)
ψ:Γ(A)×Γ(A)𝔛(M),ψ(s1,s2)(f):=12s11s21((2)f¨)(s1)(s2)(f),\displaystyle\psi:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(A)\rightarrow\mathfrak{X}(M),\quad\psi(s_{1},s_{2})(f):={\frac{1}{2}}{s_{1}}^{\langle{-1}\rangle}{s_{2}}^{\langle{-1}\rangle}((\sharp^{2})^{\ast}\ddot{f})-(\sharp s_{1})(\sharp s_{2})(f), (3.33)
ε=C:ATM,\displaystyle\varepsilon=\sharp^{C}\circ\partial-\sharp:A\rightarrow\mathrm{T}M, (3.34)
εk(s1,s2)=[s1k,s22+k][s1,s2]2𝔛2(E2)Γ(C),\displaystyle\varepsilon_{k}(s_{1},s_{2})=[{s_{1}}^{\langle{-k}\rangle},{s_{2}}^{\langle{-2+k}\rangle}]-{[s_{1},s_{2}]}^{\langle{-2}\rangle}\in\mathfrak{X}_{-2}(E^{2})\simeq\operatorname{\Gamma}(C), (3.35)

where k=0k=0 or 11.

Lemma 3.11.

The maps ξ\xi, ψ\psi, ε\varepsilon, ε0\varepsilon_{0}, ε1\varepsilon_{1} introduced in Definition 3.10 have the following properties:

  1. (i)

    The maps ξ\xi and ε1\varepsilon_{1} are tensorial in both arguments, so they give rise to the VB morphisms ξ:2ATM\xi:\bigwedge^{2}A\rightarrow\mathrm{T}M and ε1:2AC\varepsilon_{1}:\bigwedge^{2}A\rightarrow C, respectively. Moreover, 12ε1(s1,s2)=β(s1,s2)([s1,s2])\frac{1}{2}\varepsilon_{1}(s_{1},s_{2})=\beta(s_{1},s_{2})-\partial([s_{1},s_{2}]).

  2. (ii)

    The map (s1,s2)ε0(s1,s2)(s_{1},s_{2})\mapsto\varepsilon_{0}(s_{1},s_{2}) is tensorial in s2s_{2}, bot not in s1s_{1}, in general. We have ε0(fs1,s2)fε0(s1,s2)=ε(s2)(f)(s1)\varepsilon_{0}(fs_{1},s_{2})-f\varepsilon_{0}(s_{1},s_{2})=\varepsilon(s_{2})(f)\partial(s_{1}). Moreover, 12ε0(s1,s2)=s1(s2)([s1,s2]){\frac{1}{2}}\varepsilon_{0}(s_{1},s_{2})=\Box_{s_{1}}(\partial s_{2})-\partial([s_{1},s_{2}]).

  3. (iii)

    ψ(s1,s2)\psi(s_{1},s_{2}) is a derivation, hence the codomain of ψ\psi is correctly defined. Moreover, ψ(s1,s2)\psi(s_{1},s_{2}) is tensorial in s1s_{1}, but it is not tensorial in s2s_{2}, in general. Namely,

    ψ(s1,gs2)=gψ(s1,s2)+(s1)(g)ε(s2),\psi(s_{1},gs_{2})=g\psi(s_{1},s_{2})+(\sharp s_{1})(g)\cdot{\varepsilon(s_{2})},

In coordinates,

ψ(ek,ek)=12(QkkμQμa+Qkka2QkbQkaxb)xa.\psi(e_{k^{\prime}},e_{k})={\frac{1}{2}}\left(Q^{\mu}_{k^{\prime}k}Q^{a}_{\mu}+Q^{a}_{k^{\prime}k}-2Q^{b}_{k^{\prime}}\frac{\partial Q^{a}_{k}}{\partial x^{b}}\right)\partial_{x^{a}}. (3.36)

The skew-symmetric part of ψ\psi, ψalt(s1,s2)=12ψ(s1,s2)12ψ(s2,s1){\psi}^{\mathrm{alt}}(s_{1},s_{2})=\frac{1}{2}\psi(s_{1},s_{2})-\frac{1}{2}\psi(s_{2},s_{1}), is expressed in terms of the other structure maps:

ψalt(s1,s2)=12C(β(s1,s2))12[s1,s2].{\psi}^{\mathrm{alt}}(s_{1},s_{2})={\frac{1}{2}}\sharp^{C}(\beta(s_{1},s_{2}))-{\frac{1}{2}}[\sharp s_{1},\sharp s_{2}]. (3.37)

The symmetric part of ψ\psi, ψsym=ψψalt{\psi}^{\mathrm{sym}}=\psi-{\psi}^{\mathrm{alt}}, writes in coordinates as151515 It is tempting to consider a mapping (ei,ej)Q𝑤𝑖𝑑𝑒ℎ𝑎𝑡ijaxa(e_{i},e_{j})\mapsto\widehat{Q}^{a}_{ij}\partial_{x^{a}}. However, one can easily check that it does not give rise to a globally defined map.

ψsym(ei,ej)=12(QμaQ(ij)μ+QijaQwidehatija)xa,{\psi}^{\mathrm{sym}}(e_{i},e_{j})={\frac{1}{2}}\left(Q^{a}_{\mu}Q^{\mu}_{(ij)}+Q^{a}_{ij}-\widehat{Q}^{a}_{ij}\right)\partial_{x^{a}}, (3.38)

where Q𝑤𝑖𝑑𝑒ℎ𝑎𝑡ija\widehat{Q}^{a}_{ij} are defined in (3.7). Moreover, the condition 2Θ2=[2]\sharp^{2}\circ\mathrm{\Theta}^{2}=\sharp^{[2]} (compare with Theorem 3.2) is equivalent to the conjunction ψsym=0{\psi}^{\mathrm{sym}}=0 and C=\sharp^{C}\circ\partial=\sharp.

The proof is given in Appendix, subsection 4.2.

It turns out that the map ψsym{\psi}^{\mathrm{sym}} corresponds to a certain graded bundle morphism. A slightly more general result holds:

Lemma 3.12.

(a) Let (AM,:ATM)(A\rightarrow M,\sharp:A\to\mathrm{T}M) be an anchored vector bundle, and let ρ:AC\rho:A\to C be a VB morphism . Then, symmetric maps Ψ:Γ(A)×Γ(A)Γ(C)\Psi:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(A)\rightarrow\operatorname{\Gamma}(C) satisfying

Ψ(s1,fs2)=fΨ(s1,s2)+(s1)(f)ρ(s2)\Psi(s_{1},fs_{2})=f\Psi(s_{1},s_{2})+(\sharp s_{1})(f)\rho(s_{2}) (3.39)

are in a one-to-one correspondence with graded bundle morphisms Φ:A[2]C[2]\Phi:A^{[2]}\rightarrow C_{[2]}. The corresponding graded bundle  morphism Φ:A[2]C[2]\Phi:A^{[2]}\to C_{[2]} has the local form

Φ(xa,yi,y˙i)=(ρiμ(x)y˙i+12Ψijμ(x)yiyj)cμ,\Phi(x^{a},y^{i},\dot{y}^{i})=\left(\rho^{\mu}_{i}(x)\dot{y}^{i}+\frac{1}{2}\Psi^{\mu}_{ij}(x)y^{i}y^{j}\right)c_{\mu}, (3.40)

where ρ(ei)=ρiμcμ\rho(e_{i})=\rho^{\mu}_{i}c_{\mu} and Ψ(ei,ej)=Ψijμcμ\Psi(e_{i},e_{j})=\Psi^{\mu}_{ij}c_{\mu}, and (ei)(e_{i}) (respectively, (cμ)(c_{\mu})) is a local frame of sections of the vector bundle AA (respectively, CC).

The proof is given in Appendix, subsection 4.2.

The structure maps defined above, β\beta, \Box, ψsym{\psi}^{\mathrm{sym}}, and ωsym{\omega}^{\mathrm{sym}} are not 𝒞(M)\mathcal{C}^{\infty}(M)-linear in general, but satisfy certain tensor-like identities presented in the following result.

Theorem 3.13 (order-two skew HAs).

(a) Let (E2,κ2)(E^{2},\kappa^{2}) be a skew higher algebroid of order two, A=E1A=E^{1}, C=E2𝑤𝑖𝑑𝑒ℎ𝑎𝑡C=\widehat{E^{2}}. Let vΓ(C)v\in\operatorname{\Gamma}(C), s,s1,s2Γ(A)s,s_{1},s_{2}\in\operatorname{\Gamma}(A), f𝒞(M)f\in\mathcal{C}^{\infty}(M). Then

  • The map β\beta is skew-symmetric and

    β(s1,fs2)=fβ(s1,s2)+(s1)(f)(s2).\beta(s_{1},f\,s_{2})=f\,\beta(s_{1},s_{2})+(\sharp s_{1})(f)\partial(s_{2}).
  • The map (s,v)sv(s,v)\mapsto\Box_{s}v satisfies

    fsv\displaystyle\Box_{f\,s}v =fsv(Cv)(f)(s),\displaystyle=f\Box_{s}v-(\sharp^{C}v)(f)\,\partial(s),
    s(fv)\displaystyle\Box_{s}(fv) =fsv+(s)(f)v.\displaystyle=f\Box_{s}v+(\sharp s)(f)v.
  • The symmetric map ψsym{\psi}^{\mathrm{sym}} satisfies

    ψsym(s1,fs2)=fψsym(s1,s2)+12(s1)(f)ε(s2).{\psi}^{\mathrm{sym}}(s_{1},f\,s_{2})=f\,{\psi}^{\mathrm{sym}}(s_{1},s_{2})+\frac{1}{2}(\sharp s_{1})(f)\varepsilon(s_{2}).
  • The map ωssym(s1,s2){\omega}^{\mathrm{sym}}_{s}(s_{1},s_{2}) is symmetric in s1s_{1}, s2s_{2} and satisfies

    ωssym(s1,fs2)=fωssym(s1,s2)12(s1)(f)ε0(s,s2)+12ξ(s,s1)(f)(s2),{\omega}^{\mathrm{sym}}_{s}(s_{1},fs_{2})=f{\omega}^{\mathrm{sym}}_{s}(s_{1},s_{2})-\frac{1}{2}(\sharp s_{1})(f)\cdot{\varepsilon_{0}(s,s_{2})}+{\frac{1}{2}}\xi(s,s_{1})(f)\partial(s_{2}),
    ωfssym(s1,s2)=fωssym(s1,s2)+14(s1(f)ε1(s2,s)+s2(f)ε1(s1,s))+ψsym(s1,s2)(f)(s),{\omega}^{\mathrm{sym}}_{fs}(s_{1},s_{2})=f{\omega}^{\mathrm{sym}}_{s}(s_{1},s_{2})+\frac{1}{4}\left(\sharp s_{1}(f)\varepsilon_{1}(s_{2},s)+\sharp s_{2}(f)\varepsilon_{1}(s_{1},s)\right)+{{\psi}^{\mathrm{sym}}(s_{1},s_{2})(f)\partial(s)},

    (The maps ε\varepsilon, ε0\varepsilon_{0}, ε1\varepsilon_{1}, ξ\xi, and ψsym{\psi}^{\mathrm{sym}} are as in Definition 3.10.)

(b) Conversely, let (A,[,],)(A,[\cdot,\cdot],\sharp) be a skew algebroid and CMC\rightarrow M be a vector bundle. Then a system of the following maps:

  1. (i)

    VB morphisms :AC\partial:A\rightarrow C and C:CTM\sharp^{C}:C\to\mathrm{T}M covering the identity on MM,

  2. (ii)

    a skew-symmetric map β:Γ(A)×Γ(A)Γ(C)\beta:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(A)\rightarrow\operatorname{\Gamma}(C) satisfying (3.13),

  3. (iii)

    a map :Γ(A)×Γ(C)Γ(C)\Box:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(C)\to\operatorname{\Gamma}(C) satisfying (3.13) and (3.13),

  4. (iv)

    a symmetric map ψsym:Γ(A)×Γ(A)𝔛(M){\psi}^{\mathrm{sym}}:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(A)\to\mathfrak{X}(M) satisfying (3.13),

  5. (v)

    a map ωsym:Γ(A)×Sym2Γ(A)Γ(C){\omega}^{\mathrm{sym}}:\operatorname{\Gamma}(A)\times\operatorname{Sym}^{2}\operatorname{\Gamma}(A)\to\operatorname{\Gamma}(C) satisfying (3.13) and (3.13),

define a skew order-two HA on the graded bundle E2=(A[2]×MC[2])/(graph())E^{2}={\raisebox{1.79997pt}{$(A^{[2]}\times_{M}C_{[2]})$}\left/\raisebox{-1.79997pt}{$(\operatorname{graph}(-\partial))$}\right.} (see Lemma 3.6) uniquely. (Note that the maps ε\varepsilon, ε0\varepsilon_{0}, ε1\varepsilon_{1}, ξ\xi, which appear in the Leibniz-type identities of the structure maps listed here, can be expressed in terms of the aforementioned maps, see Definition 3.10 and Lemma 3.11.)

Proof.

The proof of part (a) – regarding the tensor-like properties of the structure maps β\beta, C\sharp^{C}, \Box, ψsym{\psi}^{\mathrm{sym}}, and ωsym{\omega}^{\mathrm{sym}} listed above – is technical and has been moved to Appendix, Subsection 4.2.

Proof of part (b): Let (A,[,],)(A,[\cdot,\cdot],\sharp) be a skew algebroid, and assume the structure maps listed above, \partial, β\beta, \Box, ψsym{\psi}^{\mathrm{sym}}, and ωsym{\omega}^{\mathrm{sym}}, are given.

Given the VB morphisms :AC\partial:A\to C and :ATM\sharp:A\to\mathrm{T}M, the construction of the graded bundle  E2E^{2} as the quotient of A[2]×MC[2]A^{[2]}\times_{M}C_{[2]} is well-founded, see Lemma 3.6. We shall now present the construction of the graded bundle morphism 2:E2T2M\sharp^{2}:E^{2}\rightarrow\mathrm{T}^{2}M.

There is a graded bundle morphism Φ:A[2](TM)[2]\Phi:A^{[2]}\rightarrow(\mathrm{T}M)_{[2]} corresponding to Ψ=ψsym\Psi={\psi}^{\mathrm{sym}} and ρ=ε\rho=\varepsilon, as explained in Lemma 3.12. Define a map

~2:A[2]×MC[2]T2M,(a2,v)[2](a2)+(Cv+Φ(a2)),\tilde{\sharp}^{2}:A^{[2]}\times_{M}C_{[2]}\rightarrow\mathrm{T}^{2}M,\quad(a^{2},v)\mapsto\sharp^{[2]}(a^{2})\bm{\boldsymbol{+}}(\sharp^{C}v+\Phi(a^{2})), (3.41)

where a2A[2]a^{2}\in A^{[2]} and vCv\in C project to the same point in MM, and +\bm{\boldsymbol{+}} denotes the action of the core bundle T2MwidehatTM\widehat{\mathrm{T}^{2}M}\simeq\mathrm{T}M on T2M\mathrm{T}^{2}M. We shall show that this map factors through the action of the graph of -\partial, the subbundle of the core bundle A×MCA\times_{M}C, giving rise to a map from the quotient graded bundle E2E^{2}, constructed in Lemma 3.6. It remains to show that

[2](a2+b)+(C(vb)+Φ(a2+b))\sharp^{[2]}(a^{2}\bm{\boldsymbol{+}}b)\bm{\boldsymbol{+}}(\sharp^{C}(v-\partial b)+\Phi(a^{2}\bm{\boldsymbol{+}}b))

does not depend on bAb\in A. Indeed, the change in the core is equal to

[2]widehat(b)(C)(b)+Φwidehat(b)=bCb+ε(b)=0,\widehat{\sharp^{[2]}}(b)-(\sharp^{C}\circ\partial)(b)+\widehat{\Phi}(b)=\sharp b-\sharp^{C}\circ\partial b+\varepsilon(b)=0,

since [2]widehat=\widehat{\sharp^{[2]}}=\sharp and by the definition of ε\varepsilon. A direct calculation from the coordinate formulas (3.38) and (3.40) shows that the resulting map E2T2ME^{2}\to\mathrm{T}^{2}M is indeed given by the desired formula:

(2)(x¨a)=([2])(x¨2)+(C)(x˙a)+Φ(x˙a)=Qiay˙i+12Qwidehatijayiyj+Qμacμ+(QiμQμaQia)y˙i+12(QijaQwidehatija)yiyj=Qμawμ+12Qijayiyj,\begin{split}{(\sharp^{2})}^{*}(\ddot{x}^{a})={(\sharp^{[2]})}^{*}(\ddot{x}^{2})+{(\sharp^{C})}^{*}(\dot{x}^{a})+{\Phi}^{*}(\dot{x}^{a})=\\ Q^{a}_{i}\dot{y}^{i}+\frac{1}{2}\widehat{Q}^{a}_{ij}y^{i}y^{j}+Q^{a}_{\mu}c_{\mu}^{\ast}+(Q^{\mu}_{i}Q^{a}_{\mu}-Q^{a}_{i})\dot{y}^{i}+\frac{1}{2}(Q^{a}_{ij}-\widehat{Q}^{a}_{ij})y^{i}y^{j}=Q^{a}_{\mu}w^{\mu}+\frac{1}{2}Q^{a}_{ij}y^{i}y^{j},\end{split}

where (xa,yi,wμ)(x^{a},y^{i},w^{\mu}) is the adapted coordinate system on the quotient, so cμ=wμQiμy˙ic_{\mu}^{\ast}=w^{\mu}-Q^{\mu}_{i}\dot{y}^{i} and Q(ij)μ=0Q^{\mu}_{(ij)}=0, see Definition 3.7.

We show now how to recover the comorphism κ2\kappa^{2}, which covers the graded bundle 2\sharp^{2} and governs the HA structure on the graded bundle E2E^{2}. All the local structure functions (Q)=Qia(Q^{\cdots}_{\cdots})=Q^{a}_{i}, QijaQ^{a}_{ij}, QμaQ^{a}_{\mu}, QjkiQ^{i}_{jk}, QiμQ^{\mu}_{i}, QijμQ^{\mu}_{ij}, QνiμQ^{\mu}_{\nu i}, Qij,kμ)Q^{\mu}_{ij,k}) can be derived from the structure maps listed above once we fix a graded coordinate system (xa,yi,zμ)(x^{a},y^{i},z^{\mu}) on E2E^{2}. (All these functions are defined locally, over an open subset UMU\subset M.) Without loss of generality, we may assume that (xa,yi,zμ)(x^{a},y^{i},z^{\mu}) is an adapted coordinate system (Definition 3.7), so Q(ij)μ=0Q^{\mu}_{(ij)}=0.

The local structure functions QiaQ^{a}_{i}, QμaQ^{a}_{\mu}, and QijaQ^{a}_{ij} are derived from the map 2\sharp^{2}. Next, Q[ij]μ=QijμQ^{\mu}_{[ij]}=Q^{\mu}_{ij} and QνiμQ^{\mu}_{\nu i} are derived from the maps β\beta and \Box, respectively, by means of (3.19). Finally, Qij,kμQ^{\mu}_{ij,k} is determined from ωsym{\omega}^{\mathrm{sym}}, see (3.27).

The structure functions (Q)(Q^{\cdots}_{\cdots}) establish a HA structure (EU2,κU2)(E^{2}_{U},\kappa^{2}_{U}) over the base UU, through the equations (2.21), where EU2=(σ2)1(U)E^{2}_{U}=(\sigma^{2})^{-1}(U). Moreover, the comorphism κU2\kappa^{2}_{U} determines all the structure maps U\partial_{U}, UC\sharp^{C}_{U}, βU\beta_{U}, U\Box_{U}, ψU\psi_{U}, and ωUsym{\omega}^{\mathrm{sym}}_{U} which are defined on sections of the vector bundles σU1:AUU\sigma^{1}_{U}:A_{U}\rightarrow U and σU2widehat:CUU\widehat{\sigma^{2}_{U}}:C_{U}\rightarrow U. Also the other structure maps present in the formulation of our theorem, the maps εU\varepsilon_{U}, (ε0)U(\varepsilon_{0})_{U}, (ε1)U(\varepsilon_{1})_{U} and ξ\xi are determined by κU2\kappa^{2}_{U}, as explained in Definition 3.10.

These maps are consistent with the restrictions of the corresponding maps given at the outset as the latter are local operators and satisfy the same Leibniz-type identities. For example, U(ei,cμ)=|U(ei,cμ)\Box_{U}(e_{i},c_{\mu})=\Box|_{U}(e_{i},c_{\mu}) by (3.19), where on the RHS, |U\Box|_{U} denotes the given structure map \Box restricted to ΓU(A)×ΓU(C)\operatorname{\Gamma}_{U}(A)\times\operatorname{\Gamma}_{U}(C). Moreover, both U\Box_{U} and |U\Box|_{U} satisfy the same Leibniz-type identities given in (3.13) and (3.13), because U\partial_{U} coincides with |Γ(A)\partial|_{\operatorname{\Gamma}(A)}, and similarly for \sharp and C\sharp^{C}. Therefore, (U)(s|U,v|U)=sv(\Box_{U})(s|_{U},v|_{U})=\Box_{s}v, for any sΓ(A)s\in\operatorname{\Gamma}(A) and vΓ(C)v\in\operatorname{\Gamma}(C).

Therefore, if UUU\cap U^{\prime}\neq\emptyset, then (κU2)|UU(\kappa^{2}_{U})\Big{|}_{U^{\prime}\cap U} and (κU2)|UU(\kappa^{2}_{U^{\prime}})\Big{|}_{U\cap U^{\prime}} coincide with κUU2\kappa^{2}_{U\cap U^{\prime}}, which is defined by the restrictions of the structure maps to the sections over UUU\cap U^{\prime}. Therefore, κ2\kappa^{2} is globally well-defined. ∎

Remark 3.14.

The map β\beta in the formulation of part (b) of Theorem 3.13 can be replaced by the VB morphism ε1:2AC\varepsilon_{1}:\bigwedge^{2}A\rightarrow C, defined in (3.35). Indeed, the map β\beta is related to ε1\varepsilon_{1} via the formula given in Lemma 3.11, β(s1,s2)=12ε1(s1,s2)+([s1,s2])\beta(s_{1},s_{2})=\frac{1}{2}\varepsilon_{1}(s_{1},s_{2})+\partial([s_{1},s_{2}]). Hence, the Leibniz-type identity (3.13) follows from the Leibinz rule of the bracket [,][\cdot,\cdot] on Γ(A)\operatorname{\Gamma}(A). Note also that the anchor :ATM\sharp:A\to\mathrm{T}M is uniquely determined by the bracket [,][\cdot,\cdot] on Γ(A)\operatorname{\Gamma}(A).

3.2.3 HAs over a point

We shall study HAs (σk:EkM,κk)(\sigma^{k}:E^{k}\rightarrow M,\kappa^{k}) in which the base M={pt}M=\{\operatorname{pt}\} is a point. Any such structure is fully described by a weight-respecting mapping (see Theorem 2.11)

κk:Tk𝔤𝔛0(Ek){\overrightarrow{\kappa^{k}}}:\mathrm{T}^{k}\mathfrak{g}\to\mathfrak{X}_{\leq 0}(E^{k}) (3.42)

where the algebra (𝔤,[,])(\mathfrak{g},[\cdot,\cdot]) is defined as the order-one reduction of (Ek,κk)(E^{k},\kappa^{k}). Let (ei)(e_{i}) be a basis of the vector space E1=𝔤E^{1}=\mathfrak{g}, (yi)(y^{i}) be the corresponding dual basis and let (yi,zμ)(y^{i},z^{\mu}) be a graded coordinate system for EkE^{k} in which the weight w(yi)=1\mathrm{w}(y^{i})=1 and w(zμ)2\mathrm{w}(z^{\mu})\geq 2. To define a HA (σk:EkM,κk)(\sigma^{k}:E^{k}\rightarrow M,\kappa^{k}) it amounts to provide vector fields emα𝔛α(Ek){e_{m}}^{\langle{-{\alpha}}\rangle}\in\mathfrak{X}_{-{\alpha}}(E^{k}) for 0αk0\leq{\alpha}\leq k such that

em0\displaystyle{e_{m}}^{\langle{0}\rangle} =iQlmiylyi+μfmμ(y,z)zμ,\displaystyle=\sum_{i}Q^{i}_{lm}y^{l}\partial_{y^{i}}+\sum_{\mu}f^{\mu}_{m}(y,z)\partial_{z^{\mu}}, (3.43)
em1\displaystyle{e_{m}}^{\langle{-1}\rangle} =ym+μgmμ(y,z)zμ,\displaystyle=\partial_{y^{m}}+\sum_{\mu}g^{\mu}_{m}(y,z)\partial_{z^{\mu}}, (3.44)

where QlmiQ^{i}_{lm} are the structure constants for (𝔤,[,])(\mathfrak{g},[\cdot,\cdot]), i.e., [el,em]=Qlmiei[e_{l},e_{m}]=Q^{i}_{lm}e_{i}, see Theorem 2.11. It follows that fmμf^{\mu}_{m} (resp. gmμg^{\mu}_{m}) are homogeneous functions on EkE^{k} of weight w(zμ)\mathrm{w}(z^{\mu}) (resp., w(zμ)1\mathrm{w}(z^{\mu})-1). The obtained (general) HA is AL if and only if the bracket [,][\cdot,\cdot] is skew-symmetric.

Order two.

The map Θk:Tk1𝔤Ek\mathrm{\Theta}^{k}:\mathrm{T}^{k-1}\mathfrak{g}\rightarrow E^{k} covers Θ1=id𝔤\mathrm{\Theta}^{1}=\operatorname{id}_{\mathfrak{g}}, hence it gives a canonical section of the bundle projection σ1k:EkE1=𝔤\sigma^{k}_{1}:E^{k}\rightarrow E^{1}=\mathfrak{g}. In case k=2k=2, σ12:E2𝔤\sigma^{2}_{1}:E^{2}\rightarrow\mathfrak{g} is an affine bundle projection, hence the mapping Θ2:T𝔤=𝔤[1]𝔤[2]E2\mathrm{\Theta}^{2}:\mathrm{T}\mathfrak{g}=\mathfrak{g}_{[1]}\oplus\mathfrak{g}_{[2]}\rightarrow E^{2} yields a canonical splitting

E2=𝔤[1]×C[2],E^{2}=\mathfrak{g}_{[1]}\times C_{[2]},

where C=E2widehatC=\widehat{E^{2}}, and Θ2(x,0)=(x,0)\mathrm{\Theta}^{2}(x,0)=(x,0) where x𝔤x\in\mathfrak{g}. We are going to describe the structure of the graded, finite dimensional Lie algebra 𝔛0:=𝔛0(E2)\mathfrak{X}_{\leq 0}:=\mathfrak{X}_{\leq 0}(E^{2}). In standard graded coordinates (yi,zμ)(y^{i},z^{\mu}) on 𝔤×C\mathfrak{g}\times C, vector fields of non-positive weight α{\alpha}, where2α0-2\leq{\alpha}\leq 0, have the following form

X0=cjiyjyi+(cνμzν+12cijμyiyj)zμ,X1=ciyi+ciμyizμ,X2=cμzμX_{0}=c_{j}^{i}y^{j}\partial_{y^{i}}+(c^{\mu}_{\nu}\,z^{\nu}+\frac{1}{2}\,c_{ij}^{\mu}\,y^{i}y^{j})\,\partial_{z^{\mu}},\quad X_{-1}=c^{i}\partial_{y^{i}}+c_{i}^{\mu}\,y^{i}\,\partial_{z^{\mu}},\quad X_{-2}=c^{\mu}\,\partial_{z^{\mu}}

where cc^{\cdots}_{\cdots} are some constants, and Xα𝔛α(E2)X_{\alpha}\in\mathfrak{X}_{\alpha}(E^{2}). The Lie algebra 𝔛0\mathfrak{X}_{\leq 0} acts faithfully on the linear subspace 𝒜2𝒞(𝔤×C)\mathcal{A}_{\leq 2}\subseteq\mathcal{C}^{\infty}(\mathfrak{g}\times C), spanned by homogenous functions of weight 2\leq 2. It has a \mathbb{R}-basis consisting of the functions 1,yi,yiyj,zμ1,y^{i},y^{i}y^{j},z^{\mu} and we have 𝒜2𝔤CSym2𝔤\mathcal{A}_{\leq 2}\simeq\mathbb{R}\oplus\mathfrak{g}^{\ast}\oplus C^{\ast}\oplus\operatorname{Sym}^{2}\mathfrak{g}^{\ast}. By examining the action of the vector fields X0X_{0}, X1X_{1}, X2X_{2}, we easily find the following decomposition (compare with a more general Lemma 4.1),

𝔛0End(𝔤)End(C)Hom(Sym2𝔤,C),𝔛1𝔤Hom(𝔤,C),𝔛2C.\mathfrak{X}_{0}\simeq\operatorname{End}(\mathfrak{g})\oplus\operatorname{End}(C)\oplus\operatorname{Hom}(\operatorname{Sym}^{2}\mathfrak{g},C),\quad\mathfrak{X}_{-1}\simeq\mathfrak{g}\oplus\operatorname{Hom}(\mathfrak{g},C),\quad\mathfrak{X}_{-2}\simeq C. (3.45)

The formula for the Lie bracket on 𝔛0\mathfrak{X}_{\leq 0} will be given in the proof of Theorem 3.15, see (4.9).

We shall describe algebroid lifts eα{e}^{\langle{-{\alpha}}\rangle}, where α{0,1,2}{\alpha}\in\{0,1,2\}, by means of the structure maps of (E2,κ2)(E^{2},\kappa^{2}). Then it will be straightforward to verify the condition given in Remark 2.12, ensuring that (E2,κ2)(E^{2},\kappa^{2}) is a Lie HA.

Theorem 3.15.

The structure of a skew, order-two HA over a point is fully determined by the linear maps [,]:2𝔤𝔤[\cdot,\cdot]:\bigwedge^{2}\mathfrak{g}\rightarrow\mathfrak{g}, :𝔤C\partial:\mathfrak{g}\rightarrow C, β:2𝔤C\beta:\bigwedge^{2}\mathfrak{g}\rightarrow C, :𝔤CC\Box:{\mathfrak{g}\otimes C}\rightarrow C, ωsym:𝔤Sym2𝔤C{\omega}^{\mathrm{sym}}:\mathfrak{g}\otimes\operatorname{Sym}^{2}\mathfrak{g}\rightarrow C. The associated algebroid lifts eeα𝔛αe\mapsto{e}^{\langle{-{\alpha}}\rangle}\in\mathfrak{X}_{-\alpha} are given (with respect to the isomorphisms listed in (3.45)) by

e0\displaystyle{e}^{\langle{0}\rangle} =[,e]e()2ωesym(,),\displaystyle=[\cdot,e]\oplus{\Box_{-e}(\cdot)}\oplus{2{\omega}^{\mathrm{sym}}_{e}(\cdot,\cdot),}
e1\displaystyle{e}^{\langle{-1}\rangle} =eβ(,e),\displaystyle=e\oplus\beta(\cdot,e),
e2\displaystyle{e}^{\langle{-2}\rangle} =(e)\displaystyle=\partial(e)

A skew HA (𝔤×C,[,],,β,,ωsym)(\mathfrak{g}\times C,[\cdot,\cdot],\partial,\beta,\Box,{\omega}^{\mathrm{sym}}) is Lie if and only if 𝔤\mathfrak{g} is a Lie algebra, \Box equips CC with a 𝔤\mathfrak{g}-module structure, :𝔤C\partial:{\mathfrak{g}}\rightarrow C is a 𝔤\mathfrak{g}-module morphism, ωsym=0{\omega}^{\mathrm{sym}}=0 and the mapping β\beta is given by

β(x1,x2)=([x1,x2]).\beta(x_{1},x_{2})=\partial([x_{1},x_{2}]).

Hence, order-two Lie higher algebroids over a point are in a one-to-one correspondence with 𝔤\mathfrak{g}-module morphisms :𝔤C\partial:{\mathfrak{g}}\rightarrow{C}.

The proof is straightforward but somewhat lengthy, so it has been moved to Appendix, Subsection 4.2.

Order greater than two.

The graded bundle hosting a higher Lie algebroid over a point of order greater than two need not split in a canonical way. A simple example is provided by a non-split graded space EkE^{k} such that E1={0}E^{1}=\{0\} – a vector space of dimension 0. In this case, the VB comorphism κk{\overrightarrow{\kappa^{k}}} must be the zero map, i.e., (κk)a:Tk{0}TaEk(\kappa^{k})_{a}:\mathrm{T}^{k}\{0\}\rightarrow\mathrm{T}_{a}E^{k} is the zero map for any aEka\in E^{k}, as the domain is zero-dimensional. A concrete example of this is E4=Tq2ME^{4}=\mathrm{T}_{q}^{2}M, where qMq\in M is a fixed point on a manifold MM, and the linear coordinates (x˙a)(\dot{x}^{a}) on TqM\mathrm{T}_{q}M are assigned weight 22, while the weight of x¨a\ddot{x}^{a} is 44.

Following the example given in [JR18, Section 6], a graded Lie algebra i=0k1𝔤i\bigoplus_{i=0}^{k-1}\mathfrak{g}_{i}, where 𝔤0=𝔤\mathfrak{g}_{0}=\mathfrak{g}, equipped with a graded Lie algebra morphism A:Tk1𝔤i=0k1𝔤iA:\mathrm{T}^{k-1}\mathfrak{g}\rightarrow\bigoplus_{i=0}^{k-1}\mathfrak{g}_{i} such that A0=id𝔤0A^{0}=\operatorname{id}_{\mathfrak{g}_{0}}, gives rise to a split Lie higher algebroid of order kk.

3.2.4 AL HAs of order two

There are a few relations among the structure maps of a skew HA (E2,κ2)(E^{2},\kappa^{2}) that we introduced above, ensuring that it is an almost Lie algebroid.

Theorem 3.16 (order-two AL HAs).

Let (E2,κ2)(E^{2},\kappa^{2}) be a skew order-two HA. Then (E2,κ2)(E^{2},\kappa^{2}) is AL if and only if

A is an almost Lie algebroid, i.e., ξ=0,\displaystyle\text{$A$ is an almost Lie algebroid, i.e., $\xi=0$}, (ALA\mathrm{AL_{A}})
=C,i.e., ε=0,\displaystyle\sharp=\sharp^{C}\circ\partial,\text{i.e., }\varepsilon=0, (AL\mathrm{AL_{\partial}})
ψ=0,\displaystyle\psi=0, (ALψ\mathrm{AL_{\psi}})
C(sv)=[s,Cv],\displaystyle\sharp^{C}(\Box_{s}v)=[\sharp s,\sharp^{C}v], (AL\mathrm{AL_{\Box}})
Cω=0.\displaystyle\sharp^{C}\circ\omega=0. (ALω\mathrm{AL_{\omega}})
Corollary 3.17.

An order-two AL HA on a graded bundle E2ME^{2}\rightarrow M is defined by:

  • an AL algebroid structure on the vector bundle AMA\rightarrow M;

  • VB morphisms \partial, C\sharp^{C}, and ε1\varepsilon_{1},

  • maps \Box and ωsym{\omega}^{\mathrm{sym}} that satisfy the aforementioned Leibniz-type identities,

such that conditions (AL\mathrm{AL_{\Box}}) and (AL\mathrm{AL_{\partial}}) are satisfied, and the images of the maps ε1\varepsilon_{1} and ωsym{\omega}^{\mathrm{sym}} lie in the kernel of the VB morphism C\sharp^{C}.

Proof.

This corollary follows directly from Theorem 3.16 and Remark 3.14:

Assume (E2,κ2)(E^{2},\kappa^{2}) is an AL HA. Then ψalt=0{\psi}^{\mathrm{alt}}=0 follows from (ALψ\mathrm{AL_{\psi}}), and the identity

C(β(s1,s2))=[s1,s2]\sharp^{C}(\beta(s_{1},s_{2}))=\sharp[s_{1},s_{2}]

follows from (3.37) and (ALA\mathrm{AL_{A}}). Therefore, 12Cε1=CβC[,]=(C)[,]=0\frac{1}{2}\sharp^{C}\circ\varepsilon_{1}=\sharp^{C}\circ\beta-\sharp^{C}\circ\partial\circ[\cdot,\cdot]=(\sharp-\sharp^{C}\circ\partial)\circ[\cdot,\cdot]=0, due to (AL\mathrm{AL_{\partial}}). Clearly,

Cωsym=0.\sharp^{C}\circ{\omega}^{\mathrm{sym}}=0.

follows from (ALω\mathrm{AL_{\omega}}).

Conversely, from Theorem 3.16 and Remark 3.14, it follows that the maps listed in Corollary 3.17 define a skew HA (with ψsym=0{\psi}^{\mathrm{sym}}=0).

The identity (3.2.4) holds by the assumption Cε1=0\sharp^{C}\circ\varepsilon_{1}=0, (ALA\mathrm{AL_{A}}), and (AL\mathrm{AL_{\partial}}). Hence ψalt=0{\psi}^{\mathrm{alt}}=0, and (ALψ\mathrm{AL_{\psi}}) follows from the assumption ωsym=0{\omega}^{\mathrm{sym}}=0.

We have, Cωalt(s1,s2,s)=12Jac(s1,s2,s)\sharp^{C}\circ{\omega}^{\mathrm{alt}}(s_{1},s_{2},s)=\frac{1}{2}\sharp\circ\operatorname{Jac}(s_{1},s_{2},s), from (3.28), (3.2.4), and (AL\mathrm{AL_{\Box}}), where

Jac(X1,X2,X)=[X1,[X2,X]][X2,[X1,X]][[X1,X2],X].\operatorname{Jac}(X_{1},X_{2},X)=[X_{1},[X_{2},X]]-[X_{2},[X_{1},X]]-[[X_{1},X_{2}],X]. (3.46)

Moreover, Jac(s1,s2,s)=Jac(s1,s2,s)=0\sharp\circ\operatorname{Jac}(s_{1},s_{2},s)=\operatorname{Jac}(\sharp s_{1},\sharp s_{2},\sharp s)=0 since AA is an AL algebroid. Therefore, Cωalt=0\sharp^{C}\circ{\omega}^{\mathrm{alt}}=0 and (ALω\mathrm{AL_{\omega}}) follows from the assumption Cωsym=0\sharp^{C}\circ{\omega}^{\mathrm{sym}}=0. ∎

Remark 3.18.

The equation (AL\mathrm{AL_{\partial}}) implies that ψ\psi is 𝒞(M)\mathcal{C}^{\infty}(M)-linear in both arguments. Also the difference between left and right hand side in (AL\mathrm{AL_{\Box}}) is 𝒞(M)\mathcal{C}^{\infty}(M)-linear in ss and vv thanks to (AL\mathrm{AL_{\partial}}). Moreover, in this case, also ωsym{\omega}^{\mathrm{sym}} and ω\omega are 𝒞(M)\mathcal{C}^{\infty}(M)-linear in all arguments, see (3.13), (3.13). Therefore, it is enough to check the condition listed in Theorem 3.16 for arguments from local frames (ei)(e_{i}), (cμ)(c_{\mu}) of Γ(A)\operatorname{\Gamma}(A) and Γ(C)\operatorname{\Gamma}(C), respectively.

Remark 3.19.

Note that in the AL case, (3.41) simplifies to ~2(a2,v)=[2](a2)C(v)\tilde{\sharp}^{2}(a^{2},v)=\sharp^{[2]}(a^{2})\oplus\sharp^{C}(v), giving rise to the 2nd2^{\mathrm{nd}}-order anchor map 2:E2T2M\sharp^{2}:E^{2}\rightarrow\mathrm{T}^{2}M.

Proof of Theorem 3.16.

Let (E2,κ2)(E^{2},\kappa^{2}) be an almost Lie HA. Point (ALA\mathrm{AL_{A}}) is part of the definition of an AL algebroid. We shall show (3.2.4). From Theorem 3.2 (b) and Lemma 3.11 we obtain ψsym=0{\psi}^{\mathrm{sym}}=0 and (AL\mathrm{AL_{\partial}}). We shall show (3.2.4) from which the condition ψsym{\psi}^{\mathrm{sym}} follows, hence (ALψ\mathrm{AL_{\psi}}), see again Lemma 3.11.

It is well known that if Xi𝔛(M)X_{i}\in\mathfrak{X}(M), Yi𝔛(N)Y_{i}\in\mathfrak{X}(N), and f:MNf:M\rightarrow N is a differentiable mapping such that the vector fields Xi,YiX_{i},Y_{i} are ff-related for i=1,2i=1,2 then [X1,X2][X_{1},X_{2}] and [Y1,Y2][Y_{1},Y_{2}]-are ff-related, as well. Hence, using Theorem 2.11, we find that the vector fields [s11,s21]𝔛(E2)[{s_{1}}^{\langle{-1}\rangle},{s_{2}}^{\langle{-1}\rangle}]\in\mathfrak{X}(E^{2}) and [(s1)1,(s2)1]𝔛(T2M)[{(\sharp s_{1})}^{\langle{-1}\rangle},{(\sharp s_{2})}^{\langle{-1}\rangle}]\in\mathfrak{X}(\mathrm{T}^{2}M) are 2\sharp^{2}-related. On the other hand, 12[s11,s21]\frac{1}{2}[{s_{1}}^{\langle{-1}\rangle},{s_{2}}^{\langle{-1}\rangle}] is a vector field of weight 2-2 corresponding to the section β(s1,s2)Γ(C)\beta(s_{1},s_{2})\in\operatorname{\Gamma}(C) while [(s1)1,(s2)1]=[s1,s2]2[{(\sharp s_{1})}^{\langle{-1}\rangle},{(\sharp s_{2})}^{\langle{-1}\rangle}]={[\sharp s_{1},\sharp s_{2}]}^{\langle{-2}\rangle} as (T2M,κM2)(\mathrm{T}^{2}M,\kappa^{2}_{M}) is a Lie HA (see Example 2.9). The latter corresponds to the section 2[s1,s2]Γ(TM)Γ(T2Mwidehat)2[\sharp s_{1},\sharp s_{2}]\in\operatorname{\Gamma}(\mathrm{T}M)\simeq\operatorname{\Gamma}(\widehat{\mathrm{T}^{2}M}), see Lemma 2.16. Hence, using Lemma 2.6 and (3.14), we find that Cβ(s1,s2)=2widehat(β(s1,s2))=[s1,s2]\sharp^{C}\circ\beta(s_{1},s_{2})=\widehat{\sharp^{2}}(\beta(s_{1},s_{2}))=[\sharp s_{1},\sharp s_{2}]. This completes the proof of (3.2.4).

We shall prove the next two identities, (AL\mathrm{AL_{\Box}}) and (ALω\mathrm{AL_{\omega}}), in a similar way.

Consider vΓ(C)v\in\operatorname{\Gamma}(C) as the vector field v{v^{\uparrow}} on E2E^{2}. Then the vector fields vv and 12C(v)2𝔛2(T2M)\frac{1}{2}{\sharp^{C}(v)}^{\langle{-2}\rangle}\in\mathfrak{X}_{-2}(\mathrm{T}^{2}M) are 2\sharp^{2}-related, see Lemmas 2.6,  2.16 and the definition of C\sharp^{C}. Due to the AL-assumption on (E2,κ2)(E^{2},\kappa^{2}), the vector fields s0𝔛(E2){s}^{\langle{0}\rangle}\in\mathfrak{X}(E^{2}) and (s)0𝔛(T2M){(\sharp s)}^{\langle{0}\rangle}\in\mathfrak{X}(\mathrm{T}^{2}M) are 2\sharp^{2}-related, so the corresponding Lie brackets, i.e., the vector fields sv=[s0,v]𝔛2(E2){\Box_{s}v=[{s}^{\langle{0}\rangle},v]}\in\mathfrak{X}_{-2}(E^{2}) and [(s)0,12(Cv)2]𝔛2(T2M){[{(\sharp s)}^{\langle{0}\rangle},\frac{1}{2}{(\sharp^{C}v)}^{\langle{-2}\rangle}]}\in\mathfrak{X}_{-2}(\mathrm{T}^{2}M) are 2\sharp^{2}-related, as well. Since (T2M,κM2)(\mathrm{T}^{2}M,\kappa^{2}_{M}) is Lie, the latter vector field corresponds to [s,Cv]𝔛(M)[\sharp s,\sharp^{C}v]\in\mathfrak{X}(M) and (AL\mathrm{AL_{\Box}}) follows.

For (ALω\mathrm{AL_{\omega}}), due to the AL-assumption, 2δ(s1,s2,s)=[s11,[s21s0]]𝔛2(E2){2}\delta(s_{1},s_{2},s)=[{s_{1}}^{\langle{-1}\rangle},[{s_{2}}^{\langle{-1}\rangle}{s}^{\langle{0}\rangle}]]\in\mathfrak{X}_{-2}(E^{2}) is 2\sharp^{2}-related with

[(s1)1,[(s2)1,(s)0]]=[(s1)1,[(s2)1,(s)0]]=[s1,[s2,s]]2𝔛2(T2M).[{(\sharp s_{1})}^{\langle{-1}\rangle},[{(\sharp s_{2})}^{\langle{-1}\rangle},{(\sharp s)}^{\langle{0}\rangle}]]=[{(\sharp s_{1})}^{\langle{-1}\rangle},[{(\sharp s_{2})}^{\langle{-1}\rangle},{(\sharp s)}^{\langle{0}\rangle}]]={[\sharp s_{1},[\sharp s_{2},\sharp s]]}^{\langle{-2}\rangle}\in\mathfrak{X}_{-2}(\mathrm{T}^{2}M).

Hence, the sections δ(s1,s2,s)\delta(s_{1},s_{2},s) and 12[s1,[s2,s]]\frac{1}{2}[\sharp s_{1},[\sharp s_{2},\sharp s]] are C\sharp^{C}-related. Therefore,

Cω(s1,s2,s)=Cδ(s1,s2,s)Cβ(s1,[s2,s])=12[s1,[s2,s]12[s1,[s2,s]]=0\sharp^{C}\circ\omega(s_{1},s_{2},s)=\sharp^{C}\circ\delta(s_{1},s_{2},s)-\sharp^{C}\circ\beta(s_{1},[s_{2},s])=\frac{1}{2}[\sharp s_{1},[\sharp s_{2},\sharp s]-\frac{1}{2}\sharp[s_{1},[s_{2},s]]=0

due to (3.2.4) and (ALA\mathrm{AL_{A}}).

On the other hand, we assume that the structure maps of (E2,κ2)(E^{2},\kappa^{2}) satisfy the conditions (ALA\mathrm{AL_{A}})-(ALω\mathrm{AL_{\omega}}) and shall show that the vector fields sα{s}^{\langle{{\alpha}}\rangle} and (s)α{(\sharp s)}^{\langle{{\alpha}}\rangle}, both of weight α{\alpha}, are 2\sharp^{2}-related for α=2,1,0{\alpha}=-2,-1,0. This implies that (E2,κ2)(E^{2},\kappa^{2}) is almost Lie due to Theorem 2.11.

The case α=2{\alpha}=-2 is simple: (s)=12s[2]\partial(s)=\frac{1}{2}{s}^{[-2]} and 12(s)2\frac{1}{2}{(\sharp s)}^{\langle{-2}\rangle} are 2\sharp^{2}-related due to the relation (AL\mathrm{AL_{\partial}}).

Let α=1{\alpha}=-1. The vector field s1𝔛1(E2){s}^{\langle{-1}\rangle}\in\mathfrak{X}_{-1}(E^{2}) is projectable onto E1E^{1} and its projection (Tσ12)s1(\mathrm{T}\sigma^{2}_{1}){s}^{\langle{-1}\rangle} coincides with the (E1,κ1)(E^{1},\kappa^{1})-algebroid lift s1{s}^{\langle{-1}\rangle}, see Lemma 2.14. Similarly, (s)1𝔛2(T2M){(\sharp s)}^{\langle{-1}\rangle}\in\mathfrak{X}_{-2}(\mathrm{T}^{2}M) is also a vector field projectable onto TM\mathrm{T}M and it coincides with the tangent algebroid (TM,κM)(\mathrm{T}M,\kappa_{M})-lift (s)1{(\sharp s)}^{\langle{-1}\rangle}. Thus the case α=1{\alpha}=-1 reduces to the condition that (E1,κ1)(E^{1},\kappa^{1}) is AL, and we are done.

The proof in the case α=0{\alpha}=0 is given in Appendix, Subsection 4.3 where we perform direct calculations using coordinates. ∎

3.2.5 Lie HAs of order two

In the following result, we provide conditions (referred to as the axioms of Lie HAs) on the structure maps introduced earlier, ensuring that a given AL HA (E2,κ2)(E^{2},\kappa^{2}) is a Lie HA.

Theorem 3.20 (Lie HAs of order two).

Let (E2,κ2)(E^{2},\kappa^{2}) be an AL HA. Then (E2,κ2)(E^{2},\kappa^{2}) is Lie if and only if

AA is a Lie algebroid, (LieA\mathrm{Lie_{A}})
[s1,s2]v=s1s2vs2s1v,\displaystyle\Box_{[s_{1},s_{2}]}v=\Box_{s_{1}}\Box_{s_{2}}v-\Box_{s_{2}}\Box_{s_{1}}v, (Lie\mathrm{Lie_{\Box}})
([s1,s2])=s1(s2),i.e., ε0=0,\displaystyle\partial([s_{1},s_{2}])=\Box_{s_{1}}\partial(s_{2}),\text{i.e., $\varepsilon_{0}=0$,} (Lie\mathrm{Lie_{\partial}})
β(s1,s2)=([s1,s2]),i.e., ε1=0,\displaystyle\beta(s_{1},s_{2})=\partial([s_{1},s_{2}]),\text{i.e., $\varepsilon_{1}=0$,} (Lieβ\mathrm{Lie_{\beta}})
ω=0.\displaystyle\omega=0. (Lieω\mathrm{Lie_{\omega}})
Remark 3.21.

The condition (Lieω\mathrm{Lie_{\omega}}) can be replaced with

ωsym=0.{\omega}^{\mathrm{sym}}=0. (Lieω¯\mathrm{Lie_{\bar{\omega}}})

Indeed, vanishing of ωalt=ωωsym{\omega}^{\mathrm{alt}}=\omega-{\omega}^{\mathrm{sym}} follows from (3.28), (Lie\mathrm{Lie_{\partial}}) and (Lieβ\mathrm{Lie_{\beta}}). Thus, (Lieω\mathrm{Lie_{\omega}}) follows from (Lie\mathrm{Lie_{\partial}}), (Lieβ\mathrm{Lie_{\beta}}) and (Lieω¯\mathrm{Lie_{\bar{\omega}}}).

Remark 3.22.

It is a straightforward calculation to show that, in an AL HA, the mapping curv(s1,s2,v):=s1s2vs2s1v[s1,s2]v\mathrm{curv}_{\Box}(s_{1},s_{2},v):=\Box_{s_{1}}\Box_{s_{2}}v-\Box_{s_{2}}\Box_{s_{1}}v-\Box_{[s_{1},s_{2}]}v, as well as the Jacobiator (3.46), is a tensor. Moreover, if (E2,κ2)(E^{2},\kappa^{2}) is AL, then also the difference between LHS and RHS of the remaining conditions (Lie\mathrm{Lie_{\partial}}), (Lieβ\mathrm{Lie_{\beta}}) and (Lieω\mathrm{Lie_{\omega}}) is also tensorial. Hence, it is enough to verify all the conditions given in Theorem 3.20 on sections from local frames of the VBs AA and CC.

Remark 3.23.

The structure of a Lie HA (E2,κ2)(E^{2},\kappa^{2}) is fully determined by the Lie algebroid structure on AMA\to M, along with the maps \partial, \Box, and C\sharp^{C}, such that the following compatibility conditions hold: (AL\mathrm{AL_{\partial}}), (AL\mathrm{AL_{\Box}}), (Lie\mathrm{Lie_{\Box}}), and (Lie\mathrm{Lie_{\partial}}). Indeed, we define a skew HA on the graded bundle E2E^{2} described in Lemma 3.6 by setting β\beta via (Lieβ\mathrm{Lie_{\beta}}), so that ε1=0\varepsilon_{1}=0, and ψsym=0{\psi}^{\mathrm{sym}}=0, ωsym=0{\omega}^{\mathrm{sym}}=0, see Theorem 3.13. The resulting skew HA is AL (see Corollary 3.17), and Lie as (Lieω\mathrm{Lie_{\omega}}) follows from (Lieω¯\mathrm{Lie_{\bar{\omega}}}).

Proof.

Assume that (E2,κ2)(E^{2},\kappa^{2}) is a Lie HA, hence (A,κ)(A,\kappa) is a Lie algebroid, hence (LieA\mathrm{Lie_{A}}) holds. According to Theorem 2.11, an almost Lie HA (E2,κ2)(E^{2},\kappa^{2}) is Lie if and only if

[s1,s2]i+j=[s1i,s2j]{[s_{1},s_{2}]}^{\langle{i+j}\rangle}=[{s_{1}}^{\langle{i}\rangle},{s_{2}}^{\langle{j}\rangle}] (3.47)

for s1,s2Γ(A)s_{1},s_{2}\in\operatorname{\Gamma}(A), and (i,j)=(0,0),(1,0),(1,1)(i,j)=(0,0),(-1,0),(-1,-1) and (2,0)(-2,0). We shall show first that (3.47) implies the remaining conditions (Lie\mathrm{Lie_{\Box}}) – (Lieω\mathrm{Lie_{\omega}}).

The condition (Lie\mathrm{Lie_{\Box}}) can be rewritten in the form

[[s1,s2]0,v]=[s10,[s20,v]][s20,[s10,v]][{[s_{1},s_{2}]}^{\langle{0}\rangle},v]=[{s_{1}}^{\langle{0}\rangle},[{s_{2}}^{\langle{0}\rangle},v]]-[{s_{2}}^{\langle{0}\rangle},[{s_{1}}^{\langle{0}\rangle},v]]

and it follows from (3.47) with (i,j)=(0,0)(i,j)=(0,0) and the Jacobi identity for vector fields. The conditions (Lie\mathrm{Lie_{\partial}}) and (Lieβ\mathrm{Lie_{\beta}}) can be equivalently written as (3.47) with (i,j)=(1,1)(i,j)=(-1,-1) and (i,j)=(2,0)(i,j)=(-2,0). Indeed,

β(s1,s2)=12[s11,s21]=12[s1,s2]2=([s1,s2])=[s10,12s22]=s1(s2).\beta(s_{1},s_{2})=\frac{1}{2}[{s_{1}}^{\langle{-1}\rangle},{s_{2}}^{\langle{-1}\rangle}]=\frac{1}{2}{[s_{1},s_{2}]}^{\langle{-2}\rangle}=\partial([s_{1},s_{2}])=[{s_{1}}^{\langle{0}\rangle},\frac{1}{2}{s_{2}}^{\langle{-2}\rangle}]=\Box_{s_{1}}\partial(s_{2}).

Finally, (Lieω\mathrm{Lie_{\omega}}) reads as

[s11,[s21,s0]]=[s11,[s2,s]1][{s_{1}}^{\langle{-1}\rangle},[{s_{2}}^{\langle{-1}\rangle},{s}^{\langle{0}\rangle}]]=[{s_{1}}^{\langle{-1}\rangle},{[s_{2},s]}^{\langle{-1}\rangle}]

(see (3.25)), and this equality is true thanks to (3.47) with (i,j)=(1,0)(i,j)=(-1,0).

Conversely, assume that the conditions (LieA\mathrm{Lie_{A}})-(Lieω\mathrm{Lie_{\omega}}) hold. Then, for (3.47)(i,j)=(0,2)\eqref{e:kappa_2_bracket}_{(i,j)=(0,-2)}, we write

[s10,12s22]=s1(s2)=(Lie)([s1,s2])=12[s1,s2]2.[{s_{1}}^{\langle{0}\rangle},\frac{1}{2}{s_{2}}^{\langle{-2}\rangle}]=\Box_{s_{1}}\partial(s_{2})\stackrel{{\scriptstyle\eqref{i:Lie_ax:pa}}}{{=}}\partial([s_{1},s_{2}])=\frac{1}{2}{[s_{1},s_{2}]}^{\langle{-2}\rangle}.

Similarly, for (3.47)(i,j)=(1,1)\eqref{e:kappa_2_bracket}_{(i,j)=(-1,-1)}:

12[s1[1],s21]=β(s1,s2)=(Lieβ)([s1,s2])=12[s1,s2]2.\frac{1}{2}[{s_{1}}^{[-1]},{s_{2}}^{\langle{-1}\rangle}]=\beta(s_{1},s_{2})\stackrel{{\scriptstyle\eqref{i:Lie_ax:beta}}}{{=}}\partial([s_{1},s_{2}])=\frac{1}{2}{[s_{1},s_{2}]}^{\langle{-2}\rangle}.

The case (i,j)=(0,1)(i,j)=(0,-1) is more complicated. Denote Δ:=[s10,s21][s1,s2]1\Delta:=[{s_{1}}^{\langle{0}\rangle},{s_{2}}^{\langle{-1}\rangle}]-{[s_{1},s_{2}]}^{\langle{-1}\rangle}, so Δ𝔛1(E2)\Delta\in\mathfrak{X}_{-1}(E^{2}). We shall show first that Δ\Delta is annihilated by Tσ12\mathrm{T}\sigma^{2}_{1}. We have

(Tσ12)[s10,s21]E2=[(Tσ12)s10,(Tσ12)s21]E1=[s10κ1,s21κ1].(\mathrm{T}\sigma^{2}_{1})[{s_{1}}^{\langle{0}\rangle},{s_{2}}^{\langle{-1}\rangle}]_{E^{2}}=[(\mathrm{T}\sigma^{2}_{1}){s_{1}}^{\langle{0}\rangle},(\mathrm{T}\sigma^{2}_{1}){s_{2}}^{\langle{-1}\rangle}]_{E^{1}}=[s_{1}^{\langle 0\rangle_{\kappa^{1}}},s_{2}^{\langle-1\rangle_{\kappa^{1}}}].

(We have used the compatibility of algebroid lifts with respect to κ2\kappa^{2} and its order-one reduction κ1\kappa^{1}, as guaranteed by Lemma 2.14.) The latter is [s1,s2]1κ1[s_{1},s_{2}]^{\langle-1\rangle_{\kappa^{1}}} (as (A,κ1)(A,\kappa^{1}) is Lie) and this coincides with the projection of [s1,s2]1𝔛1(E2){[s_{1},s_{2}]}^{\langle{-1}\rangle}\in\mathfrak{X}_{-1}(E^{2}) onto E1E^{1}. Hence, the vector field Δ\Delta is vertical with respect to the projection σ12:E2E1\sigma^{2}_{1}:E^{2}\rightarrow E^{1}, as we claimed. Hence, Δ𝔛1𝐕(E2)Hom(A,C)\Delta\in\mathfrak{X}^{\mathbf{V}}_{-1}(E^{2})\simeq\operatorname{Hom}(A,C) by Lemma 4.1(v), i.e., Δ\Delta can be considered a VB morphism ACA\rightarrow C.

We know that ω=0\omega=0, hence From condition (Lieω\mathrm{Lie_{\omega}}) we find that for any section sΓ(A)s\in\operatorname{\Gamma}(A) we have [Δ,s1]=0[\Delta,{s}^{\langle{-1}\rangle}]=0. For X𝔛1𝐕(E2)Hom(A,C)X\in\mathfrak{X}^{\mathbf{V}}_{-1}(E^{2})\simeq\operatorname{Hom}(A,C) and s𝔛1(E2)Γ(C)s\in\mathfrak{X}_{-1}(E^{2})\simeq\operatorname{\Gamma}(C), the Lie bracket [X,s][X,s] of vector fields on E2E^{2} reads as

[X,s1]E2=XsΓ(C)=𝔛2(E2),[X,{s}^{\langle{-1}\rangle}]_{E^{2}}=-X\circ s\in\operatorname{\Gamma}(C)=\mathfrak{X}_{-2}(E^{2}),

see Lemma 4.1. We take X:=ΔX:=\Delta. Vanishing of Δs\Delta\circ s for any sΓ(A)s\in\operatorname{\Gamma}(A) implies Δ=0\Delta=0.

We follow a similar idea in the case (i,j)=(0,0)(i,j)=(0,0). Consider =[s1,s2]0[s1,s2]0\heartsuit={[s_{1},s_{2}]}^{\langle{0}\rangle}-{[s_{1},s_{2}]}^{\langle{0}\rangle}, so 𝔛2(E2)\heartsuit\in\mathfrak{X}_{-2}(E^{2}), and refer to the exact sequence (4.1) in Lemma 4.1. We aim to show that \heartsuit is in the kernel of the projection π:X(Tσ12,X|C)\pi:X\mapsto(\mathrm{T}\sigma^{2}_{1},X|_{C}). Indeed, the vector fields s10,s20{s_{1}}^{\langle{0}\rangle},{s_{2}}^{\langle{0}\rangle} are tangent to the submanifold CE2C\subset E^{2} (as they have weight 0 and CC is given in E2E^{2} by the equations yi=0y^{i}=0), so [s10,s20]|C=[s10|C,s20|C][{s_{1}}^{\langle{0}\rangle},{s_{2}}^{\langle{0}\rangle}]|_{C}=[{s_{1}}^{\langle{0}\rangle}|_{C},{s_{2}}^{\langle{0}\rangle}|_{C}]. Thus, |C=0\heartsuit|_{C}=0. Analogously to the case (i,j)=(1,0)(i,j)=(-1,0), we have (Tσ12)=0(\mathrm{T}\sigma^{2}_{1})\heartsuit=0, as (Tσ12)[s10,s20]E2=[s10κ1,s20κ1]E1=[s1,s2]0κ1(\mathrm{T}\sigma^{2}_{1})[{s_{1}}^{\langle{0}\rangle},{s_{2}}^{\langle{0}\rangle}]_{E^{2}}=[s_{1}^{\langle 0\rangle_{\kappa^{1}}},s_{2}^{\langle 0\rangle_{\kappa^{1}}}]_{E^{1}}=[s_{1},s_{2}]^{\langle 0\rangle_{\kappa^{1}}}. Hence we know, that Hom(Sym2A,E2widehat)𝔛0(E2)\heartsuit\in\operatorname{Hom}(\operatorname{Sym}^{2}A,\widehat{E^{2}})\subset\mathfrak{X}_{0}(E^{2}), i.e., it has a form

=12cijμ(x)yiyjzμ.\heartsuit=\frac{1}{2}c^{\mu}_{ij}(x)y^{i}y^{j}\partial_{z^{\mu}}.

for some functions cijμc^{\mu}_{ij} on MM. Next, we notice that for any section sΓ(A)s\in\operatorname{\Gamma}(A) we have

[,s1]=0.[\heartsuit,{s}^{\langle{-1}\rangle}]=0.

Indeed, [[s1,s2]0,s1]=(i,j)=(0,1)[[s1,s2],s]1[{[s_{1},s_{2}]}^{\langle{0}\rangle},{s}^{\langle{-1}\rangle}]\stackrel{{\scriptstyle(i,j)=(0,-1)}}{{=}}{[[s_{1},s_{2}],s]}^{\langle{-1}\rangle} and

[[s10,s20],s1]=[s10,[s20,s1]][s20,[s10,s1]]=(i,j)=(0,1)[s10,[s2,s]1][s20,[s1,s]1]=[s1,[s2,s]]1[s2,[s1,s]]1=[[s1,s2],s]1.\begin{split}[[{s_{1}}^{\langle{0}\rangle},{s_{2}}^{\langle{0}\rangle}],{s}^{\langle{-1}\rangle}]=[{s_{1}}^{\langle{0}\rangle},[{s_{2}}^{\langle{0}\rangle},{s}^{\langle{-1}\rangle}]]-[{s_{2}}^{\langle{0}\rangle},[{s_{1}}^{\langle{0}\rangle},{s}^{\langle{-1}\rangle}]]\stackrel{{\scriptstyle(i,j)=(0,-1)}}{{=}}\\ [{s_{1}}^{\langle{0}\rangle},{[s_{2},s]}^{\langle{-1}\rangle}]-[{s_{2}}^{\langle{0}\rangle},{[s_{1},s]}^{\langle{-1}\rangle}]={[s_{1},[s_{2},s]]}^{\langle{-1}\rangle}-{[s_{2},[s_{1},s]]}^{\langle{-1}\rangle}={[[s_{1},s_{2}],s]}^{\langle{-1}\rangle}.\end{split}

Therefore, [[,s11],s21]=0[[\heartsuit,{s_{1}}^{\langle{-1}\rangle}],{s_{2}}^{\langle{-1}\rangle}]=0 for any s1,s2Γ(A)s_{1},s_{2}\in\operatorname{\Gamma}(A). On the other hand, for any χHom(Sym2A,C)\chi\in\operatorname{Hom}(\operatorname{Sym}^{2}A,C), we have

[[χ,s11],s21]=χ(s1,s2)Γ(C),[[\chi,{s_{1}}^{\langle{-1}\rangle}],{s_{2}}^{\langle{-1}\rangle}]=\chi(s_{1},s_{2})\in\operatorname{\Gamma}(C),

up to isomorphisms given in Lemma 4.1. Therefore, =0\heartsuit=0. ∎

3.2.6 HAs of order two and representations up to homotopy of Lie algebroids

The notion of the representation up to homotopy of Lie algebroids was introduced in [AC12]. Some recollection on this subject is given in Appendix, Subsection 4.4. In our case of interest (2-term representations), the definition given in [AC12] boils down to the following data: a Lie algebroid (AM,[,],)(A\rightarrow M,[\cdot,\cdot],\sharp), a 2-term complex F0F1F_{0}\xrightarrow{\partial}F_{1} of vector bundles over MM concentrated in degrees 0 and 11, AA-connections i\nabla^{i} on FiF_{i}, for i=0,1i=0,1, and AA-form KΩ2(A;Hom(F1,F0))K\in\Omega^{2}(A;\operatorname{Hom}(F_{1},F_{0})) such that

  1. (i)

    s1=s0:Γ(F0)Γ(F1)\nabla_{s}^{1}\circ\partial={\partial}\circ\nabla_{s}^{0}:\operatorname{\Gamma}(F_{0})\rightarrow\operatorname{\Gamma}(F_{1}) for any sΓ(A)s\in\operatorname{\Gamma}(A);

  2. (ii)

    curv0=K\mathrm{curv}_{\nabla^{0}}=-K\circ\partial, and curv1=K\mathrm{curv}_{\nabla^{1}}=-\partial\circ K where curv\mathrm{curv}_{\nabla} denotes the curvature of an AA-connection \nabla, see (4.22);

  3. (iii)

    the covariant derivative of KK vanishes, i.e., dHomK=0\mathrm{d}_{\nabla^{\operatorname{Hom}}}K=0 where Hom\nabla^{\operatorname{Hom}} is the AA-connection on Hom(F1,F0)\operatorname{Hom}(F_{1},F_{0}) induced by 0\nabla^{0} and 1\nabla^{1}, see (4.25).

(Note that K=KK\circ\partial=K\bm{\wedge}\partial and K=K\partial\circ K=\partial\bm{\wedge}K, where \bm{\wedge} denotes an operation on AA-forms induced by the composition of maps, see (4.23).) All this data can be gathered together to a so called the structure operator D:Ω(A;F)Ω(A;F)D:\Omega(A;F)\to\Omega(A;F), determined by the triple (,=(0,1),K)(\partial,\nabla=(\nabla^{0},\nabla^{1}),K) (also denoted by DD) defined by means of the wedge product, as D:=widehat+d+KwidehatD:=\widehat{\partial}+d_{\nabla}+\widehat{K}, see (4.28). The compatibility conditions (i) - (iii) can be shortly written as DD=0D\circ D=0, see Appendix. A morphism (E0EE1;E,KE)(E_{0}\xrightarrow{\partial^{E}}E_{1};\nabla^{E},K^{E}) to (F0FF1;F,KF)(F_{0}\xrightarrow{\partial^{F}}F_{1};\nabla^{F},K^{F}) consists of a morphism of complexes Φ0:(E,E)(F,F)\Phi_{0}:(E,\partial^{E})\to(F,\partial^{F}) (i.e., Φ0E=FΦ0\Phi_{0}\circ\partial^{E}=\partial^{F}\circ\Phi_{0} ) and a 1-form Φ1Ω1(A;Hom(E1,F0))\Phi_{1}\in\Omega^{1}(A;\operatorname{Hom}(E_{1},F_{0})) such that

  1. (i)

    HomΦ1+dΦ0=0\partial^{\operatorname{Hom}}\Phi_{1}+\mathrm{d}_{\nabla}\Phi_{0}=0,

  2. (ii)

    dΦ1+KFΦ0Φ0KE=0\mathrm{d}_{\nabla}\Phi_{1}+K^{F}\bm{\wedge}\Phi_{0}-\Phi_{0}\bm{\wedge}K^{E}=0.

These conditions can be shortened to [Φwidehat,D]=0[\widehat{\Phi},D]=0 and can be rewritten in a more explicit form as

Φ1(s;e)Φ0(sE0e)+sF0Φ0(e)=0 for eΓ(E0);{-}\Phi_{1}(s;\partial e)-\Phi_{0}(\nabla_{s}^{E_{0}}e)+\nabla_{s}^{F_{0}}\Phi_{0}(e)=0\text{ for $e\in\operatorname{\Gamma}(E_{0})$;} (3.48)
(Φ1(s;v))Φ0(sE1v)+sF1Φ0(v)=0 for vΓ(E1);{-}\partial(\Phi_{1}(s;v))-\Phi_{0}(\nabla_{s}^{E_{1}}v)+\nabla_{s}^{F_{1}}\Phi_{0}(v)=0\text{ for $v\in\operatorname{\Gamma}(E_{1})$}; (3.49)
KF(s1,s2;Φ0(v))Φ0(KE(s1,s2;v))Φ1([s1,s2];v)Φ1(s2;s1E1v)+s1F0Φ1(s2;v)+Φ1(s1;s2E1v)s2F0Φ1(s1;v)=0, for s1,s2Γ(A),vΓ(E1).\begin{split}K^{F}(s_{1},s_{2};\Phi_{0}(v))-\Phi_{0}(K^{E}(s_{1},s_{2};v))-\Phi_{1}([s_{1},s_{2}];v){-}\Phi_{1}(s_{2};\nabla_{s_{1}}^{E_{1}}v)+\nabla_{s_{1}}^{F_{0}}\Phi_{1}(s_{2};v)\\ {+}\Phi_{1}(s_{1};\nabla_{s_{2}}^{E_{1}}v)-\nabla^{F_{0}}_{s_{2}}\Phi_{1}(s_{1};v)=0,\text{ for }s_{1},s_{2}\in\operatorname{\Gamma}(A),v\in\operatorname{\Gamma}(E_{1}).\end{split} (3.50)

The advantage of the framework of representations u.t.h. of Lie algebroids is that it is more flexible and contains generalizations of some important concepts from the theory of Lie algebras. The example is the adjoint representation. It is modelled on the complex

ATM,A\xrightarrow{\sharp}\mathrm{T}M,

and the AA-connections on this complex is induced by a linear connection :(X,s)Xs\nabla:(X,s)\mapsto\nabla_{X}s on the vector bundle AMA\to M in the following way161616Two choices of connections on AA leads to isomorphic representations.

s1As2\displaystyle\nabla^{A}_{s_{1}}s_{2} =s2s1+[s1,s2],\displaystyle=\nabla_{\sharp s_{2}}s_{1}+[s_{1},s_{2}], (3.51)
sTMX\displaystyle\nabla^{\mathrm{T}M}_{s}X =(Xs)+[s,X]τM.\displaystyle=\sharp(\nabla_{X}s)+[\sharp s,X]_{\tau_{M}}. (3.52)

The curvatures of the AA-connections A\nabla^{A} and TM\nabla^{\mathrm{T}M} are expressed in the terms of the following 22-form, called the basic curvature RbasΩ2(A;Hom(TM,A))R_{\nabla}^{\mathrm{bas}}\in\Omega^{2}(A;\operatorname{Hom}(\mathrm{T}M,A)), as curvA=Rbas\mathrm{curv}_{\nabla^{A}}=-R_{\nabla}^{\mathrm{bas}}\circ\sharp, curvTM=Rbas\mathrm{curv}_{\nabla^{\mathrm{T}M}}=-\sharp\circ R_{\nabla}^{\mathrm{bas}}, where

Rbas(s1,s2;X)=X[s1,s2][Xs1,s2][s1,Xs2]s2TMXs1+s1TMXs2,R_{\nabla}^{\mathrm{bas}}(s_{1},s_{2};X)=\nabla_{X}[s_{1},s_{2}]-[\nabla_{X}s_{1},s_{2}]-[s_{1},\nabla_{X}s_{2}]-\nabla_{\nabla_{s_{2}}^{\mathrm{T}M}X}s_{1}+\nabla_{\nabla_{s_{1}}^{\mathrm{T}M}X}s_{2}, (3.53)

see [AC12]. The structure operator for the adjoint representation of a Lie algebroid is denoted by ad=(,(A,TM),Rbas)\operatorname{ad}_{\nabla}=(\sharp,(\nabla^{A},\nabla^{\mathrm{T}M}),R^{\mathrm{bas}}).

From order-two Lie HA to 2-term representations.

Let (E2,κ2)(E^{2},\kappa^{2}) be a Lie HA of order two. Recall that it is determined by the Lie algebroid structure on the vector bundle AMA\rightarrow M (being the order-one reduction of E2E^{2}), and the structure maps \partial, \Box, C\sharp^{C}, see Remark 3.23. We shall define a Lie algebroid representation u.t.h. on the complex ACA\xrightarrow{\partial}C, AA in degree 0, CC – degree 1. Our construction mimics the adjoint representation of a Lie algebroid.

Definition 3.24.

Let us choose a linear connection \nabla on the vector bundle σ:AM\sigma:A\rightarrow M and define:

  • an AA-connection C\nabla^{C} on CC:

    sCv:=sv+C(v)s,\nabla^{C}_{s}v:=\Box_{s}v+\partial\circ\nabla_{\sharp^{C}(v)}s,

    where sΓ(A)s\in\operatorname{\Gamma}(A), vΓ(C)v\in\operatorname{\Gamma}(C);

  • an AA-connection A\nabla^{A} on AA:

    s1As2:=s2s1+[s1,s2];\nabla^{A}_{s_{1}}{s_{2}}:=\nabla_{\sharp s_{2}}s_{1}+[s_{1},s_{2}];
  • an two-form KΩ2(A,Hom(C,A))K\in\Omega^{2}(A,\operatorname{Hom}(C,A)):

    K(s1,s2;v):=Cv[s1,s2][Cvs1,s2][s1,Cvs2]C(s2Cv)s1+C(s1Cv)s2.K(s_{1},s_{2};v):=\nabla_{\sharp^{C}v}[s_{1},s_{2}]-[\nabla_{\sharp^{C}v}s_{1},s_{2}]-[s_{1},\nabla_{\sharp^{C}v}s_{2}]-\nabla_{\sharp^{C}(\nabla_{s_{2}}^{C}v)}s_{1}+\nabla_{\sharp^{C}(\nabla_{s_{1}}^{C}v)}s_{2}.

We assume that in both constructions, the adjoint representation and the representation on the complex :AC\partial:A\rightarrow C, we have chosen the same linear connection on AA. Then, the AA-connections A\nabla^{A} defined above and in the adjoint representation, also coincide. Moreover,

K=RbasC,K=R^{\mathrm{bas}}_{\nabla}\circ\sharp^{C}, (3.54)

where RbasR^{\mathrm{bas}}_{\nabla} is given in (3.53). Indeed, by comparing the formulas for KK and RbasR^{\mathrm{bas}}_{\nabla}, for (3.54) we need to show that

CsCv=sTM(Cv)\sharp^{C}\circ\nabla^{C}_{s}v=\nabla_{s}^{\mathrm{T}M}(\sharp^{C}v) (3.55)

This can be rewritten as

C(sv)+CCvs=(Cvs)+[s,Cv]\sharp^{C}(\Box_{s}v)+\sharp^{C}\circ\partial\,\nabla_{\sharp^{C}v}s=\sharp\left(\nabla_{\sharp^{C}v}s\right)+[\sharp s,\sharp^{C}v]

and it is true due to the AL assumption (see (AL\mathrm{AL_{\Box}}), (AL\mathrm{AL_{\partial}}) in Theorem 3.16).

Lemma 3.25.

An order-two Lie higher algebroid (E2,κ2)(E^{2},\kappa^{2}) gives rise, as explained in Definition 3.24, to a representation u.t.h. of the Lie algebroid AA (the order-one reduction of (E2,κ2)(E^{2},\kappa^{2})) on the complex

ACA\xrightarrow{\partial}C (3.56)

with the structure operator given by D=(,(A,C),K)D=(\partial,(\nabla^{A},\nabla^{C}),K). Two choices of the connection on the vector bundle σ:AM\sigma:A\rightarrow M result in isomorphic representations. Moreover, idAC\operatorname{id}_{A}\oplus\sharp^{C} gives rise to a morphism from the constructed representation (A[0]C[1],D)(A_{[0]}\oplus C_{[1]},D) to the adjoint representation (A[0](TM)[1],ad)(A_{[0]}\oplus(\mathrm{T}M)_{[1]},\operatorname{ad}_{\nabla}) of AA.

Proof.

It is straightforward to check that C\nabla^{C} is an AA-connection. Indeed, using tensor-like properties of \Box described in Theorem 3.13, we get

fsCv=f(sv)(Cv)(f)s+fCvs+((Cv)(f)s)=fsCv.\nabla^{C}_{fs}v={f(\Box_{s}v)}-(\sharp^{C}v)(f)\,\partial s+f\,\partial\nabla_{\sharp^{C}v}s+\partial\left((\sharp^{C}v)(f)\,s\right)=f\nabla_{s}^{C}v.

Similarly, we check that sC:Γ(C)Γ(C)\nabla^{C}_{s}:\operatorname{\Gamma}(C)\rightarrow\operatorname{\Gamma}(C) is a derivative endomorphism,

sC(fv)fsCv=fsvfsv+(fcvsfCvs)=(s)(f)v.\nabla^{C}_{s}(fv)-f\nabla^{C}_{s}v=\Box_{fs}v-f\Box_{s}v+\partial\circ\left(\nabla_{f\sharp^{c}v}s-f\nabla_{\sharp^{C}v}s\right)=(\sharp s)(f)v.

The AA-connections A\nabla^{A} and C\nabla^{C} are compatible with :AC\partial:A\to C. Indeed,

s1As2=([s1,s2]+s2s1)=s1(s2)+C(s2)s1=s1Cs2.\partial\nabla^{A}_{s_{1}}s_{2}=\partial\left([s_{1},s_{2}]+\nabla_{\sharp s_{2}}s_{1}\right)=\Box_{s_{1}}(\partial s_{2})+\partial\nabla_{\sharp^{C}(\partial s_{2})}s_{1}=\nabla^{C}_{s_{1}}{\partial s_{2}}.

Here, we used (AL\mathrm{AL_{\partial}}) and (Lie\mathrm{Lie_{\partial}}) which are true in any Lie HA. In analogy to [AC12, Proposition 2.11] we shall prove that

  1. (i)

    curvA=K\mathrm{curv}_{\nabla^{A}}=-K\circ\partial and curvC=K\mathrm{curv}_{\nabla^{C}}=-\partial\circ K;

  2. (ii)

    dHomK=0\mathrm{d}_{\nabla^{\operatorname{Hom}}}K=0, i.e., KK is closed with respect to the AA-connection Hom\nabla^{\operatorname{Hom}} on Hom(C,A)\operatorname{Hom}(C,A) induced by A\nabla^{A} and C\nabla^{C}.

Recall that the curvature of A\nabla^{A} is Rbas-R^{\mathrm{bas}}\circ\sharp, hence

curvA(s1,s2;s)=Rbas(s1,s2;s)=Rbas(s1,s2;C(s)),\mathrm{curv}_{\nabla^{A}}(s_{1},s_{2};s)=-R^{\mathrm{bas}}_{\nabla}(s_{1},s_{2};\sharp s)=-R^{\mathrm{bas}}_{\nabla}(s_{1},s_{2};\sharp^{C}(\partial s)),

i.e., curvA=(RbasC)=K\mathrm{curv}_{\nabla^{A}}=-(R_{\nabla}^{\mathrm{bas}}\circ\sharp^{C})\circ\partial=-K\circ\partial due to (3.54). For the curvature of C\nabla^{C} we apply \partial to (3.53) and get

Rbas(s1,s2;Cv)\displaystyle\partial\circ R^{\mathrm{bas}}_{\nabla}(s_{1},s_{2};\sharp^{C}v) =Cv[s1,s2]I[Cvs1,s2]II[s1,Cvs2]IIs2TMCvs1III+s1TMCvs2III\displaystyle=\underbrace{\partial\nabla_{\sharp^{C}v}[s_{1},s_{2}]}_{\mathrm{I}}-\underbrace{\partial[\nabla_{\sharp^{C}v}s_{1},s_{2}]}_{\mathrm{II}}-\underbrace{\partial[s_{1},\nabla_{\sharp^{C}v}s_{2}]}_{\mathrm{II}^{\prime}}-\underbrace{\partial\nabla_{\nabla_{s_{2}}^{\mathrm{T}M}\sharp^{C}v}s_{1}}_{\mathrm{III}}+\underbrace{\partial\nabla_{\nabla_{s_{1}}^{\mathrm{T}M}\sharp^{C}v}s_{2}}_{\mathrm{III}^{\prime}}

On the other hand,

[s1,s2]Cv\displaystyle\nabla^{C}_{[s_{1},s_{2}]}v =[s1,s2]vJ+Cv[s1,s2]I,\displaystyle=\underbrace{\Box_{[s_{1},s_{2}]}v}_{\mathrm{J}}+\underbrace{\partial\nabla_{\sharp^{C}v}[s_{1},s_{2}]}_{\mathrm{I}},
s1Cs2Cv\displaystyle\nabla_{s_{1}}^{C}\nabla_{s_{2}}^{C}v =s1C(s2v+Cvs2)=\displaystyle=\nabla_{s_{1}}^{C}\left(\Box_{s_{2}}v+\partial\nabla_{\sharp^{C}v}s_{2}\right)=
=s1s2vJ+C(s2v)s1III1+s1(Cvs2)II+C(Cvs2)s1III2\displaystyle=\underbrace{\Box_{s_{1}}\Box_{s_{2}}v}_{\mathrm{J}^{\prime}}+\underbrace{\partial\nabla_{\sharp^{C}(\Box_{s_{2}}v)}s_{1}}_{\mathrm{III}_{1}}+\underbrace{\Box_{s_{1}}\left(\partial\nabla_{\sharp^{C}v}s_{2}\right)}_{\mathrm{II}}+\underbrace{\partial\nabla_{\sharp^{C}(\partial\nabla_{\sharp^{C}v}s_{2})}\,s_{1}}_{\mathrm{III}_{2}}

We have analogous expressions (J′′)(\mathrm{J}^{\prime\prime}), (III1)(\mathrm{III}_{1}^{\prime}), (II)(\mathrm{II}^{\prime}) and (III2)(\mathrm{III}_{2}^{\prime}) for s2Cs1Cv\nabla_{s_{2}}^{C}\nabla_{s_{1}}^{C}v. We should show that

Rbas(s1,s2;Cv)+s1Cs2Cvs2Cs1Cv[s1,s2]Cv=0.\partial\circ R^{\mathrm{bas}}_{\nabla}(s_{1},s_{2};\sharp^{C}v)+\nabla_{s_{1}}^{C}\nabla_{s_{2}}^{C}v-\nabla_{s_{2}}^{C}\nabla_{s_{1}}^{C}v-\nabla^{C}_{[s_{1},s_{2}]}v=0.

We see that the expressions (J)(\mathrm{J}), (J)(\mathrm{J}^{\prime}) and (J′′)(\mathrm{J}^{\prime\prime}) cancel, due to (Lie\mathrm{Lie_{\Box}}). Similarly for the two expressions denoted by (I)(\mathrm{I}). Next, the expressions (II)(\mathrm{II}) cancel due to (Lie\mathrm{Lie_{\partial}}), and similarly for (II)(\mathrm{II}^{\prime}). Finally, for (III)(\mathrm{III}) we have

s2TMCvs1=Cvs2s1III2+[s2,Cv]s1III1,\nabla_{\nabla_{s_{2}}^{\mathrm{T}M}\sharp^{C}v}\,s_{1}=\underbrace{\nabla_{\sharp\nabla_{\sharp^{C}v}s_{2}}\,s_{1}}_{\mathrm{III}_{2}}+\underbrace{\nabla_{[\sharp s_{2},\sharp^{C}v]}s_{1}}_{\mathrm{III}_{1}},

hence (III)(\mathrm{III}) equals (III1)+(III2)(\mathrm{III}_{1})+(\mathrm{III}_{2}). Similarly, (III)(\mathrm{III}^{\prime}) cancels with the sum of (III1)(\mathrm{III}_{1}^{\prime}) and (III2)(\mathrm{III}_{2}^{\prime}). Here we used Theorem 3.16: (AL\mathrm{AL_{\Box}}) and (AL\mathrm{AL_{\partial}}).

We shall prove the second claim that the 2-form KK is closed. We shall use (3.54) and the equality dbasRbas=0\mathrm{d}_{\nabla^{\mathrm{bas}}}R^{\mathrm{bas}}=0 which is proved in [AC12]. We have

dHomK(s1,s2,s3;v)\displaystyle\mathrm{d}_{\nabla^{\operatorname{Hom}}}K(s_{1},s_{2},s_{3};v) =cyclic(s1HomK(s2,s3))(v)K([s1,s2],s3;v)=\displaystyle=\sum_{\text{cyclic}}\left(\nabla^{\operatorname{Hom}}_{s_{1}}K(s_{2},s_{3})\right)(v)-K([s_{1},s_{2}],s_{3};v)=
=cyclics1A(K(s2,s3;v))Rbas(s2,s3;Cv)K(s2,s3)(s1Cv)Rbas(s2,s3)(Cs1Cv)Rbas([s1,s2],s3;Cv)=\displaystyle=\sum_{\text{cyclic}}\underbrace{\nabla_{s_{1}}^{A}\left(K(s_{2},s_{3};v)\right)}_{R^{\mathrm{bas}}_{\nabla}(s_{2},s_{3};\sharp^{C}v)}-\underbrace{K(s_{2},s_{3})(\nabla^{C}_{s_{1}}v)}_{R^{\mathrm{bas}}_{\nabla}(s_{2},s_{3})(\sharp^{C}\nabla^{C}_{s_{1}}v)}-R^{\mathrm{bas}}_{\nabla}([s_{1},s_{2}],s_{3};\sharp^{C}v)=
=dbasRbas(s1,s2,s3;Cv)=0,\displaystyle=\mathrm{d}_{\nabla^{\mathrm{bas}}}R^{\mathrm{bas}}_{\nabla}(s_{1},s_{2},s_{3};\sharp^{C}v)=0,

by (3.55). The proof that D=(,,K)D=(\partial,\nabla,K) is a structure operator is completed.

Let us assume that we have chosen two linear connections \nabla and ~\widetilde{\nabla} on the vector bundle σ:AM\sigma:A\rightarrow M. We define Φ0=idAC\Phi_{0}=\operatorname{id}_{A\oplus C} and Φ1(s)(v)=Cvs~Cvs\Phi_{1}(s)(v)=\nabla_{\sharp^{C}v}s-\widetilde{\nabla}_{\sharp^{C}v}s, Φ1Ω1(A,Hom(C,A))\Phi_{1}\in\Omega^{1}(A,\operatorname{Hom}(C,A)). Note that Hom(C,A)=End1(A[0]C[1])\operatorname{Hom}(C,A)=\operatorname{End}^{-1}(A_{[0]}\oplus C_{[1]}). Then Φ0+Φ1\Phi_{0}+\Phi_{1} establishes an isomorphism between the representations u.t.h. of the Lie algebroid AA, induced from a given HA (E2,κ2)(E^{2},\kappa^{2}), defined by means of the linear connections \nabla and ~\widetilde{\nabla}, respectively. Indeed, the equation (3.48) writes as

Φ1(s1;s2)=s1As2~s1As2.\Phi_{1}(s_{1};\partial s_{2})=\nabla^{A}_{s_{1}}s_{2}-\widetilde{\nabla}^{A}_{s_{1}}s_{2}.

The RHS is s2s1~s2s1\nabla_{\sharp s_{2}}s_{1}-\widetilde{\nabla}_{\sharp s_{2}}s_{1} and the same is LHS as C=\sharp^{C}\circ\partial=\sharp. The second equation (3.49) writes as

(Φ1(s;v))=sCv~sCv\partial(\Phi_{1}(s;v))=\nabla^{C}_{s}v-\widetilde{\nabla}^{C}_{s}v

and both sides are equal to (Cvs~CvCs)\partial\circ\left(\nabla_{\sharp^{C}v}s-\widetilde{\nabla}^{C}_{\sharp^{C}v}s\right) due to the definitions of Φ1\Phi_{1} and the AA-connection on CC (see Definition 3.24). The third equation (3.50) is a consequence of a similar result for the adjoint representation. Namely, if Ψ0=idATM\Psi_{0}=\operatorname{id}_{A\oplus\mathrm{T}M}, Ψ1(s;X)=Xs~Xs\Psi_{1}(s;X)=\nabla_{X}s-\widetilde{\nabla}_{X}s, Ψ1Ω1(A,End¯1(ATM))\Psi_{1}\in\Omega^{1}(A,\underline{\operatorname{End}}^{-1}(A\oplus\mathrm{T}M)), then Ψ0+Ψ1\Psi_{0}+\Psi_{1} is an isomorphism (ATM,ad)(ATM,ad~)(A\oplus\mathrm{T}M,\operatorname{ad}_{\nabla})\rightarrow(A\oplus\mathrm{T}M,\operatorname{ad}_{\widetilde{\nabla}}) between the adjoint representations of the Lie algebroid AA associated with the linear connections \nabla and ~\widetilde{\nabla}, respectively. We have

Φ1(s;v)=Ψ1(s;Cv),\Phi_{1}(s;v)=\Psi_{1}(s;\sharp^{C}v),

hence, from (3.54) and (3.55), we find that

Φ1(s2;s1Cv)=Ψ1(s2;Cs1Cv)=Ψ1(s2;s1TMCv).\Phi_{1}(s_{2};\nabla^{C}_{s_{1}}v)=\Psi_{1}(s_{2};\sharp^{C}\nabla^{C}_{s_{1}}v)=\Psi_{1}(s_{2};\nabla_{s_{1}}^{\mathrm{T}M}\sharp^{C}v).

Hence, (3.50) can be written in our case as:

K~(s1,s2;v)K(s1,s2;v)=Φ1([s1,s2];v)Ψ1([s1,s2];Cv)+Φ1(s2;s1Cv)Φ1(s1;s2Cv)+\displaystyle K_{\widetilde{\nabla}}(s_{1},s_{2};v)-K_{\nabla}(s_{1},s_{2};v)=\underbrace{\Phi_{1}([s_{1},s_{2}];v)}_{\Psi_{1}([s_{1},s_{2}];\sharp^{C}v)}+\Phi_{1}(s_{2};\nabla^{C}_{s_{1}}v){-}\Phi_{1}(s_{1};\nabla^{C}_{s_{2}}v)+
~s2A(Φ1(s1;v))~s1A(Φ1(s2;v)),\displaystyle\widetilde{\nabla}^{A}_{s_{2}}(\Phi_{1}(s_{1};v))-\widetilde{\nabla}^{A}_{s_{1}}(\Phi_{1}(s_{2};v)),

and it follows from the same equation (3.50) applied to the adjoint representation, i.e., with KK, vv, and Φ\Phi replaced with RbasR^{\mathrm{bas}}, Cv\sharp^{C}v, and Ψ\Psi, respectively.

For the last statement, we clearly see that Φ0=(idA,C):ACATM\Phi_{0}=(\operatorname{id}_{A},\sharp^{C}):A\oplus C\to A\oplus\mathrm{T}M is a morphism of complexes, due to =C\sharp=\sharp^{C}\circ\partial (see Theorem 3.16). We set Φ1=0\Phi_{1}=0 and find that equation (3.48) holds automatically, equation (3.49) is true due to (3.55), and (3.50) reduces to (3.54). ∎

Recovering HA.

Assume we are given a Lie algebroid (AM,[,],)(A\to M,[\cdot,\cdot],\sharp), the structure operator D=(,(A,C),K)D=(\partial,(\nabla^{A},\nabla^{C}),K), which provides a representation u.t.h. of AA on the complex :AC\partial:A\to C. Let Φ\Phi be a morphism to the adjoint representation (A,ad)(A,\operatorname{ad}_{\nabla}), where \nabla is a chosen linear connection on AA. We assume that Φ0|A=idA\Phi_{0}|_{A}=\operatorname{id}_{A} and Φ1=0\Phi_{1}=0, where Φ0\Phi_{0}, Φ1\Phi_{1} are the components of Φ\Phi:

A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}idA\scriptstyle{\operatorname{id}_{A}}\scriptstyle{\partial}C\textstyle{C\ignorespaces\ignorespaces\ignorespaces\ignorespaces}C:=Φ0|C\scriptstyle{\sharp^{C}:=\Phi_{0}|_{C}}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\sharp}TM\textstyle{\mathrm{T}M}

We shall show how to recover the structure of a Lie HA on the graded bundle E2E^{2} constructed in Lemma 3.6 by means of the VB morphism :AC\partial:A\to C given already. The structure map C\sharp^{C} of the HA are taken from the diagram above, as C=Φ0|C\sharp^{C}=\Phi_{0}|_{C}. The structure map :Γ(A)×Γ(C)Γ(C)\Box:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(C)\rightarrow\operatorname{\Gamma}(C) is recovered by means of the formula given in Definition 3.24,

sv=sCvC(v)s.\Box_{s}v=\nabla^{C}_{s}v-\partial\circ\nabla_{\sharp^{C}(v)}s.

We easily check the tensor-like properties of the action (s,v)sv(s,v)\mapsto\Box_{s}v:

  • fsvf(sv)=(Cv(fs)fCvs)=(Cv)(f)(s)\Box_{fs}v-f(\Box_{s}v)=-\partial\circ(\nabla_{\sharp^{C}v}(fs)-f\nabla_{\sharp^{C}v}s)=-(\sharp^{C}v)(f)\,\partial(s), as sCv\nabla_{s}^{C}v is 𝒞(M)\mathcal{C}^{\infty}(M)-linear in ss.

  • s(fv)f(sv)=sC(fv)fsCv=(s)(f)v\Box_{s}(fv)-f\,(\Box_{s}v)=\nabla^{C}_{s}(fv)-f\nabla_{s}^{C}v=(\sharp s)(f)\,v due to the properties of AA-connections.

We shall show that the compatibility conditions given in Theorem 3.20, which ensure Lie HA structure are satisfied.

Obviously, (AL\mathrm{AL_{\partial}}) is true due to the commutativity of the diagram above. Since Φ1=0\Phi_{1}=0 the conditions (3.48), (3.49), (3.50) simplify to:

  1. (i)

    The AA-connections on the vector bundle σ:AM\sigma:A\to M, being part of the structure operators DD and ad\operatorname{ad}_{\nabla}, coincide;

  2. (ii)

    C(sCv)=sTM(Cv)\sharp^{C}\left(\nabla^{C}_{s}v\right)=\nabla^{\mathrm{T}M}_{s}\left(\sharp^{C}v\right);

  3. (iii)

    Rbas(s1,s2;Cv)=K(s1,s2;v)R^{\mathrm{bas}}_{\nabla}(s_{1},s_{2};\sharp^{C}v)=K(s_{1},s_{2};v).

We have

C(sv)=C(sCvCvs)=sTM(Cv)(C)Cvs=[s,Cv]\sharp^{C}(\Box_{s}v)=\sharp^{C}\left(\nabla^{C}_{s}v-\partial\circ\nabla_{\sharp^{C}v}s\right)=\nabla^{\mathrm{T}M}_{s}\left(\sharp^{C}v\right)-(\sharp^{C}\circ\partial)\nabla_{\sharp^{C}v}s=[\sharp s,\sharp^{C}v]

due to the formula (3.52) for TM\nabla^{\mathrm{T}M}. It proves (AL\mathrm{AL_{\Box}}). Next,

s1(s2)=s1C(s2)(C)s2s1=s1C(s2)(s1As2[s1,s2])=[s1,s2]+(s1Cs2s1As2)=[s1,s2],\begin{split}\Box_{s_{1}}\partial(s_{2})=\nabla^{C}_{s_{1}}(\partial s_{2})-\partial\nabla_{(\sharp^{C}\circ\partial)s_{2}}s_{1}=\\ \nabla^{C}_{s_{1}}(\partial s_{2})-\partial\left(\nabla^{A}_{s_{1}}s_{2}-[s_{1},s_{2}]\right)=\partial[s_{1},s_{2}]+\left(\nabla_{s_{1}}^{C}\partial s_{2}-\partial\nabla^{A}_{s_{1}}s_{2}\right)=\partial[s_{1},s_{2}],\end{split}

by the compatibility C\nabla^{C} with A\nabla^{A}, so (Lie\mathrm{Lie_{\partial}}) is true. It remains to prove that \Box satisfies (Lie\mathrm{Lie_{\Box}}), see Remark 3.23.

[s1,s2]v\displaystyle\Box_{[s_{1},s_{2}]}v =[s1,s2]CvCv[s1,s2],\displaystyle=\nabla^{C}_{[s_{1},s_{2}]}v-\partial\circ\nabla_{\sharp^{C}v}[s_{1},s_{2}],
s1s2v\displaystyle\Box_{s_{1}}\Box_{s_{2}}v =s1(s2CvCvs2)=s1C(s2CvCvs2)Cs2vs1,\displaystyle=\Box_{s_{1}}\left(\nabla^{C}_{s_{2}}v-\partial\circ\nabla_{\sharp^{C}v}s_{2}\right)=\nabla^{C}_{s_{1}}\left(\nabla^{C}_{s_{2}}v-\partial\circ\nabla_{\sharp^{C}v}s_{2}\right)-\partial\circ\nabla_{\sharp^{C}\Box_{s_{2}}v}s_{1},

hence, using (Lie\mathrm{Lie_{\partial}}), we get

curv(s2,s2;v):=[s1,s2]vs1s2v+s2s1v=curvC(s1,s2;v)(Cv[s1,s2][s2,Cv]s1+[s1,Cv]s2)+s1C(Cvs2)s2C(Cvs1).\begin{split}-\mathrm{curv}_{\Box}(s_{2},s_{2};v):=\Box_{[s_{1},s_{2}]}v-\Box_{s_{1}}\Box_{s_{2}}v+\Box_{s_{2}}\Box_{s_{1}}v=-\mathrm{curv}_{\nabla^{C}}(s_{1},s_{2};v)\\ -\partial\circ\left(\nabla_{\sharp^{C}v}[s_{1},s_{2}]-\nabla_{[\sharp s_{2},\sharp^{C}v]}s_{1}+\nabla_{[\sharp s_{1},\sharp^{C}v]}s_{2}\right){+}\nabla^{C}_{s_{1}}(\partial\circ\nabla_{\sharp^{C}v}s_{2}){-}\nabla^{C}_{s_{2}}(\partial\circ\nabla_{\sharp^{C}v}s_{1}).\end{split}

We replace curvC(s1,s2;v)-\mathrm{curv}_{\nabla^{C}}(s_{1},s_{2};v) with K(s1,s2;v)=Rbas(s1,s2;Cv)\partial\circ K(s_{1},s_{2};v)=\partial\circ R^{\mathrm{bas}}_{\nabla}(s_{1},s_{2};\sharp^{C}v), see (iii), and C\nabla^{C}\circ\partial with A\partial\circ\nabla^{A}, and find that curv(s2,s2;v)=(s1,s2;Cv)-\mathrm{curv}_{\Box}(s_{2},s_{2};v)=\partial\circ\triangle(s_{1},s_{2};\sharp^{C}v), where

(s1,s2;X)=Rbas(s1,s2;X)(X[s1,s2][s2,X]s1+[s1,X]s2)+(s1AXs2s2AXs1).\triangle(s_{1},s_{2};X)=R^{\mathrm{bas}}_{\nabla}(s_{1},s_{2};X)-\left(\nabla_{X}[s_{1},s_{2}]-\nabla_{[\sharp s_{2},X]}s_{1}+\nabla_{[\sharp s_{1},X]}s_{2}\right)+\left(\nabla_{s_{1}}^{A}\nabla_{X}s_{2}-\nabla^{A}_{s_{2}}\nabla_{X}s_{1}\right).

We expand the last bracket using the formula (3.51) for A\nabla^{A} and replace Rbas(s1,s2)(X)R^{\mathrm{bas}}_{\nabla}(s_{1},s_{2})(X) with (3.53), and after cancelling similar terms we get

(s1,s2)(X)=s2TMXs1+s1TMXs2+([s2,X]s1[s1,X]s2)+(Xs2s1Xs1s2)=0\begin{split}\triangle(s_{1},s_{2})(X)=-\nabla_{\nabla_{s_{2}}^{\mathrm{T}M}X}s_{1}+\nabla_{\nabla_{s_{1}}^{\mathrm{T}M}X}s_{2}+\left(\nabla_{[\sharp s_{2},X]}s_{1}-\nabla_{[\sharp s_{1},X]}s_{2}\right)+\\ \left(\nabla_{\sharp\nabla_{X}s_{2}}s_{1}-\nabla_{\sharp\nabla_{X}s_{1}}s_{2}\right)=0\end{split}

thanks to the formula for TM\nabla^{\mathrm{T}M} given in (3.52). The proof of (Lie\mathrm{Lie_{\Box}}) is completed. We have obtained the following result.

Theorem 3.26.

Let (AM,[,],)(A\to M,[\cdot,\cdot],\sharp) be a Lie algebroid and let us fix a vector bundle CMC\to M, a linear connection \nabla on AA and a VB morphism :AC\partial:A\to C over idM\operatorname{id}_{M}.

Assume, in addition, that we are given a representation u.t.h. of the Lie algebroid AA on the complex :AC\partial:A\rightarrow C, and a morphism Φ=(Φ0,Φ1)\Phi=(\Phi_{0},\Phi_{1}) from this to the adjoint representation (A,ad)(A,\operatorname{ad}_{\nabla}) such that Φ1=0\Phi_{1}=0 and Φ0|A=idA\Phi_{0}|_{A}=\operatorname{id}_{A}. Here, ΦiΩi(A;Endi(A[0]C[1]))\Phi_{i}\in\Omega^{i}(A;\operatorname{End}^{i}(A_{[0]}\oplus C_{[1]})), i=0,1i=0,1, are the components of Φ\Phi.

Then, there exists a unique HA structure on the graded bundle E2E^{2} constructed in Lemma 3.6, such that the representation u.t.h. of the Lie algebroid AA, and the morphism Φ\Phi, described in Lemma 3.25, are the given ones. This establishes a one-to-one correspondence between order-two Lie HA structures on the graded bundle E2E^{2} and morphisms Φ\Phi of the above form.

3.2.7 HAs, VB-alegbroids and representations up to homotopy

A brief account of VB-algebroids is given in Preliminaries.

Recall that the constructions from Definition 3.24 and the adjoint representation depend on the choice of a linear connection on the vector bundle AMA\to M. However, there is a way to avoid this choice. The motivation comes from description of 2-term representations in a framework of VB-algebroids, as discovered in [GSM10]. In this framework, the adjoint representation of AA is the VB-algebroid (TA;TM,A;M)(\mathrm{T}A;\mathrm{T}M,A;M) – the tangent prolongation of the algebroid AA.

Corollary 3.27.

A Lie HA (E2,κ2)(E^{2},\kappa^{2}) can be described by means of the following data:

  1. (i)

    a VB-algebroid structure on a DVB DD whose side bundles are CC and AA and the core is also AA;

  2. (ii)

    a VB-algebroid morphism Ψ\Psi from DD to TA\mathrm{T}A (the adjoint representation of AA) such that Ψ\Psi is the identity on the side bundle AA and also on the core bundle AA:

    D\textstyle{D\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}C\textstyle{C\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M\textstyle{M}   Ψ\scriptstyle{\Psi}   TA\textstyle{\mathrm{T}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τA\scriptstyle{\tau_{A}}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}TM\textstyle{\mathrm{T}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M\textstyle{M} (3.57)

We proceed with the proof by recalling the correspondence between 2-term representations and VB-algebroids. Details are nicely presented in [GSJLMM18].

Let (D;σE,σA;M)(D;\sigma_{E},\sigma_{A};M) be a DVB with the core CC, as in (2.2). As shown in [GSM10], a VB-algebroid structure on the DVB (DE;AM)(D\rightarrow E;A\rightarrow M), as in (2.2), together with a horizontal lift θA:Γ(A)ΓE(D)\theta_{A}:\operatorname{\Gamma}(A)\rightarrow\operatorname{\Gamma}^{\ell}_{E}(D), i.e., a splitting of the short exact sequence (2.3), gives rise to a representation u.t.h. of the Lie algebroid AA. (Recall that such horizontal lifts are in bijective correspondence with decompositions DE×MA×MCD\rightarrow E\times_{M}A\times_{M}C of the DVB DD, and with the inclusions σ:E×MAD\sigma:E\times_{M}A\rightarrow D.) We shall review this construction. First of all, it is a representation on the 2-term complex :CE\partial:C\to E, where \partial is the core of the anchor map D:DTE\sharp^{D}:D\to\mathrm{T}E. (Note that Dwidehat=C\widehat{D}=C and TEwidehat=E\widehat{\mathrm{T}E}=E.) The AA-connections on CC and EE, denoted by core\nabla^{\mathrm{core}} and aside\nabla_{a}^{\mathrm{side}} are the following (see [GSJLMM18]):

(acorec)=[θA(a),c]D,aside~=D(θA(a)),\left(\nabla^{\mathrm{core}}_{a}c\right)^{\dagger}=[\theta_{A}(a),c^{\dagger}]_{D},\quad\widetilde{\nabla_{a}^{\mathrm{side}}}=\sharp_{D}(\theta_{A}(a)), (3.58)

where aΓ(A)a\in\operatorname{\Gamma}(A), cΓ(C)c\in\operatorname{\Gamma}(C) and ξξ~\xi\mapsto\widetilde{\xi} denotes 1-1 correspondence between derivative endomorphism of Γ(E)\operatorname{\Gamma}(E) and linear vector fields on EE. The last component, AA-form KΩ2(A;Hom(E,C))K\in\Omega^{2}(A;\operatorname{Hom}(E,C)) is defined as

K(a1,a2)=θA([a1,a2]A)[θA(a1),θA(a2)]D.K(a_{1},a_{2})=\theta_{A}([a_{1},a_{2}]_{A})-[\theta_{A}(a_{1}),\theta_{A}(a_{2})]_{D}. (3.59)

A DVB morphism Ψ\Psi between decomposed vector bundles, E×MA×MCE×MA×MCE\times_{M}A\times_{M}C\to E^{\prime}\times_{M}A^{\prime}\times_{M}C^{\prime}, covering idM\operatorname{id}_{M}, is uniquely defined by restrictions of Ψ\Psi to the side bundles A,EA,E and the core CC and a 1-form χΩ1(A,Hom(E,C))=Γ(AEC)\chi\in\Omega^{1}(A,\operatorname{Hom}(E,C^{\prime}))=\operatorname{\Gamma}(A^{\ast}\otimes E^{\ast}\otimes C^{\prime}),

Ψ(e,a,c)=(Ψ|E(e),Ψ|A(a),Ψ|C(c)+χ(a,e)).\Psi(e,a,c)=(\Psi|_{E}(e),\Psi|_{A}(a),\Psi|_{C}(c)+\chi(a,e)). (3.60)

If A=AA=A^{\prime} and Ψ|A=idA\Psi|_{A}=\operatorname{id}_{A} then Ψ\Psi defines a graded VB morphism Φ0:C[0]E[1]C[0]E[1]\Phi_{0}:C_{[0]}\oplus E_{[1]}\rightarrow C^{\prime}_{[0]}\oplus E^{\prime}_{[1]}, Φ0=Ψ|CΨ|E\Phi_{0}=\Psi|_{C}\oplus\Psi|_{E}. Note that Hom(E,C)=Hom1(C[0]E[1],C[0]E[1])\operatorname{Hom}(E,C^{\prime})=\operatorname{Hom}_{-1}(C_{[0]}\oplus E_{[1]},C^{\prime}_{[0]}\oplus E^{\prime}_{[1]}).

Theorem 3.28.

Let AMA\to M be a Lie algebroid.

  1. (i)

    [GSM10] Let DD be a DVB as in (2.2), and θA\theta_{A} be a horizontal lift. (It gives rise to a decomposition DE×MA×MCD\simeq E\times_{M}A\times_{M}C.) Then the formulas (3.58) and (3.59) establish a one-to-one correspondence between algebroid structures on DED\to E that provide a VB-algebroid structure on the DVB DD and 2-term representations u.t.h. of the Lie algebroid AA on the complex :CE\partial:C\to E.

  2. (ii)

    [DJLO15] Let the decomposed DVBs D:A×ME×CD:A\times_{M}E\times_{C}, D=A×ME×CD^{\prime}=A^{\prime}\times_{M}E^{\prime}\times_{C}^{\prime} carry VB-algebroid structures and assume that the Lie algebroids AA, AA^{\prime} are the same. Then a DVB morphism Ψ:DD\Psi:D\to D^{\prime} such that Ψ|A=idA\Psi|_{A}=\operatorname{id}_{A}, is a VB-alegbroid morphism if and only if Φ=(Φ0,Φ1)\Phi=(\Phi_{0},\Phi_{1}), where Φ0=idAΨ|C\Phi_{0}=\operatorname{id}_{A}\oplus\Psi|_{C}, and Ψ1=χ\Psi_{1}=\chi, is a morphism between the associated 2-term representations.

Proof of Corollary 3.27.

Let us assume that we are given a VB-algebroid morphism Ψ\Psi, as above. Denote C:=Ψ|C:CTM\sharp^{C}:=\Psi|_{C}:C\to\mathrm{T}M. Let \nabla be any linear connection on AA. This corresponds to a decomposition :A×MTMTA\sum^{\nabla}:A\times_{M}\mathrm{T}M\to\mathrm{T}A of the DVB TA\mathrm{T}A. Thanks to presence and properties of Ψ\Psi, the DVB DD has a decomposition, induced by \nabla, as well. Indeed, Ψ\Psi is an affine bundle morphism from DD to TA\mathrm{T}A covering Ψ¯=idA×C:A×MCA×MTM\underline{\Psi}=\operatorname{id}_{A}\times\sharp^{C}:A\times_{M}C\to A\times_{M}\mathrm{T}M. It is fiber-wise bijective since Ψ\Psi is the identity on the core bundle AA. Hence, there exists a unique decomposition D:A×MCD\sum^{D}:A\times_{M}C\to D such that ΦD=Φ¯\Phi\circ\sum^{D}=\sum^{\nabla}\circ\underline{\Phi}.

In our case, the VB-algebroid structure on the DVB DD, given in (3.57), induces a representation of the Lie algebroid AA on the complex :AC\partial:A\to C. Besides, Ψ\Psi as a morphism of VB-algebroids, induces a morphism Φ=(Φ0,Φ1)\Phi=(\Phi_{0},\Phi_{1}) of 2-term representations, as described in Theorem 3.28. In our case, Φ1Ω1(A;Hom(C,A))\Phi_{1}\in\Omega^{1}(A;\operatorname{Hom}(C,A)) vanishes, since Ψ\Psi respects the decompositions of DD and TA\mathrm{T}A. Therefore, Φ\Phi is of the form described in Theorem 3.26, i.e., Φ=(Φ0,Φ1)\Phi=(\Phi_{0},\Phi_{1}), Φ1=0\Phi_{1}=0, Φ0|A=idA\Phi_{0}|_{A}=\operatorname{id}_{A}.

We shall prove that the AA-connection on AA in the complex ACA\to C is the same as in the adjoint representation. According to (3.58), these AA-connections on the core bundles of DD and TA\mathrm{T}A, denoted by core(D)\nabla^{\mathrm{core}(D)} and core(TA)\nabla^{\mathrm{core}(\mathrm{T}A)}, respectively, are given by

(a1core(D)a2)=[θAD(a1),a2]D,(a1core(TA)a2)=[θAD(a1),a2]TA\left(\nabla^{\mathrm{core}(D)}_{a_{1}}a_{2}\right)^{\dagger}=[\theta^{D}_{A}(a_{1}),a_{2}^{\dagger}]_{D},\quad\left(\nabla^{\mathrm{core}(\mathrm{T}A)}_{a_{1}}a_{2}\right)^{\dagger}=[\theta^{D}_{A}(a_{1}),a_{2}^{\dagger}]_{\mathrm{T}A}

where a1,a2Γ(A)a_{1},a_{2}\in\operatorname{\Gamma}(A). We have Ψ([θAD(a1),a2]D)=[θATA(a1),a2]TA\Psi([\theta_{A}^{D}(a_{1}),a_{2}^{\dagger}]_{D})=[\theta_{A}^{\mathrm{T}A}(a_{1}),a_{2}^{\dagger}]_{\mathrm{T}A} since Ψ:DTA\Psi:D\to\mathrm{T}A is a Lie algebroid morphism (covering the projection EME\to M), the corresponding decompositions of DVBs DD and TA\mathrm{T}A are Ψ\Psi-related (ΨθAD=θATA\Psi\circ\theta^{D}_{A}=\theta_{A}^{\mathrm{T}A}) and Ψ\Psi induces the identity on the core bundles. It follows that core(D)=core(TA)\nabla^{\mathrm{core}(D)}=\nabla^{\mathrm{core}(\mathrm{T}A)}.

Hence, due to Theorem 3.26, we get a HA structure on a certain canonically constructed graded bundle E2E^{2}, defined in Lemma 3.6. We shall prove that the obtained HA (E2,κ2)(E^{2},\kappa^{2}) does not depend on the choice of the linear connection \nabla on AA. It amounts to showing that the structure map

sv=sCvC(v)s\Box_{s}v=\nabla^{C}_{s}v-\partial\circ\nabla_{\sharp^{C}(v)}s

does not depends on the choice of \nabla. (The AA-connection C\nabla^{C} on CMC\to M associated with the VB-algebroid (D;C,A;M)(D;C,A;M) in (3.57) is the AA-connection denoted by Δside\Delta^{\mathrm{side}} in (3.58).) Let ~\widetilde{\nabla} be another linear connection on AA, so ~=:φHom(ATM,A)\widetilde{\nabla}-\nabla=:\varphi\in\operatorname{Hom}(A\otimes\mathrm{T}M,A). From [GSJLMM18, Remark 2.12], we find that

~sCvsCv=ϕ(s,v)Γ(C),\widetilde{\nabla}^{C}_{s}v-\nabla^{C}_{s}v=\partial\circ\phi(s,v)\in\operatorname{\Gamma}(C),

where :AC\partial:A\to C is as above and ϕHom(AC,A)\phi\in\operatorname{Hom}(A\otimes C,A) is the difference of the decompositions of the DVB DD induced by the linear connections ~\widetilde{\nabla} and \nabla. We have Φ(D(s,v))=TA(s,C(v))\Phi(\sum^{D}(s,v))=\sum^{\mathrm{T}A}(s,\sharp^{C}(v)), hence ϕ(s,v)=φ(s,C(v))\phi(s,v)=\varphi(s,\sharp^{C}(v)). Moreover,

(~C(v)sC(v)s)=φ(s,C(v)),\partial\circ\left(\widetilde{\nabla}_{\sharp^{C}(v)}s-\nabla_{\sharp^{C}(v)}s\right)=\partial\circ\varphi(s,\sharp^{C}(v)),

what finishes the proof of our claim, ~sv=sv\widetilde{\Box}_{s}v=\Box_{s}v, as ~svsv=~sCvsCv\widetilde{\Box}_{s}v-\Box_{s}v=\widetilde{\nabla}^{C}_{s}v-\nabla^{C}_{s}v, see Definition 3.24. ∎

Example 3.29.

We shall describe the representation u.t.h. of AA associated with the HA (A[2],κ[2])(A^{[2]},\kappa^{[2]}) – the 2nd2^{\mathrm{nd}} prolongation of a Lie algebroid (AM,[,],)(A\rightarrow M,[\cdot,\cdot],\sharp). From Example 3.8, it follows from that this is a representation on the complex =idA:AA[2]widehatA\partial=\operatorname{id}_{A}:A\rightarrow{\widehat{A^{[2]}}}\simeq A, and the AA-connections defined in Definition 3.24, denoted by A\nabla^{A} and C\nabla^{C}, coincide. Moreover, the 2-form KK given in Definition 3.24 is the curvature of the AA-connection A\nabla^{A}. Indeed, we know from [AC12] that curv(A)=Rbas\mathrm{curv}(\nabla^{A})={-}R_{\nabla}^{\mathrm{bas}}\circ\sharp while K=RbasCK=R_{\nabla}^{\mathrm{bas}}\circ\sharp^{C}, see (3.54), so curv(A)=K\mathrm{curv}(\nabla^{A})={-}K as =C\sharp=\sharp^{C} in our case.

Now consider the linearisation lin(A[2]){\operatorname{lin}}(A^{[2]}) of the graded bundle A[2]A^{[2]} as a DVB, where we shall recognize a VB-algebroid structure and a morphism to the adjoint representation corresponding to the HA structure on A[2]A^{[2]}, as described in Corollary 3.27. It was shown in [BGG15b, Theorem 2.3.8] that lin(A[k])A×TMTA[k1]{\operatorname{lin}}(A^{[k]})\simeq A\times_{\mathrm{T}M}\mathrm{T}A^{[k-1]} and that it carries a natural weighted algebroid structure. In the special case k=2k=2, we find that

lin(A[2])A×TMTA={(a,X)A×TA:a=(Tσ)X},{\operatorname{lin}}(A^{[2]})\simeq A\times_{\mathrm{T}M}\mathrm{T}A=\{(a,X)\in A\times\mathrm{T}A:\sharp a=(\mathrm{T}\sigma)X\},

and the DVB lin(A[2]){\operatorname{lin}}(A^{[2]}) carries a canonical structure of a VB-algebroid. Note that the side bundles and the core of lin(A[2]){\operatorname{lin}}(A^{[2]}) are naturally identified with the VB AMA\rightarrow M. Moreover, the Lie algebroid structure on the vector bundle pr1:A×TMTAA\operatorname{pr}_{1}:A\times_{\mathrm{T}M}\mathrm{T}A\rightarrow A, where pr1\operatorname{pr}_{1} is the projection onto the first factor, is a special case of the construction called the prolongation of a Lie algebroid, see [Mar01] and [BGG16]. The morphism Ψ\Psi of VB-algebroids has a straightforward form, Ψ:A×TMTATA\Psi:A\times_{\mathrm{T}M}\mathrm{T}A\rightarrow\mathrm{T}A is induced by the projection onto the second factor:

A×TMTA\textstyle{A\times_{\mathrm{T}M}\mathrm{T}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pr1\scriptstyle{\operatorname{pr}_{1}}τApr2\scriptstyle{\tau_{A}\circ\operatorname{pr}_{2}}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M\textstyle{M}   Ψ\scriptstyle{\Psi}   TA\textstyle{\mathrm{T}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τA\scriptstyle{\tau_{A}}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}TM\textstyle{\mathrm{T}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M\textstyle{M}

We shall illustrate now the procedure of reconstructing an HA from a given representation a Lie algebroid and a morphism to the adjoint representation.

Example 3.30.

We shall reconstruct a HA (E2,κ2)(E^{2},\kappa^{2}) out of the adjoint representation of a Lie algebroid (σ:AM,[,],)(\sigma:A\to M,[\cdot,\cdot],\sharp) and the morphism Φ\Phi being the identity on A[0](TM)[1]A_{[0]}\oplus(\mathrm{T}M)_{[1]}. According to Lemma 3.6, E2E^{2} is the quotient E2=A[2]×M(TM)[2]/E^{2}={\raisebox{1.79997pt}{$A^{[2]}\times_{M}(\mathrm{T}M)_{[2]}$}\left/\raisebox{-1.79997pt}{$\sim$}\right.} where the relation \sim is induced by the graph of :A[2]widehatATM=T2Mwidehat-\sharp:\widehat{A^{[2]}}\simeq A\rightarrow\mathrm{T}M=\widehat{T^{2}M}. Note that order-one reduction of E2E^{2} is AA, and its core is TM\mathrm{T}M. We can geometrically describe the graded bundle E2E^{2} as follows:

Take (X,v)(X,v) and (Y,w)(Y,w) in A[2]×M(TM)[2]A^{[2]}\times_{M}(\mathrm{T}M)_{[2]}. Then (X,v)(Y,w)(X,v)\sim(Y,w) if and only if τA(X)=τA(Y)\tau_{A}(X)=\tau_{A}(Y), (Tσ)X=(Tσ)Y(\mathrm{T}\sigma)X=(\mathrm{T}\sigma)Y, and (YX)=vw\sharp(Y-X)=v-w, where YXY-X is consider as an element of AA via the isomorphism between AA and 𝐕MATA\mathbf{V}_{M}A\subset\mathrm{T}A. (Recall, A[2]A^{[2]} consists of XTAX\in\mathrm{T}A such that (Tσ)X=τA(X)TM(\mathrm{T}\sigma)X=\sharp\tau_{A}(X)\in\mathrm{T}M.)

The structure maps of the HA (E2,κ2)(E^{2},\kappa^{2}) are easy to describe: =\partial=\sharp, C=idTM\sharp^{C}=\operatorname{id}_{\mathrm{T}M} and sv=sTMvvs=[s,v]\Box_{s}v=\nabla^{\mathrm{T}M}_{s}v-\partial\nabla_{v}s=[\sharp s,v], according to the definition of TM\nabla^{\mathrm{T}M}, see (3.52). Since (E2,κ2)(E^{2},\kappa^{2}) is a Lie HA we have ω=0\omega=0, and β(s1,s2)=[s1,s2]\beta(s_{1},s_{2})=\sharp[s_{1},s_{2}].

3.3 Final remarks and questions

The results presented in this paper (e.g. Theorems 3.263.15) are the source of new examples of order-two graded bundles and HAs, eg. Example 3.30, and raise questions about the classification of HAs under certain natural assumptions. This represents one potential direction for further development based on the findings of this paper.

Another avenue of research on HAs involves exploring how HAs of order 3\geq 3 are related to representations u.t.h. of Lie algebroids.

In light of the paper [BO19] on the integration of 2-term representations of Lie algebroids, a natural question arises about the integration of HAs. What higher-order groupoids are and how they relate to HAs?

Recall that HAs were introduced as geometric-algebraic structures providing a proper language to formulate a geometric formalism of higher-order variational calculus (generalizing the first-order case). We hope this work will encourage further developments in the area of HAs and higher-order geometric mechanics.

4 Appendix

In what follows, (xa,ywi)(x^{a},y^{i}_{w}) denotes graded coordinates on a graded bundle EkME^{k}\rightarrow M. In the case k=2k=2, we continue using the notation from Subsection 3.2 and Example 2.8. In particular, A=E1A=E^{1}, C=E2widehatC=\widehat{E^{2}}, and (ei)(e_{i}), (cμ)(c_{\mu}) denote local frames of the VBs AA and CC, respectively; (xa,yi,zμ)(x^{a},y^{i},z^{\mu}) are graded coordinates on E2E^{2} compatible with the chosen frames (ei)(e_{i}), (cμ)(c_{\mu}).

4.1 Vector fields of non-positive weight on graded bundles

In the following lemma, we study the structure of the space of vector fields of non-negative weight on a graded bundle EkE^{k}.

Lemma 4.1.

Let σk:EkM\sigma^{k}:E^{k}\rightarrow M be a graded bundle of order kk and let 𝔛0(Ek)=j=k0𝔛j(Ek)\mathfrak{X}_{\leq 0}(E^{k})=\bigoplus_{j=-k}^{0}\mathfrak{X}_{j}(E^{k}) denotes the Lie algebra of non-positively graded vector fields on EkE^{k}.

  1. (i)

    The Lie subalgebra 𝔛0(E)\mathfrak{X}_{0}(E) of linear vector fields on the total space EE of a vector bundle σ:EM\sigma:E\rightarrow M coincides with the Lie algebra of derivative endomorphisms of the dual bundle, 𝔛0(E)𝔇(E)\mathfrak{X}_{0}(E)\simeq\mathfrak{D}({E}^{\ast}).

  2. (ii)

    A vector field X𝔛0(Ek)X\in\mathfrak{X}_{0}(E^{k}) of weight zero is projectable onto EjE^{j} for any 0jk0\leq j\leq k, in particular on M=E0M=E^{0}.

  3. (iii)

    𝔛0(E2)\mathfrak{X}_{0}(E^{2}) is an abelian extension by Γ(Hom(Sym2E1,E2widehat))\operatorname{\Gamma}(\operatorname{Hom}(\operatorname{Sym}^{2}E^{1},\widehat{E^{2}})) of the Lie subalgebra of 𝔛0(E1)𝔛0(E2widehat)\mathfrak{X}_{0}(E^{1})\oplus\mathfrak{X}_{0}(\widehat{E^{2}}) consisting of pairs (X1,X2)(X_{1},X_{2}) such that X1X_{1} and X2X_{2} project onto the same vector field on MM:

    0Hom(Sym2E1,E2widehat)𝔛0(E2)𝜋𝔛0(E1)×𝔛(M)𝔛0(E2widehat)00\to\operatorname{Hom}(\operatorname{Sym}^{2}E^{1},\widehat{E^{2}})\to\mathfrak{X}_{0}(E^{2})\xrightarrow{\pi}\mathfrak{X}_{0}(E^{1})\times_{\mathfrak{X}(M)}\mathfrak{X}_{0}(\widehat{E^{2}})\to 0 (4.1)

    where the projection π\pi is given by π(X)=((Tσ21)(X),X|E2widehat)\pi(X)=((\mathrm{T}\sigma^{2}_{1})(X),X|_{\widehat{E^{2}}}) and the kernel of π\pi can be canonically identified with the space of VB morphisms Sym2E1E2widehat\operatorname{Sym}^{2}E^{1}\to\widehat{E^{2}}.

  4. (iv)

    There is a short exact sequence of graded Lie algebras

    0𝔛𝐕<0(Ek)𝔛<0(Ek)𝔛<0(Ek1)00\to\mathfrak{X}^{\mathbf{V}}_{<0}(E^{k})\to\mathfrak{X}_{<0}(E^{k})\to\mathfrak{X}_{<0}(E^{k-1})\to 0 (4.2)

    where 𝔛𝐕<0(Ek)\mathfrak{X}^{\mathbf{V}}_{<0}(E^{k}) denotes the subspace of 𝔛<0(Ek)\mathfrak{X}_{<0}(E^{k}) of those vector fields which are vertical with respect to the projection σkk1:EkEk1\sigma^{k}_{k-1}:E^{k}\rightarrow E^{k-1}.

  5. (v)

    In case k=2k=2, the homogeneous part of weight 1-1 of (4.2) reads as

    0𝔛𝐕1(E2)Hom(E1,E2widehat)𝔛1(E2)𝔛1(E1)Γ(E1)00\to\mathfrak{X}^{\mathbf{V}}_{-1}(E^{2})\simeq\operatorname{Hom}(E^{1},\widehat{E^{2}})\to\mathfrak{X}_{-1}(E^{2})\to\mathfrak{X}_{-1}(E^{1})\simeq\operatorname{\Gamma}(E^{1})\to 0
Proof.

Point (i) is well known, see e.g. [KSM02] or [EVT19, Remark 2.1]. For the proof of (ii), write a vector field X𝔛0(Ek)X\in\mathfrak{X}_{0}(E^{k}) in a general local form

X=fa(x)xa+fi(x,y)yiw,X=f^{a}(x)\partial_{x^{a}}+f^{i}(x,y)\partial_{y^{i}_{w}},

where functions fi(x,y)f_{i}(x,y) are homogenous of weight w=w(yiw)w=\mathrm{w}(y^{i}_{w}). It follows that the function fi(x,y)f_{i}(x,y) does not depend on coordinates of weights greater than w(yi)\mathrm{w}(y^{i}), so it is the pullback of a function on Ew(i)E^{\mathrm{w}(i)}.

Obviously, Tσk\mathrm{T}\sigma^{k} annihilates yiw\partial_{y^{i}_{w}} and defines a projection of the vector field XX onto MM, which is fa(x)xaf_{a}(x)\partial_{x^{a}}. A very similar proof works for the projections σkj:EkEj\sigma^{k}_{j}:E^{k}\rightarrow E^{j} where j>0j>0. For the proof of (iii) write X𝔛0(E2)X\in\mathfrak{X}_{0}(E^{2}) in the form

X=fa(x)xa+fij(x)yjyi+(fμν(x)zν+fμij(x)yiyj)zμ.X=f^{a}(x)\partial_{x^{a}}+f^{i}_{j}(x)y^{j}\partial_{y^{i}}+\left(f^{\mu}_{\nu}(x)z^{\nu}+f^{\mu}_{ij}(x)y^{i}y^{j}\right)\partial_{z_{\mu}}.

It follows that the vector field XX restricted to the submanifold E2widehat\widehat{E^{2}} is tangent to it and X|E2widehat=fa(x)xa+fμν(x)zνzμX{|}_{\widehat{E^{2}}}=f^{a}(x)\partial_{x^{a}}+f^{\mu}_{\nu}(x)z^{\nu}\partial_{z_{\mu}}. Besides, (Tσ21)X=fa(x)xa+fij(x)yjyi(\mathrm{T}\sigma^{2}_{1})X=f^{a}(x)\partial_{x^{a}}+f^{i}_{j}(x)y^{j}\partial_{y^{i}}, hence X|E2widehatX{|}_{\widehat{E^{2}}} and (Tσ21)X(\mathrm{T}\sigma^{2}_{1})X project to the same vector field on MM. Moreover, the kernel of the projection π\pi consists of vector fields of the form fμij(x)yiyjzμf^{\mu}_{ij}(x)y^{i}y^{j}\partial_{z_{\mu}} which can be identified with a VB morphism from Hom(C,Sym2(E1))=Hom(Sym2E1,C)\operatorname{Hom}(C^{\ast},\operatorname{Sym}^{2}(E^{1})^{\ast})=\operatorname{Hom}(\operatorname{Sym}^{2}E^{1},C), where C=E2widehatC=\widehat{E^{2}}.

For (iv) it is enough to notice that a vector field X𝔛(Ek)X\in\mathfrak{X}(E^{k}) of weight 1\leq-1 has a well defined projection on Ek1E^{k-1}. Point (v) is a direct consequence of (iv). ∎

4.2 Leibniz-type identities of the structure maps of HAs

Proof of Lemma 3.12.

Let (ej)(e_{j}) and (e¯j)(\underline{e}_{j}) be local frames of sections of the vector bundle AMA\rightarrow M, related by ej=Tij(x)e¯ie_{j}=T^{i}_{j}(x)\underline{e}_{i}. Let (cμ)(c_{\mu}) be a frame of the vector bundle CMC\rightarrow M. The graded bundle  morphism Φ:A[2]C[2]\Phi:A^{[2]}\to C_{[2]} has the local form

Φ(xa,yi,y˙i)=(Φμi(x)y˙i+12Φμij(x)yiyj)cμ,\Phi(x^{a},y^{i},\dot{y}^{i})=(\Phi^{\mu}_{i}(x)\dot{y}^{i}+\frac{1}{2}\Phi^{\mu}_{ij}(x)y^{i}y^{j})c_{\mu},

where Φμij=Φμji\Phi^{\mu}_{ij}=\Phi^{\mu}_{ji}. On the other hand, a map Ψ:Γ(A)×Γ(A)Γ(C)\Psi:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(A)\rightarrow\operatorname{\Gamma}(C) satisfying the Leibniz-type identity (3.39) is locally determined by the VB morphism ρ\rho and local functions Ψμij\Psi^{\mu}_{ij}, where Ψμij=Ψμji\Psi^{\mu}_{ij}=\Psi^{\mu}_{ji}, as follows:

Ψ(ei,ej)=Ψμij(x)cμ.\Psi(e_{i},e_{j})=\Psi^{\mu}_{ij}(x)c_{\mu}.

In the given correspondence, ρ\rho corresponds to the core VB morphism Φwidehat:AC\widehat{\Phi}:A\to C, via the isomorphism A[2]widehatA\widehat{A^{[2]}}\simeq A. To complete the proof, we shall show that the change, (ej)(e¯j)(e_{j})\mapsto(\underline{e}_{j}), of local frames of Γ(A)\operatorname{\Gamma}(A) results in the same transition functions for the local functions (Φμij)(\Phi^{\mu}_{ij}) as for (Ψμij)(\Psi^{\mu}_{ij}). By calculating the differential of y¯i\underline{y}^{i}, we find that the local coordinates (xa,yi,y˙i)(x^{a},y^{i},\dot{y}^{i}) on A[2]A^{[2]} transform as x¯a=xa\underline{x}^{a}=x^{a}, y¯i=Tijyj\underline{y}^{i}=T^{i}_{j}y^{j},

y˙¯i=Tijy˙j+12αijkyjyk,where αijk=TijxaQak+TikxaQaj.\underline{\dot{y}}^{i}=T^{i}_{j}\dot{y}^{j}+\frac{1}{2}\alpha^{i}_{jk}y^{j}y^{k},\text{where }\alpha^{i}_{jk}=\frac{\partial T^{i}_{j}}{x^{a}}Q^{a}_{k}+\frac{\partial T^{i}_{k}}{x^{a}}Q^{a}_{j}.

It follows that if Φ¯μiy˙¯i+12Φ¯μijy¯iy¯j=Φμjy˙j+12Φμjkyjyk\underline{\Phi}^{\mu}_{i}\underline{\dot{y}}^{i}+\frac{1}{2}\underline{\Phi}^{\mu}_{ij}\underline{y}^{i}\underline{y}^{j}=\Phi^{\mu}_{j}\dot{y}^{j}+\frac{1}{2}\Phi^{\mu}_{jk}y^{j}y^{k} then Φμj=Φ¯μiTij\Phi^{\mu}_{j}=\underline{\Phi}^{\mu}_{i}T^{i}_{j}, and

Φμjk=Φ¯μiαijk+Φ¯μj,kTjjTkk.\Phi^{\mu}_{jk}=\underline{\Phi}^{\mu}_{i}\alpha^{i}_{jk}+\underline{\Phi}^{\mu}_{j^{\prime},k^{\prime}}T^{j^{\prime}}_{j}T^{k^{\prime}}_{k}.

On the other hand,

Ψ(ej,ek)=Ψ(Tjje¯j,Tkke¯k)=TjjTkkΨ¯μjkcμ+12(ej)(Tkk)ρ(e¯k)+12(ek)(Tjj)ρ(e¯j)=cμ(TjjTkkΨ¯μjk+12αjkiρ¯μi).\begin{split}\Psi(e_{j},e_{k})=\Psi(T^{j^{\prime}}_{j}\underline{e}_{j^{\prime}},T^{k^{\prime}}_{k}\underline{e}_{k^{\prime}})=T^{j^{\prime}}_{j}T^{k^{\prime}}_{k}\underline{\Psi}^{\mu}_{j^{\prime}k^{\prime}}c_{\mu}+\frac{1}{2}(\sharp e_{j})(T^{k^{\prime}}_{k})\rho(\underline{e}_{k^{\prime}})+\frac{1}{2}(\sharp e_{k})(T^{j^{\prime}}_{j})\rho(\underline{e}_{j^{\prime}})=\\ c_{\mu}\left(T^{j^{\prime}}_{j}T^{k^{\prime}}_{k}\underline{\Psi}^{\mu}_{j^{\prime}k^{\prime}}+\frac{1}{2}\alpha_{jk}^{i}\underline{\rho}^{\mu}_{i}\right).\end{split}

Therefore, the transformations for Ψμij\Psi^{\mu}_{ij} are the same as those for Φμij\Phi^{\mu}_{ij}, as we claimed. ∎

The following three lemmas concern the calculus with algebroid lifts introduced in (2.29). The first one, Lemma 4.2, is the most general – we do not assume any HA structure.

Lemma 4.2.

Let kk\in\mathbb{N} and MM be a smooth manifold.

  1. (i)

    If X𝔛(M)X\in\mathfrak{X}(M), f𝒞(M)f\in\mathcal{C}^{\infty}(M) then

    (τkM)X(f)=1k!Xk(f(k)),{(\tau^{k}_{M})}^{*}X(f)=\frac{1}{k!}{X}^{\langle{-k}\rangle}(f^{(k)}), (4.3)

    where Xk𝔛k(TkM){X}^{\langle{-k}\rangle}\in\mathfrak{X}_{-k}(\mathrm{T}^{k}M) is (TkM,κkM)(\mathrm{T}^{k}M,\kappa^{k}_{M})-lift of XX in weight k-k.

  2. (ii)

    Let ρk:EkTkM\rho^{k}:E^{k}\to\mathrm{T}^{k}M be any morphism of graded bundles covering idM\operatorname{id}_{M}. Let vΓ(Ekwidehat)v\in\operatorname{\Gamma}(\widehat{E^{k}}), f𝒞(M)f\in\mathcal{C}^{\infty}(M). Then

    (v)((ρk)f(k))=ρkwidehat(v)(f),({v^{\uparrow}})({(\rho^{k})}^{*}f^{(k)})=\widehat{\rho^{k}}(v)(f), (4.4)

    where v{v^{\uparrow}} is the image of vv in 𝔛k(Ek)\mathfrak{X}_{-k}(E^{k}) (see Lemma 2.6) and ρkwidehat(v)Γ(TkMwidehat)𝔛(M)\widehat{\rho^{k}}(v)\in\operatorname{\Gamma}(\widehat{\mathrm{T}^{k}M})\simeq\mathfrak{X}(M) is understood as a vector field on MM thanks to the isomorphism jkM:TMTkMwidehatj^{k}_{M}:\mathrm{T}M\to\widehat{\mathrm{T}^{k}M} given in (2.7).

Proof.

(i) Recall that the vector field 1k!Xk\frac{1}{k!}{X}^{\langle{-k}\rangle} is constructed in two steps. First, we take the vertical lift X(0)ΓTkM(TkE)X^{(0)}\in\operatorname{\Gamma}_{\mathrm{T}^{k}M}(\mathrm{T}^{k}E) of XX where E=TME=\mathrm{T}M, and then we compose it with κkM:TkTMTTkM\kappa^{k}_{M}:\mathrm{T}^{k}\mathrm{T}M\rightarrow\mathrm{T}\mathrm{T}^{k}M, see (2.29). We shall describe 1k!Xk\frac{1}{k!}{X}^{\langle{-k}\rangle} by means of its flow.

We take E=TME=\mathrm{T}M and α=k\alpha=k in (2.25) and read from (2.26) that the vertical lift X(0)X^{(0)} sends [γ]kTkM[\gamma]_{k}\in\mathrm{T}^{k}M to [t1k!tkXγ(t)]kTkγ(t)TM[t\mapsto{\frac{1}{k!}}t^{k}X_{\gamma(t)}]_{k}\in\mathrm{T}^{k}_{\gamma(t)}\mathrm{T}M. Hence, the vector field 1k!Xk𝔛(TkM)\frac{1}{k!}{X}^{\langle{-k}\rangle}\in\mathfrak{X}(\mathrm{T}^{k}M) is given by

1k!Xk[γ]k=[u[tϕXutk/k!(γ(t))]k]1T[γ]kTkM,\frac{1}{k!}{X}^{\langle{-k}\rangle}_{[\gamma]_{k}}=[u\mapsto[t\mapsto\phi^{X}_{ut^{k}/k!}(\gamma(t))]_{k}]_{1}\in\mathrm{T}_{{[\gamma]_{k}}}\mathrm{T}^{k}M,

where (t,x)ϕXt(x)(t,x)\mapsto\phi^{X}_{t}(x), xMx\in M, is the flow of the vector field XX. Hence,

1k!Xk(f(k))=ddu|u=0f(k)([t1k!ϕXutk/k!(γ(t))]k)=dkdtk|t=0tk/k!ddu|u=0f(ϕXu(γ(t)))==ddu|u=0f(ϕXu(γ(0)))=X(f)(γ(0))\frac{1}{k!}{X}^{\langle{-k}\rangle}(f^{(k)})=\left.{\frac{\mathrm{d}}{\mathrm{d}u}}\right|_{u=0}f^{(k)}([t\mapsto\frac{1}{k!}\phi^{X}_{ut^{k}/k!}(\gamma(t))]_{k})=\left.\frac{\mathrm{d}^{k}}{\mathrm{d}t^{k}}\right|_{t=0}t^{k}/k!\,\left.{\frac{\mathrm{d}}{\mathrm{d}u}}\right|_{u=0}f(\phi^{X}_{u}(\gamma(t)))=\\ =\left.{\frac{\mathrm{d}}{\mathrm{d}u}}\right|_{u=0}f(\phi^{X}_{u}(\gamma(0)))=X(f)(\gamma(0))

as we claimed.

(ii): First of all, note that the function (v)((ρk)f(k))𝒞(Ek)({v^{\uparrow}})({(\rho^{k})}^{*}f^{(k)})\in\mathcal{C}^{\infty}(E^{k}) has weight k+k=0-k+k=0; hence, it is the pullback of a function on the base MM, and it is enough to verify the equality (4.4) at a point mMm\in M.

The tangent vector (v)mTmEk({v^{\uparrow}})_{m}\in\mathrm{T}_{m}E^{k} is represented by the curve ttEkwidehatvmt\mapsto t\cdot_{\widehat{E^{k}}}v_{m}, which is equal to htkEk(vm)h_{\sqrt[k]{t}}^{E^{k}}(v_{m}) if t0t\geq 0, where hEkh^{E^{k}} is the homogeneity structure on EkE^{k}. Hence,

LHS of (4.4)=ddt|t=0f(k)(ρk(tEkwidehatvm)).\text{LHS\ of \eqref{e:rhoVF}}=\left.\frac{\mathrm{d}}{\mathrm{d}t}\right|_{t=0}f^{(k)}(\rho^{k}(t\cdot_{\widehat{E^{k}}}v_{m})).

Assume that the image of ρkwidehat(vm)\widehat{\rho^{k}}(v_{m}) in TmMwidehat\widehat{\mathrm{T}_{m}M} is represented by a curve γ:M\gamma:\mathbb{R}\to M, γ(0)=m\gamma(0)=m, i.e.,

RHS of (4.4)=dds|s=0f(γ(s)).\text{RHS\ of \eqref{e:rhoVF}}=\left.\frac{\mathrm{d}}{\mathrm{d}s}\right|_{s=0}f(\gamma(s)).

Then ρk(vm)=ρkwidehat(vm)=[sγ(sk/k!)]k\rho^{k}(v_{m})=\widehat{\rho^{k}}(v_{m})=[s\mapsto\gamma(s^{k}/k!)]_{k} as kthk^{\mathrm{th}}-velocity in TkMwidehatTkM\widehat{\mathrm{T}^{k}M}\subset\mathrm{T}^{k}M, hence

ρk(tEkwidehatvm)=tTkMwidehatρk(vm)=[sγ(tsk/k!)]k.\rho^{k}(t\cdot_{\widehat{E^{k}}}v_{m})=t\cdot_{\widehat{\mathrm{T}^{k}M}}\rho^{k}(v_{m})=[s\mapsto\gamma(t\cdot s^{k}/k!)]_{k}.

Therefore,

LHS of (4.4)=ddt|t=0dkdsk|s=0f(γ(tsk/k!)).\text{LHS\ of \eqref{e:rhoVF}}=\left.\frac{\mathrm{d}}{\mathrm{d}t}\right|_{t=0}\left.\frac{\mathrm{d}^{k}}{\mathrm{d}s^{k}}\right|_{s=0}f(\gamma(ts^{k}/k!)).

Let (xa)(x^{a}) be local coordinates on MM around mm such that xa(m)=0x^{a}(m)=0. It is enough to prove (4.4) for f=xaf=x^{a}. If γa(t):=xa(γ(t))=cat+o(t)\gamma^{a}(t):=x^{a}(\gamma(t))=c^{a}t+o(t) then dkdsk|s=0γa(tsk/k!)=cat+o(t)\left.\frac{\mathrm{d}^{k}}{\mathrm{d}s^{k}}\right|_{s=0}\gamma^{a}(ts^{k}/k!)=c^{a}t+o(t). Hence, the left and right hand sides of (4.4) coincide with cac^{a}. ∎

Lemma 4.3.

Let (Ek,κk)(E^{k},\kappa^{k}) be an AL HA.

  1. (i)

    For sΓ(E1)s\in\operatorname{\Gamma}(E^{1}) and f𝒞(M)f\in\mathcal{C}^{\infty}(M), the following identities hold

    1k!sk((k)f(k))=(σk)1(s)(f)=1α!sα((k)f(α))\frac{1}{k!}{s}^{\langle{-k}\rangle}((\sharp^{k})^{\ast}f^{(k)})={(\sigma^{k})^{\ast}}\sharp^{1}(s)(f)=\frac{1}{{\alpha}!}{s}^{\langle{-{\alpha}}\rangle}((\sharp^{k})^{\ast}f^{({\alpha})})

    for any 0αk10\leq{\alpha}\leq k-1.

  2. (ii)

    1=Ekwidehatk\sharp^{1}=\sharp^{\widehat{E^{k}}}\circ\partial^{k} where Ekwidehat:EkwidehatTM\sharp^{\widehat{E^{k}}}:\widehat{E^{k}}\rightarrow\mathrm{T}M and k:E1Ekwidehat\partial^{k}:E^{1}\rightarrow\widehat{E^{k}} are the VB morphisms given in Remark 3.9.

Proof.

Proof of (i): Denote X=1s𝔛(M)X=\sharp^{1}s\in\mathfrak{X}(M) for time being. We know from Theorem 2.11 that the vector fields sk{s}^{\langle{-k}\rangle} and Xk{X}^{\langle{-k}\rangle} are k\sharp^{k}-related, hence for any function ψ𝒞(TkM)\psi\in\mathcal{C}^{\infty}(\mathrm{T}^{k}M) we have sk((k)ψ)=(k)Xk(ψ){s}^{\langle{-k}\rangle}((\sharp^{k})^{\ast}\psi)=(\sharp^{k})^{\ast}{X}^{\langle{-k}\rangle}(\psi). We take ψ=f(k)\psi=f^{(k)}, use (4.3) and get

1k!sk((k)f(k))=1k!(k)Xk(f(k))=(k)(τkM)X(f)=(σk)X(f)\frac{1}{k!}{s}^{\langle{-k}\rangle}((\sharp^{k})^{\ast}f^{(k)})=\frac{1}{k!}(\sharp^{k})^{\ast}{X}^{\langle{-k}\rangle}(f^{(k)})=(\sharp^{k})^{\ast}(\tau^{k}_{M})^{\ast}X(f)=(\sigma^{k})^{\ast}X(f)

as σk=τkMk\sigma^{k}=\tau^{k}_{M}\circ\sharp^{k}. This proves the first equality.

The second one follows from Lemma 2.14. Indeed, consider the reduction of (Ek,κk)(E^{k},\kappa^{k}) to weight α\alpha. We find that the vector field sα𝔛α(Ek){s}^{\langle{-{\alpha}}\rangle}\in\mathfrak{X}_{-{\alpha}}(E^{k}) is projectable onto EαE^{{\alpha}} and its projection is sακαs^{\langle-{\alpha}\rangle_{\kappa^{{\alpha}}}} hence the equality (σk)1(s)(f)=1α!sα((k)f(α))(\sigma^{k})^{\ast}\sharp^{1}(s)(f)=\frac{1}{{\alpha}!}{s}^{\langle{-{\alpha}}\rangle}((\sharp^{k})^{\ast}f^{({\alpha})}) follows from the previous one by replacing kk with α{\alpha}.

Proof of (ii): The claim follows from the commutativity of the following diagram

Γ(E1)\textstyle{\operatorname{\Gamma}(E^{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ek\scriptstyle{\partial_{E^{k}}}\scriptstyle{\sharp}𝔛k(Ek)\textstyle{\mathfrak{X}_{-k}(E^{k})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}k\scriptstyle{\sharp^{k}}\scriptstyle{\simeq}Γ(Ekwidehat)\textstyle{\operatorname{\Gamma}(\widehat{E^{k}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}kwidehat\scriptstyle{\widehat{\sharp^{k}}}𝔛(M)\textstyle{\mathfrak{X}(M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}TkM\scriptstyle{\partial_{\mathrm{T}^{k}M}}𝔛k(TkM)\textstyle{\mathfrak{X}_{-k}(\mathrm{T}^{k}M)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\simeq}Γ(TkMwidehat)\textstyle{\operatorname{\Gamma}(\widehat{T^{k}M})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\simeq}𝔛(M)\textstyle{\mathfrak{X}(M)}

where the arrow in the middle, labelled by k\sharp^{k}, denotes a relation: (X,Y)k(X,Y)\in\sharp^{k} if the vector fields X𝔛(Ek)X\in\mathfrak{X}(E^{k}) and Y𝔛(TkM)Y\in\mathfrak{X}(\mathrm{T}^{k}M) are k\sharp^{k}-related. Actually, this relation restricted to the lowest degree k-k becomes a mapping k:𝔛k(Ek)𝔛k(TkM)\sharp^{k}:\mathfrak{X}_{-k}(E^{k})\rightarrow\mathfrak{X}_{-k}(\mathrm{T}^{k}M). Moreover, Ek=k\partial_{E^{k}}=\partial^{k} and TkM{\partial_{\mathrm{T}^{k}M}} are defined by means of algebroid lifts, as in Remark 3.9. It follows from Lemma 2.16 that the composition of maps in the lower row is the identity on 𝔛(M)\mathfrak{X}(M). All maps in the diagram are 𝒞(M)\mathcal{C}^{\infty}(M)-linear, hence they give rise to VB morphisms. The square on the left is commutative due to Theorem 2.11 and the AL assumption. The square on the right is also commutative and it is a more general fact: k\sharp^{k} can be replaced there with any graded bundle morphism ρk:EkFk\rho^{k}:E^{k}\rightarrow F^{k}, as stated in Lemma 2.6. ∎

Recall, that for k=2k=2, the formula (2.31) gives

(fs)2=fs2=2f(s),(fs)1=fs1+(f˙)s2,(fs)0=fs0+(f˙)s1+12((2)f¨)s2.\begin{split}{(fs)}^{\langle{-2}\rangle}&=f{s}^{\langle{-2}\rangle}=2f\partial(s),\\ {(fs)}^{\langle{-1}\rangle}&=f{s}^{\langle{-1}\rangle}+({\sharp}^{*}\dot{f}){s}^{\langle{-2}\rangle},\\ {(fs)}^{\langle{0}\rangle}&=f{s}^{\langle{0}\rangle}+({\sharp}^{*}\dot{f}){s}^{\langle{-1}\rangle}+\frac{1}{2}((\sharp^{2})^{\ast}\ddot{f}){s}^{\langle{-2}\rangle}.\end{split} (4.5)

We shall need the following lemma for proving tensor-like properties of some structure maps associated with a skew HA (E2,κ2)(E^{2},\kappa^{2}).

Lemma 4.4.

Let (E2,κ2)(E^{2},\kappa^{2}) be a skew HA. Let f𝒞(M)f\in\mathcal{C}^{\infty}(M), s,s1,s2Γ(A)s,s_{1},s_{2}\in\operatorname{\Gamma}(A) and vΓ(C)𝔛2(E2)v\in\operatorname{\Gamma}(C)\simeq\mathfrak{X}_{-2}(E^{2}). Then

  1. (i)

    (Cv)(f)=v((2)f¨)(\sharp^{C}v)(f)={v^{\uparrow}}((\sharp^{2})^{\ast}\ddot{f}), where v{v^{\uparrow}} is given in Lemma 2.6,

  2. (ii)

    12s2((2)f¨)=(C)(s)(f)\frac{1}{2}{s}^{\langle{-2}\rangle}((\sharp^{2})^{\ast}\ddot{f})=(\sharp^{C}\circ\partial)(s)(f),

  3. (iii)

    (s)(f)=s1(f˙)=s0(f)(\sharp s)(f)={s}^{\langle{-1}\rangle}({\sharp}^{*}\dot{f})={s}^{\langle{0}\rangle}(f),

  4. (iv)

    s11s20(f˙)=([s1,s2]+s2s1)(f){s_{1}}^{\langle{-1}\rangle}{s_{2}}^{\langle{0}\rangle}({\sharp}^{*}\dot{f})=\left(\sharp[s_{1},s_{2}]+\sharp s_{2}\circ\sharp s_{1}\right)(f).

Remark 4.5.

We consider f𝒞(M)f\in\mathcal{C}^{\infty}(M) and (f˙){\sharp}^{*}(\dot{f}) as functions on E2E^{2} using the pullbacks of ff by σ2:E2M\sigma^{2}:E^{2}\rightarrow M and σ21:E2E1\sigma^{2}_{1}:E^{2}\rightarrow E^{1}, respectively. Note that if (E2,κ2)(E^{2},\kappa^{2}) is AL, then C=\sharp^{C}\circ\partial=\sharp, hence (ii) coincides in this case with Lemma 4.3 (i) with k=2k=2. It is tempting to add in (iii) the equality (s)(f)=12s2((2)f¨)(\sharp s)(f)=\frac{1}{2}{s}^{\langle{-2}\rangle}(\left(\sharp^{2}\right)^{\ast}\ddot{f}), but this requires the assumption (AL\mathrm{AL_{\partial}}), see Lemma 4.3. In the AL case, point (iv) simplifies to s11s20(f˙)=s1(s2(f)){s_{1}}^{\langle{-1}\rangle}{s_{2}}^{\langle{0}\rangle}({\sharp}^{*}\dot{f})=\sharp s_{1}(\sharp s_{2}(f)).

Proof.

In general, the (α)(\alpha)-lift f(α)𝒞(TkM)f^{(\alpha)}\in\mathcal{C}^{\infty}(\mathrm{T}^{k}M) has weight 0αk0\leq\alpha\leq k, the anchor map k:EkTkM\sharp^{k}:E^{k}\rightarrow\mathrm{T}^{k}M preserves the gradings on EkE^{k} and TkM\mathrm{T}^{k}M, while the vector field sβ𝔛(Ek){s}^{\langle{\beta}\rangle}\in\mathfrak{X}(E^{k}), where sΓ(E1)s\in\operatorname{\Gamma}(E^{1}), has weight kβ0-k\leq\beta\leq 0. Hence sβ(f(α))=0{s}^{\langle{\beta}\rangle}(f^{(\alpha)})=0 whenever α+β<0\alpha+\beta<0.

Clearly, the equation (i) is 𝒞(M)\mathcal{C}^{\infty}(M)-linear in vv, so it is enough to show (i) for vv from a frame (cμ)(c_{\mu}) of local sections of CMC\rightarrow M. We have f¨=fxax¨a+2fxaxbx˙ax˙b\ddot{f}=\frac{\partial f}{\partial x^{a}}\ddot{x}^{a}+\frac{\partial^{2}f}{\partial x^{a}\partial x^{b}}\dot{x}^{a}\dot{x}^{b}, so

(2)(f¨)=fxa(Qaμzμ+12Qaijyiyj)+2fxaxbQaiQbjyiyj,(\sharp^{2})^{\ast}(\ddot{f})=\frac{\partial f}{\partial x^{a}}\left(Q^{a}_{\mu}z^{\mu}+\frac{1}{2}Q^{a}_{ij}y^{i}y^{j}\right)+{\frac{\partial^{2}f}{\partial x^{a}\partial x^{b}}Q^{a}_{i}Q^{b}_{j}y^{i}y^{j},} (4.6)

hence cμ((2)(f¨))=Qaμfxa=C(cμ)(f){c_{\mu}^{\uparrow}}((\sharp^{2})^{\ast}(\ddot{f}))=Q^{a}_{\mu}\frac{\partial f}{\partial x^{a}}=\sharp^{C}(c_{\mu})(f), see (3.14), and (i) follows immediately.

Similarly, using (4.5), we find that (ii) and (iii) are 𝒞(M)\mathcal{C}^{\infty}(M)-linear in ss. Thus it is enough to verify these equalities for s=eks=e_{k}. This is straightforward: we use the formulas (3.18)\eqref{e:alg_lifts_coord} for ekα{e_{k}}^{\langle{\alpha}\rangle} and find that 12ek2((2)f¨)=QμkQaμfxa\frac{1}{2}{e_{k}}^{\langle{-2}\rangle}((\sharp^{2})^{\ast}\ddot{f})=Q^{\mu}_{k}Q^{a}_{\mu}\frac{\partial f}{\partial x^{a}}, which coincides with (C)(ek)(f)=QμkC(cμ)(f)(\sharp^{C}\circ\partial)(e_{k})(f)=Q^{\mu}_{k}\sharp^{C}(c^{\mu})(f) due to (3.19), thereby proving (ii). Next, f˙=fxaQaiyi{\sharp}^{*}\dot{f}=\frac{\partial f}{\partial x^{a}}Q^{a}_{i}y^{i}, and all three expressions in (iii) are equal to QakfxaQ^{a}_{k}\frac{\partial f}{\partial x^{a}}.

It remains to prove (iv). We claim that the left and right hand sides of (iv) are 𝒞(M)\mathcal{C}^{\infty}(M)-linear in s1s_{1}, and their difference is also 𝒞(M)\mathcal{C}^{\infty}(M)-linear in s2s_{2}. Indeed, consider (iv) with s1s_{1} replaced with gs1gs_{1} and expand (gs1)1{(gs_{1})}^{\langle{-1}\rangle} as in (4.5). Note that s20(f˙){s_{2}}^{\langle{0}\rangle}({\sharp}^{*}\dot{f}) has weight 1 and so it is killed by s12{s_{1}}^{\langle{-2}\rangle}, hence (gs1)1s20(f˙)=gs11s20(f˙){(g\,s_{1})}^{\langle{-1}\rangle}{s_{2}}^{\langle{0}\rangle}({\sharp}^{*}\dot{f})=g\,{s_{1}}^{\langle{-1}\rangle}{s_{2}}^{\langle{0}\rangle}({\sharp}^{*}\dot{f}) while

[gs1,s2]+s2(gs1)=(g[s1,s2](s2)(g)s1)+s2(g)s1)+gs2s1=gRHS(iv).\sharp[gs_{1},s_{2}]+\sharp s_{2}\circ\sharp(gs_{1})=\left(g\sharp[s_{1},s_{2}]-(\sharp s_{2})(g)\,\sharp s_{1}\right)+\sharp s_{2}(g)\,\sharp s_{1})+g\sharp s_{2}\circ\sharp s_{1}=g\,\cdot\mathrm{RHS}_{\eqref{e:sharp_lift2ddf}}.

Similarly, using point (iii) and a weight argument, we obtain

s11(gs2)0(f˙)\displaystyle{s_{1}}^{\langle{-1}\rangle}{(gs_{2})}^{\langle{0}\rangle}({\sharp}^{*}\dot{f}) =s11(gs20+(g˙)s21+12(2)(g¨)s22)(f˙)=\displaystyle={s_{1}}^{\langle{-1}\rangle}\left(g{s_{2}}^{\langle{0}\rangle}+{\sharp}^{*}(\dot{g}){s_{2}}^{\langle{-1}\rangle}+\frac{1}{2}(\sharp^{2})^{\ast}(\ddot{g}){s_{2}}^{\langle{-2}\rangle}\right)({\sharp}^{*}\dot{f})=
(gs11s20+s11(g˙)s21)(f˙)=gLHS(iv)+(s1)(g)(s2)(f),\displaystyle(g{s_{1}}^{\langle{-1}\rangle}{s_{2}}^{\langle{0}\rangle}+{s_{1}}^{\langle{-1}\rangle}({\sharp}^{*}\dot{g})\,{s_{2}}^{\langle{-1}\rangle})({\sharp}^{*}\dot{f})=g\cdot\mathrm{LHS}_{\eqref{e:sharp_lift2ddf}}+(\sharp s_{1})(g)\cdot(\sharp s_{2})(f),

and, in the same way, ([s1,gs2]+(gs2)(s1))(f)=gRHS(iv)+(s1)(g)(s2)(f)\left(\sharp[s_{1},gs_{2}]+\sharp(gs_{2})\circ(\sharp s_{1})\right)(f)=g\cdot\mathrm{RHS}_{\eqref{e:sharp_lift2ddf}}+(\sharp s_{1})(g)\cdot(\sharp s_{2})(f). It proves our claim, and thus it is enough to check (iv) with s1=eks_{1}=e_{k^{\prime}} and s2=eks_{2}=e_{k}. We have

ek0(f˙)=ek0(fxaQaiyi)=Qbkyixb(fxaQai)+QijkQaiyjfxa,{e_{k}}^{\langle{0}\rangle}({\sharp}^{*}\dot{f})={e_{k}}^{\langle{0}\rangle}\left(\frac{\partial f}{\partial x^{a}}Q^{a}_{i}y^{i}\right)=Q^{b}_{k}y^{i}\frac{\partial}{\partial x^{b}}(\frac{\partial f}{\partial x^{a}}Q^{a}_{i})+Q^{i}_{jk}Q^{a}_{i}y^{j}\frac{\partial f}{\partial x^{a}},

hence zμ\partial_{z^{\mu}} kills above expressions. By applying ek1{e_{k^{\prime}}}^{\langle{-1}\rangle} we get

ek1ek0(f˙)=Qbkxb(fxaQak)+QikkQaixa=ek(ek(f))+([ek,ek])(f),{e_{k^{\prime}}}^{\langle{-1}\rangle}{e_{k}}^{\langle{0}\rangle}({\sharp}^{*}\dot{f})=Q^{b}_{k}\frac{\partial}{\partial x^{b}}(\frac{\partial f}{\partial x^{a}}Q^{a}_{k^{\prime}})+Q^{i}_{{k^{\prime}}k}Q^{a}_{i}\partial_{x^{a}}=\sharp e_{k}(\sharp e_{k^{\prime}}(f))+(\sharp[e_{k^{\prime}},e_{k}])(f), (4.7)

and we are done. ∎

Proof of Lemma 3.11 and Theorem 3.13 part (a).

The formulas 12ε1(s1,s2)=β(s1,s2)([s1,s2])\frac{1}{2}\varepsilon_{1}(s_{1},s_{2})=\beta(s_{1},s_{2})-\partial([s_{1},s_{2}]) and 12ε0(s1,s2)=s1(s2)([s1,s2]){\frac{1}{2}}\varepsilon_{0}(s_{1},s_{2})=\Box_{s_{1}}(\partial s_{2})-\partial([s_{1},s_{2}]) come from the definitions of the corresponding maps, compare (3.35) with (3.16), (3.15) and (3.17). The other properties of the maps ε0\varepsilon_{0} and ε1\varepsilon_{1} given in Lemma 3.11 follow immediately from the properties of the maps β\beta and \Box given in Theorem 3.13, which we are going to prove first.

  • Proof of (3.13): As the Lie bracket of vector fields is skew symmetric, so β(s1,s2)=β(s2,s1)\beta(s_{1},s_{2})=-\beta(s_{2},s_{1}). There are no vector fields on E2E^{2} of weight less than 2-2. Hence, using (4.5), we get

    β(s1,fs2)=12[s11,fs21]+12[s11,(f˙)s22]=12f[s11,s21]+12s11(f˙)s22=fβ(s1,s2)+(s1)(f)(s2),{\beta}(s_{1},fs_{2})={\frac{1}{2}}[{s_{1}}^{\langle{-1}\rangle},f{s_{2}}^{\langle{-1}\rangle}]+{\frac{1}{2}}[{s_{1}}^{\langle{-1}\rangle},({\sharp}^{*}\dot{f}){s_{2}}^{\langle{-2}\rangle}]=\\ {\frac{1}{2}}f[{s_{1}}^{\langle{-1}\rangle},{s_{2}}^{\langle{-1}\rangle}]+{\frac{1}{2}}{s_{1}}^{\langle{-1}\rangle}({\sharp}^{*}\dot{f}){s_{2}}^{\langle{-2}\rangle}=f{\beta}(s_{1},s_{2})+(\sharp s_{1})(f)\partial(s_{2}), (4.8)

    by a weight argument and Lemma 4.4(iii).

  • Proof of (3.13) and (3.13): We expand (fs)0{(fs)}^{\langle{0}\rangle} as in (4.5) and find that for v𝔛2(E2)Γ(C)v\in\mathfrak{X}_{-2}(E^{2})\simeq\operatorname{\Gamma}(C) and sΓ(E)s\in\operatorname{\Gamma}(E) we have

    fsv\displaystyle\Box_{fs}v =[fs0,v]+[(f˙)s1,v]+[12(2)f¨s2,v]=\displaystyle=[f{s}^{\langle{0}\rangle},v]+{[({\sharp}^{*}\dot{f}){s}^{\langle{-1}\rangle},v]}+[{\frac{1}{2}}(\sharp^{2})^{\ast}\ddot{f}{s}^{\langle{-2}\rangle},v]=
    =f[s0,v]v(2f¨)s2=Lemma (4.4)f(sv)(Cv)(f)(s),\displaystyle=f[{s}^{\langle{0}\rangle},v]{-}v({\sharp^{2}}^{*}\ddot{f}){s}^{\langle{-2}\rangle}\stackrel{{\scriptstyle\text{Lemma \eqref{l:alifts_skew}}}}{{=}}f\cdot(\Box_{s}v)-(\sharp^{C}v)(f)\,\partial(s),

    as v(f)v(f), v((f˙))v({\sharp}^{*}(\dot{f})), [s1,v][{s}^{\langle{-1}\rangle},v], [s2,v][{s}^{\langle{-2}\rangle},v] vanish by inspecting weights. Similarly for (3.13):

    sfv=[s0,fv]=f[s0,v]+s0(f)v=f(sv)+(s)(f)v\Box_{s}fv=[{s}^{\langle{0}\rangle},fv]=f[{s}^{\langle{0}\rangle},v]{+}{s}^{\langle{0}\rangle}(f)v=f(\Box_{s}v){+}(\sharp s)(f)v

    by Lemma 4.4(iii).

  • Proof of the properties of the map ψ\psi given in Lemma 3.11:

    Let h:=12(2)f¨h:=\frac{1}{2}(\sharp^{2})^{\ast}\ddot{f} for time being, so hh is a function on E2E^{2} of weight two. From (4.5) and the definition (3.33) of ψ\psi we get

    (ψ(gs1,s2)gψ(s1,s2))(f)=12(g˙)s12s21(h)=0\left(\psi(gs_{1},s_{2})-g\psi(s_{1},s_{2})\right)(f)={\frac{1}{2}}({\sharp}^{*}\dot{g})\,{s_{1}}^{\langle{-2}\rangle}{s_{2}}^{\langle{-1}\rangle}(h)=0

    by a weight argument, hence ψ\psi is tensorial in its first argument. Also

    (ψ(s1,gs2)gψ(s1,s2))(f)=12s11(gs21(h)+g˙s22(h))12gs11s21(h)(s1g)(s2)(f)=12s11(g˙)s22(h)(s1)(g)(s2)(f)=(s1)(g)(C)(s2)(f),\left(\psi(s_{1},gs_{2})-g\psi(s_{1},s_{2})\right)(f)={\frac{1}{2}}{s_{1}}^{\langle{-1}\rangle}\left(g\,{s_{2}}^{\langle{-1}\rangle}(h)+{\sharp}^{*}\dot{g}\,{s_{2}}^{\langle{-2}\rangle}(h)\right)-{\frac{1}{2}}g{s_{1}}^{\langle{-1}\rangle}{s_{2}}^{\langle{-1}\rangle}(h)\\ -(\sharp s_{1}\,g)\,(\sharp s_{2})(f)={\frac{1}{2}}{s_{1}}^{\langle{-1}\rangle}({\sharp}^{*}\dot{g})\cdot{s_{2}}^{\langle{-2}\rangle}(h)-(\sharp s_{1})(g)\,(\sharp s_{2})(f)=(\sharp s_{1})(g)\left(\sharp^{C}\circ\partial-\sharp\right){(s_{2})}(f),

    by Lemma 4.4(ii) and (iii), hence we get (iii). Set A(f)=s11s21(h)A(f)={s_{1}}^{\langle{-1}\rangle}{s_{2}}^{\langle{-1}\rangle}(h), B(f)=(s1)(s2)(f)B(f)=(\sharp s_{1})\circ(\sharp s_{2})(f), so ψ(s1,s2)=AB\psi(s_{1},s_{2})=A-B. By inspecting weights and using (fg)(2)=fg¨+f¨g+2f˙g˙(fg)^{(2)}=f\ddot{g}+\ddot{f}g+2\dot{f}\dot{g} and Lemma 4.4(iii) we find that

    A(fg)=A(f)g+fA(g)+(s1)(f)(s2)(g)+(s1)(g)(s2)(f),A(fg)=A(f)g+fA(g)+(\sharp s_{1})(f)(\sharp s_{2})(g)+(\sharp s_{1})(g)(\sharp s_{2})(f),

    while

    B(fg)=fB(g)+B(f)g+(s1)(f)(s2)(g)+(s1)(g)(s2)(f)B(fg)=fB(g)+B(f)g+(\sharp s_{1})(f)(\sharp s_{2})(g)+(\sharp s_{1})(g)(\sharp s_{2})(f)

    hence ψ(s1,s2)\psi(s_{1},s_{2}) is a derivation. The coordinate formula (3.36) for ψ(ek,ek)\psi(e_{k^{\prime}},e_{k}) follows directly from (3.18) and (4.6):

    ek1ek1(h)\displaystyle{e_{k^{\prime}}}^{\langle{-1}\rangle}{e_{k}}^{\langle{-1}\rangle}(h) =ek1(12QμikQaμfxayi+12Qaikfxayi+2fxaxbQaiQbkyi)=\displaystyle={e_{k^{\prime}}}^{\langle{-1}\rangle}\left(\frac{1}{2}Q^{\mu}_{ik}Q^{a}_{\mu}\frac{\partial f}{\partial x^{a}}y^{i}+\frac{1}{2}Q^{a}_{ik}\frac{\partial f}{\partial x^{a}}y^{i}+\frac{\partial^{2}f}{\partial x^{a}\partial x^{b}}Q^{a}_{i}Q^{b}_{k}y^{i}\right)=
    =12(QμkkQaμ+Qakk)fxa+2fxaxbQakQbk,\displaystyle=\frac{1}{2}\left(Q^{\mu}_{k^{\prime}k}Q^{a}_{\mu}+Q^{a}_{k^{\prime}k}\right)\frac{\partial f}{\partial x^{a}}+\frac{\partial^{2}f}{\partial x^{a}\partial x^{b}}Q^{a}_{k^{\prime}}Q^{b}_{k},
    (ek)(ek)(f)\displaystyle(\sharp e_{k^{\prime}})(\sharp e_{k})(f) =QbkQak2fxaxb+QbkQakxbfxa.\displaystyle=Q^{b}_{k^{\prime}}Q^{a}_{k}\frac{\partial^{2}f}{\partial x^{a}\partial x^{b}}+Q^{b}_{k^{\prime}}\frac{\partial Q^{a}_{k}}{\partial x^{b}}\frac{\partial f}{\partial x^{a}}.

    The formula (3.38) for ψsym{\psi}^{\mathrm{sym}} follows immediately from (3.36). The skew-symmetric part of ψ\psi is derived from (3.33):

    ψalt(s1,s2)(f)=14[s11,s21]((2)f¨)12[s1,s2](f){\psi}^{\mathrm{alt}}(s_{1},s_{2})(f)={\frac{1}{4}}[{s_{1}}^{\langle{-1}\rangle},{s_{2}}^{\langle{-1}\rangle}]((\sharp^{2})^{\ast}\ddot{f})-{\frac{1}{2}}[\sharp s_{1},\sharp s_{2}](f)

    and this coincides with the formula (3.37) due to the definition of β\beta and Lemma 4.4(i). The direct computation of 2Θ2\sharp^{2}\circ\mathrm{\Theta}^{2} using (3.3) gives

    2Θ2(xa,yi,y˙i)=(xa,x˙a=Qaiyi,x¨a=12Qaijyiyj+Qaμ(Qμiy˙i+12Qμ(ij)yiyj))\sharp^{2}\circ\mathrm{\Theta}^{2}(x^{a},y^{i},\dot{y}^{i})=\left(x^{a},\dot{x}^{a}=Q^{a}_{i}y^{i},\ddot{x}^{a}=\frac{1}{2}Q^{a}_{ij}y^{i}y^{j}+Q^{a}_{\mu}\left(Q^{\mu}_{i}\dot{y}^{i}+\frac{1}{2}Q^{\mu}_{(ij)}y^{i}y^{j}\right)\right)

    and comparing it with

    [2](xa,yi,y˙i)=(xa,Qaiyi,12Qwidehataijyiyj+Qaiy˙i)\sharp^{[2]}(x^{a},y^{i},\dot{y}^{i})=(x^{a},Q^{a}_{i}y^{i},\frac{1}{2}\widehat{Q}^{a}_{ij}\,y^{i}y^{j}+Q^{a}_{i}\,\dot{y}^{i})

    as read from Example 3.8, gives the desired equivalence: 2Θ2=[2]\sharp^{2}\circ\mathrm{\Theta}^{2}=\sharp^{[2]} if and only if ψsym=0{\psi}^{\mathrm{sym}}=0 and =C\sharp=\sharp^{C}\circ\partial, see (3.7) and (3.38).

  • Proof of (3.13) and (3.13): The map δ\delta defined in (3.22) satisfies

    δs(fs1,s2)=fδs(s1,s2)+([s,s2])(f)(s1),\delta_{s}(fs_{1},s_{2})=f\delta_{s}(s_{1},s_{2})+(\sharp[s,s_{2}])(f)\,\partial(s_{1}), (Eq1δ\mathrm{Eq^{1}_{\delta}})
    δs(s1,fs2)=fδs(s1,s2)(s)(f)β(s1,s2)(s1)(f)s(s2)([s1,s]+ss1)(f)(s2),\delta_{s}(s_{1},fs_{2})=f\delta_{s}(s_{1},s_{2})-(\sharp s)(f)\beta(s_{1},s_{2}){-(\sharp s_{1})(f)\,\Box_{s}\partial(s_{2})-(\sharp[s_{1},s]+\sharp s\circ\sharp s_{1})(f)\,\partial(s_{2})}, (Eq2δ\mathrm{Eq^{2}_{\delta}})
    δfs(s1,s2)=fδs(s1,s2)+(s1)(f)β(s2,s)+(s2)(f)β(s1,s)+((s1)(s2)+ψ(s1,s2))(f)(s).\delta_{fs}(s_{1},s_{2})=f\delta_{s}(s_{1},s_{2})+(\sharp s_{1})(f)\beta(s_{2},s)+(\sharp s_{2})(f)\beta(s_{1},s)+\left((\sharp s_{1})\circ(\sharp s_{2})+\psi(s_{1},s_{2})\right)(f)\,\partial(s). (Eq3δ\mathrm{Eq^{3}_{\delta}})

    where δs(s1,s2)=δ(s1,s2,s)\delta_{s}(s_{1},s_{2})=\delta(s_{1},s_{2},s). We expand (fs)0{(f\,s)}^{\langle{0}\rangle} as in (4.5) and using Lemma 4.4 we get

    [s11,[s21,fs0]]=f[s11,[s21,s0]],\displaystyle[{s_{1}}^{\langle{-1}\rangle},[{s_{2}}^{\langle{-1}\rangle},f{s}^{\langle{0}\rangle}]]=f[{s_{1}}^{\langle{-1}\rangle},[{s_{2}}^{\langle{-1}\rangle},{s}^{\langle{0}\rangle}]],
    [s11,[s21,f˙s1]]=[s11,2f˙β(s2,s)+s2(f)s1]=2s1(f)β(s2,s)+2s2(f)β(s1,s),\displaystyle[{s_{1}}^{\langle{-1}\rangle},[{s_{2}}^{\langle{-1}\rangle},{\sharp}^{*}\dot{f}{s}^{\langle{-1}\rangle}]]=[{s_{1}}^{\langle{-1}\rangle},2{\sharp}^{*}\dot{f}\beta(s_{2},s)+\sharp s_{2}(f)\,{s}^{\langle{-1}\rangle}]=2\sharp s_{1}(f)\beta(s_{2},s)+2\sharp s_{2}(f)\beta(s_{1},s),
    [s11,[s21,(12(2)(f¨))s2]]=s11s21122(f¨)s2=(ψ(s1,s2)(f)+(s1)(s2)(f)) 2(s),\displaystyle[{s_{1}}^{\langle{-1}\rangle},[{s_{2}}^{\langle{-1}\rangle},(\frac{1}{2}{(\sharp^{2})}^{*}(\ddot{f})){s}^{\langle{-2}\rangle}]]={s_{1}}^{\langle{-1}\rangle}{s_{2}}^{\langle{-1}\rangle}\frac{1}{2}\sharp_{2}^{\ast}(\ddot{f}){s}^{\langle{-2}\rangle}=\left(\psi(s_{1},s_{2})(f)+(\sharp s_{1})(\sharp s_{2})(f)\right)\,{}2\partial(s),

    where ψ\psi is defined in (3.33). Summing up these three equalities, we get (Eq3δ\mathrm{Eq^{3}_{\delta}}). The equalities (Eq1δ\mathrm{Eq^{1}_{\delta}}) and (Eq2δ\mathrm{Eq^{2}_{\delta}}) can be derived in a very similar way and we omit the proof. The direct use of the definition of ω\omega (see (3.25)) and the properties of δ\delta and β\beta lead to

    ωs(fs1,s2)\displaystyle\omega_{s}(fs_{1},s_{2}) =fωs(s1,s2),\displaystyle=f\omega_{s}(s_{1},s_{2}),
    ωs(s1,fs2)\displaystyle\omega_{s}(s_{1},fs_{2}) fωs(s1,s2)=(s1)(f)(([s,s2]s(s2)))+([s,s1][s,s1])(f)(s2)=\displaystyle-f\omega_{s}(s_{1},s_{2})=(\sharp s_{1})(f)\left(\partial([s,s_{2}]-\Box_{s}\partial(s_{2}))\right)+\left(\sharp[s,s_{1}]-[\sharp s,\sharp s_{1}]\right)(f)\partial(s_{2})=
    (s1)(f)ε0(s,s2)+ξ(s,s1)s2,\displaystyle-(\sharp s_{1})(f)\varepsilon_{0}(s,s_{2})+\xi(s,s_{1})\partial s_{2},
    ωfs(s1,s2)\displaystyle\omega_{fs}(s_{1},s_{2}) fωs(s1,s2)=(s1)(f)(([s,s2])β(s,s2))+ψ(s1,s2)(f)(s)=\displaystyle-f\omega_{s}(s_{1},s_{2})=(\sharp s_{1})(f)\left(\partial([s,s_{2}])-\beta(s,s_{2})\right)+\psi(s_{1},s_{2})(f)\,\partial(s)=
    12(s1)(f)ε1(s2,s)+ψ(s1,s2)(f)(s),\displaystyle\frac{1}{2}(\sharp s_{1})(f)\varepsilon_{1}(s_{2},s)+\psi(s_{1},s_{2})(f)\,\partial(s),

    where ωs(s1,s2)=ω(s1,s2,s)\omega_{s}(s_{1},s_{2})=\omega(s_{1},s_{2},s) and ε0\varepsilon_{0}, ε1\varepsilon_{1}, ξ\xi and ψ\psi are as in Definition 3.10. From this, the equations (3.13) and (3.13) follow immediately.

Proof of Theorem 3.15.

First, we shall describe the structure of the Lie algebra 𝔛0(𝔤×C)\mathfrak{X}_{\leq 0}(\mathfrak{g}\times C) with respect to the decomposition given in (3.45). We write ϕψχ𝔛0\phi\oplus\psi\oplus\chi\in\mathfrak{X}_{0}, xf𝔛1x\oplus f\in\mathfrak{X}_{-1}, v𝔛2v\in\mathfrak{X}_{-2}, where ϕEnd(𝔤)\phi\in\operatorname{End}(\mathfrak{g}), ψEnd(C)\psi\in\operatorname{End}(C), χHom(Sym2𝔤,C)\chi\in\operatorname{Hom}(\operatorname{Sym}^{2}\mathfrak{g},C), x𝔤x\in\mathfrak{g}, fHom(𝔤,C)f\in\operatorname{Hom}(\mathfrak{g},C), and vCv\in C. We have

[𝔛0,𝔛0]:[ϕ1,ϕ2]=ϕ2ϕ1ϕ1ϕ2,[ψ1,ψ2]=ψ2ψ1ψ1ψ2,[χ1,χ2]=0[ϕ,χ](x1,x2)=χ(ϕ(x1),x2)+χ(x1,ϕ(x2)),[ψ,χ](x1,x2)=ψ(χ(x1,x2)),[ϕ,ψ]=0,[𝔛0,𝔛1]:[ϕ,x]=ϕ(x),[ϕ,f]=fϕ,[ψ,x]=0,[ψ,f]=ψf,[χ,x]=χ(x,),[χ,f]=0[𝔛1,𝔛1]:[x1,x2]=0,[f1,f2]=0,[f,x]=f(x),[𝔛0,𝔛2]:[ϕ,v]=0,[ψ,v]=ψ(v),[χ,v]=0.\begin{split}[\mathfrak{X}_{0},\mathfrak{X}_{0}]:&[\phi_{1},\phi_{2}]=\phi_{2}\circ\phi_{1}-\phi_{1}\circ\phi_{2},[\psi_{1},\psi_{2}]=\psi_{2}\circ\psi_{1}-\psi_{1}\circ\psi_{2},[\chi_{1},\chi_{2}]=0\\ &[\phi,\chi](x_{1},x_{2})=\chi(\phi(x_{1}),x_{2})+\chi(x_{1},\phi(x_{2})),[\psi,\chi](x_{1},x_{2})=-\psi(\chi(x_{1},x_{2})),[\phi,\psi]=0,\\ [\mathfrak{X}_{0},\mathfrak{X}_{-1}]:&[\phi,x]=-\phi(x),[\phi,f]=f\circ\phi,[\psi,x]=0,[\psi,f]=-\psi\circ f,[\chi,x]=-\chi(x,\cdot),[\chi,f]=0\\ [\mathfrak{X}_{-1},\mathfrak{X}_{-1}]:&[x_{1},x_{2}]=0,[f_{1},f_{2}]=0,[f,x]=-f(x),\\ [\mathfrak{X}_{0},\mathfrak{X}_{-2}]:&[\phi,v]=0,[\psi,v]=-\psi(v),[\chi,v]=0.\end{split} (4.9)

For example, the formula for [ϕ,x][\phi,x] can be derived as follows. A vector x=(xi)𝔤x=(x^{i})\in\mathfrak{g} is idenified with the vector field xiyix^{i}\partial_{y^{i}}, and an endomorphism ϕ:𝔤𝔤\phi:\mathfrak{g}\to\mathfrak{g}, such that ϕ(yi)=ϕijyj{\phi}^{*}(y^{i})=\phi^{i}_{j}y^{j}, is identified with the vector field ϕ=ϕijyjyi\phi=\phi^{i}_{j}y^{j}\partial_{y^{i}}. Then, [ϕ,x]=ϕijxjyi=ϕ(x)[\phi,x]=-\phi^{i}_{j}x^{j}\partial_{y^{i}}=-\phi(x). In a similar way we derive the remaining formulas.

The formulas for algebroid lifts eα{e}^{\langle{{\alpha}}\rangle}, where α=2,1,0{\alpha}=-2,-1,0, given in the formulation of our theorem, define vector fields which have the form as in (3.43) since the projection of e1{e}^{\langle{-1}\rangle} onto 𝔤\mathfrak{g} is ee and the bracket [,][\cdot,\cdot] is skew-symmetric. Therefore, these vector fields define an AL higher algebroid.

Conversely, let (𝔤×C,κ2)(\mathfrak{g}\times C,\kappa^{2}) be a skew HA defined by means of algebroid lifts eα{e}^{\langle{{\alpha}}\rangle} given above. Let us temporarily denote by ~\widetilde{\partial}, β~\widetilde{\beta}, ~\widetilde{\Box}, ωsym~\widetilde{{\omega}^{\mathrm{sym}}} the maps associated with κ2\kappa^{2}, defined in Subsection 3.2 by formulas (3.15), (3.16), (3.17), (3.26), respectively. We shall show that ~=\widetilde{\partial}=\partial, β~=β\widetilde{\beta}=\beta, ~=\widetilde{\Box}=\Box, ωsym~=ωsym\widetilde{{\omega}^{\mathrm{sym}}}={\omega}^{\mathrm{sym}}.

The definitions of ~\widetilde{\partial} and \partial coincide, hence ~=\widetilde{\partial}=\partial. For the proof of the equality β~=β\widetilde{\beta}=\beta we have

β~(x,y)=12[x1,y1]=12[xβ(,x),yβ(,y)]=β(x,y)\widetilde{\beta}(x,y)=\frac{1}{2}[{x}^{\langle{-1}\rangle},{y}^{\langle{-1}\rangle}]=\frac{1}{2}[x\oplus\beta(\cdot,x),y\oplus\beta(\cdot,y)]=\beta(x,y)

due to the skew-symmetry of β\beta and the formulas (4.9) for the bracket on 𝔛1𝔛1\mathfrak{X}_{-1}\oplus\mathfrak{X}_{-1}. For ~=\widetilde{\Box}=\Box we write

~xv=[x0,v]=[[,x]ϕx()ψ2ωsymx(,)χ,v]=ψ(v)=xv,{\widetilde{\Box}_{x}v=[{x}^{\langle{0}\rangle},v]=[\underbrace{[\cdot,x]}_{\phi}\oplus\underbrace{\Box_{-x}(\cdot)}_{\psi}\oplus\underbrace{2{\omega}^{\mathrm{sym}}_{x}(\cdot,\cdot)}_{\chi},v]=-\psi(v)=\Box_{x}v},

due to the formulas for the bracket restricted to 𝔛2𝔛0\mathfrak{X}_{-2}\oplus\mathfrak{X}_{0}. The proof of δ~=δ\widetilde{\delta}=\delta is a bit longer. First, we calculate

[y1,z0]=[yβ(,y),[,z]z()2ωsymz(,)]=[y,z]Xβ([,z],y)+zβ(,y)+2ωsymz(y,)F,\begin{split}[{y}^{\langle{-1}\rangle},{z}^{\langle{0}\rangle}]&=[y\oplus\beta(\cdot,y),[\cdot,z]\oplus\Box_{-z}(\cdot)\oplus 2{\omega}^{\mathrm{sym}}_{z}(\cdot,\cdot)]=\\ &\underbrace{[y,z]}_{X}\oplus\underbrace{\beta([\cdot,-z],y)+{\Box_{-z}\beta(\cdot,y)}+{2{\omega}^{\mathrm{sym}}_{z}(y,\cdot)}}_{F},\end{split} (4.10)

hence

δ~(x,y,z)\displaystyle\widetilde{\delta}(x,y,z) =12[x1,[y1,z0]]=12[xβ(,x),XF]=12F(x)12β(X,x)=\displaystyle={\frac{1}{2}}[{x}^{\langle{-1}\rangle},[{y}^{\langle{-1}\rangle},{z}^{\langle{0}\rangle}]]=\frac{1}{2}[x\oplus\beta(\cdot,x),X\oplus F]=\frac{1}{2}F(x)-\frac{1}{2}\beta(X,x)=
12β([y,z],x)+12β([z,x],y)12zβ(x,y)+ωsymz(y,x).\displaystyle-\frac{1}{2}\beta([y,z],x)+\frac{1}{2}\beta([z,x],y)-\frac{1}{2}\Box_{z}{\beta(x,y)}+{\omega}^{\mathrm{sym}}_{z}(y,x).

Thus ωsymx~(y,z)\widetilde{{\omega}^{\mathrm{sym}}_{x}}(y,z), which is obtained by symmetrizing δ~(x,y,z)β(x,[y,z])\widetilde{\delta}(x,y,z)-\beta(x,[y,z]) in x,yx,y, coincides, due to the skew-symmetry of β\beta, with the symmetrization of ωsymz(x,y)+12β([y,z],x)+12β([z,x],y){\omega}^{\mathrm{sym}}_{z}(x,y)+\frac{1}{2}\beta([y,z],x)+\frac{1}{2}\beta([z,x],y), and the latter simplifies to ωsymz(x,y){\omega}^{\mathrm{sym}}_{z}(x,y), as was claimed.

We now examine the Lie condition for HAs given in Remark 2.12. From (4.9) we get

[x0,y0]=[[x,]ϕ1x()ψ12ωsymx(,)χ1,[y,]ϕ2y()ψ22ωsymy(,)χ2]=([y,[x,]][x,[y,]])[y,x]χ,[{x}^{\langle{0}\rangle},{y}^{\langle{0}\rangle}]=[\underbrace{[-x,\cdot]}_{\phi_{1}}\oplus\underbrace{\Box_{-x}(\cdot)}_{\psi_{1}}\oplus\underbrace{2{\omega}^{\mathrm{sym}}_{x}(\cdot,\cdot)}_{\chi_{1}},\underbrace{[-y,\cdot]}_{\phi_{2}}\oplus\underbrace{\Box_{-y}(\cdot)}_{\psi_{2}}\oplus\underbrace{2{\omega}^{\mathrm{sym}}_{y}(\cdot,\cdot)}_{\chi_{2}}]=\\ \left([y,[x,\cdot]]-[x,[y,\cdot]]\right)\oplus[\Box_{y},\Box_{x}]\oplus\chi,

for some χHom(Sym2𝔤,C)\chi\in\operatorname{Hom}(\operatorname{Sym}^{2}\mathfrak{g},C). From the condition [x0,y0]=[x,y]0[{x}^{\langle{0}\rangle},{y}^{\langle{0}\rangle}]={[x,y]}^{\langle{0}\rangle} we read that 𝔤\mathfrak{g} is a Lie algebra, CC is a left 𝔤\mathfrak{g}-module. We shall show that ωsym=0{\omega}^{\mathrm{sym}}=0, from which it follows that the identity obtain by comparing the Hom(Sym2𝔤,C)\operatorname{Hom}(\operatorname{Sym}^{2}\mathfrak{g},C)-components is satisfied automatically.

The equation [x1,y1]=[x,y]2[{x}^{\langle{-1}\rangle},{y}^{\langle{-1}\rangle}]={[x,y]}^{\langle{-2}\rangle} and [x0,y2]=[x,y]2[{x}^{\langle{0}\rangle},{y}^{\langle{-2}\rangle}]={[x,y]}^{\langle{-2}\rangle} write as

x(y)=([x,y]) and β(x,y)=([x,y]).\Box_{x}\partial(y)=\partial([x,y])\text{ and }\beta(x,y)=\partial([x,y]). (4.11)

Finally, for (i,j)=(0,1)(i,j)=(0,-1), the Lie condition from (4.10) is given by

β(,[x,y])=β([,x],y)+xβ(,y)+ωsymx(y,).\beta(\cdot,[x,y])=\beta([\cdot,-x],y)+\Box_{-x}\beta(\cdot,y)+{\omega}^{\mathrm{sym}}_{x}(y,\cdot).

This, along with the equalities in (4.11) and the Jacobi identity, yields ωsym=0{\omega}^{\mathrm{sym}}=0, and completes the proof. ∎

4.3 Equations for AL and Lie HAs

We shall use Theorem 2.11 to write equations for structure functions corresponding to almost Lie HAs. The obtained equations will be used to complete the proof of Theorem 3.16.

AL HAs.

Let (E2,κ2)(E^{2},\kappa^{2}) be an AL HA and let (ek)(e_{k}), (cμ)(c_{\mu}) be as in Subsection 3.2. The vector fields ekα𝔛α(E2){e_{k}}^{\langle{{\alpha}}\rangle}\in\mathfrak{X}_{\alpha}(E^{2}) for α=0,1,2{\alpha}=0,-1,-2 are given in (3.18). The formulas for (T2M,κ2M)(\mathrm{T}^{2}M,\kappa^{2}_{M})-algebroid lifts e¯kα:=(ek)α=(Qakxa)α𝔛α(T2M){\underline{e}_{k}}^{\langle{{\alpha}}\rangle}:={(\sharp e_{k})}^{\langle{{\alpha}}\rangle}={(Q^{a}_{k}\partial_{x^{a}})}^{\langle{{\alpha}}\rangle}\in\mathfrak{X}_{\alpha}(\mathrm{T}^{2}M) are easily derived from (2.31) by notting that xaα{\partial_{x^{a}}}^{\langle{{\alpha}}\rangle} is equal to xa\partial_{x^{a}}, x˙a\partial_{\dot{x}^{a}} and 2x¨a2\partial_{\ddot{x}^{a}} for α=0,1,2{\alpha}=0,-1,-2, respectively. Thus

{e¯k0=Qakxa+Qakxbx˙bx˙a+(Qakxbx¨b+2Qakxbxcx˙bx˙c)x¨a,e¯k1=Qakx˙a+2Qakxbx˙bx¨a,e¯k2=2Qakx¨a.\begin{cases}{\underline{e}_{k}}^{\langle{0}\rangle}&=Q^{a}_{k}\partial_{x^{a}}+\frac{\partial Q^{a}_{k}}{\partial{x^{b}}}\dot{x}^{b}\partial_{\dot{x}^{a}}+\left(\frac{\partial Q^{a}_{k}}{\partial x^{b}}\ddot{x}^{b}+\frac{\partial^{2}Q^{a}_{k}}{\partial{x^{b}}\partial{x^{c}}}\dot{x}^{b}\dot{x}^{c}\right)\partial_{\ddot{x}^{a}},\\ {\underline{e}_{k}}^{\langle{-1}\rangle}&=Q^{a}_{k}\partial_{\dot{x}^{a}}+{2}\frac{\partial Q^{a}_{k}}{\partial x^{b}}\dot{x}^{b}\partial_{\ddot{x}^{a}},\\ {\underline{e}_{k}}^{\langle{-2}\rangle}&=2Q^{a}_{k}\,\partial_{\ddot{x}^{a}}.\end{cases} (4.12)

(Note that (Qak)(1)=Qakxbx˙b(Q^{a}_{k})^{(1)}=\frac{\partial Q^{a}_{k}}{\partial{x^{b}}}\dot{x}^{b} and (Qak)(2)=Qakxbx¨b+2Qakxbxcx˙bx˙c(Q^{a}_{k})^{(2)}=\frac{\partial Q^{a}_{k}}{\partial x^{b}}\ddot{x}^{b}+\frac{\partial^{2}Q^{a}_{k}}{\partial{x^{b}}\partial{x^{c}}}\dot{x}^{b}\dot{x}^{c}. The above formulas for e¯kα{\underline{e}_{k}}^{\langle{{\alpha}}\rangle} can also be obtained from (3.18) and (3.29).) We check whether vector fields ekα{e_{k}}^{\langle{{\alpha}}\rangle} and e¯kα{\underline{e}_{k}}^{\langle{{\alpha}}\rangle} are 2\sharp^{2}-related, where 2(xa,yi,zμ)=(xa,x˙a=Qakyk,x¨a=Qaμzμ+12Qaijyiyj)\sharp^{2}(x^{a},y^{i},z^{\mu})=(x^{a},\dot{x}^{a}=Q^{a}_{k}y^{k},\ddot{x}^{a}=Q^{a}_{\mu}z^{\mu}+\frac{1}{2}Q^{a}_{ij}y^{i}y^{j}). Straightforward calculations leads to the following system of equations (referred to as AL HA equations):

QaμQμk=Qak\displaystyle Q^{a}_{\mu}Q^{\mu}_{k}=Q^{a}_{k} (4.13a)
Qakj+QaμQμjk=2QakxbQbj\displaystyle Q^{a}_{kj}+Q^{a}_{\mu}Q^{\mu}_{jk}=2\,\frac{\partial Q^{a}_{k}}{\partial x^{b}}Q^{b}_{j} (4.13b)
QakQijk=Qwidecheckaij\displaystyle Q^{a}_{k}Q_{ij}^{k}=\widecheck{Q}^{a}_{ij} (4.13c)
QaνQνμi+QaμxbQbi=QaixbQbμ\displaystyle Q^{a}_{\nu}Q^{\nu}_{\mu i}+\frac{\partial Q^{a}_{\mu}}{\partial x^{b}}Q^{b}_{i}=\frac{\partial Q^{a}_{i}}{\partial x^{b}}Q^{b}_{\mu} (4.13d)
QaμQμij,k+QaijxbQbk+QaliQljk+QaljQlik=QakxbQbij+22QakxbxcQbiQcj\displaystyle Q^{a}_{\mu}Q^{\mu}_{ij,k}+\frac{\partial Q^{a}_{ij}}{\partial x^{b}}Q^{b}_{k}+Q^{a}_{li}Q^{l}_{jk}+Q^{a}_{lj}Q^{l}_{ik}=\frac{\partial Q^{a}_{k}}{\partial x^{b}}Q^{b}_{ij}+2\,\frac{\partial^{2}Q^{a}_{k}}{\partial x^{b}\partial x^{c}}Q^{b}_{i}Q^{c}_{j} (4.13e)

where

Qwidecheckaij:=QbiQajxbQbjQaixb,\widecheck{Q}^{a}_{ij}:=Q^{b}_{i}\frac{\partial Q^{a}_{j}}{\partial x^{b}}-Q^{b}_{j}\frac{\partial Q^{a}_{i}}{\partial x^{b}}, (4.14)

(The equations (4.13a), (4.13b) correspond to the cases α=2{\alpha}=-2 and α=1{\alpha}=-1, respectively; while (4.13c) (4.13d) and (4.13e) correspond to the case α=0{\alpha}=0. Note also that the equations (4.13a), (4.13b), (4.13c), and (4.13d) follows immadiately from (AL\mathrm{AL_{\partial}}), (ALψ\mathrm{AL_{\psi}}), (ALA\mathrm{AL_{A}}), and (AL\mathrm{AL_{\Box}}), respectively.) Note that

[ei,ej]=Qwidecheckaijxa.[\sharp e_{i},\sharp e_{j}]=\widecheck{Q}^{a}_{ij}\partial_{x^{a}}. (4.15)

The equation (4.13b) can be replaced with

QaμQ[ij]=Qwidecheckaij,\displaystyle Q^{a}_{\mu}Q_{[ij]}=\widecheck{Q}^{a}_{ij}, (4.16a)
QaμQμ(ij)+Qaij=Qwidehataij\displaystyle Q^{a}_{\mu}Q^{\mu}_{(ij)}+Q^{a}_{ij}=\widehat{Q}^{a}_{ij} (4.16b)

where Qwidehataij\widehat{Q}^{a}_{ij} is given in (3.7).

Completion of the proof of Theorem 3.16.

We shall prove that if a skew HA (E2,κ2)(E^{2},\kappa^{2}) satisfy the conditions listed in Theorem 3.16, then it is almost Lie. It amounts to proving that the vector fields ek[α]{e_{k}}^{[{\alpha}]} and e¯k[α]{\underline{e}_{k}}^{[{\alpha}]}, see (3.18) and (4.12), are 2\sharp^{2}-related for α=2,1,0{\alpha}=-2,-1,0. We have already proved this for α=2,1{\alpha}=-2,-1, so it remains to prove this for α=0{\alpha}=0, i.e., to verify the equations (4.13c,4.13d 4.13e).

The equation (4.13c) means [ei,ej]=Qwidecheckaijxa\sharp[e_{i},e_{j}]=\widecheck{Q}^{a}_{ij}\partial_{x^{a}}, which is true since the algebroid (E1,κ1)(E^{1},\kappa^{1}) is AL. Next, the condition (4.13d) means Ceicμ=[ei,Ccμ]\sharp^{C}\Box_{e_{i}}c_{\mu}=[\sharp e_{i},\sharp^{C}c_{\mu}], and it follows from (AL\mathrm{AL_{\Box}}).

The proof of (4.13e) is a bit more involved. We claim that

Cδsym=δsymT2M×3:A×A×ATM,\sharp^{C}\circ{\delta}^{\mathrm{sym}}={\delta}^{\mathrm{sym}}_{\mathrm{T}^{2}M}\circ\sharp^{\times 3}:A\times A\times A\rightarrow\mathrm{T}M, (4.17)

where δsym{\delta}^{\mathrm{sym}} is given in (3.23) and δsymT2M{\delta}^{\mathrm{sym}}_{\mathrm{T}^{2}M} is the same structure map but associated with the HA (T2M,κ2M)(\mathrm{T}^{2}M,\kappa^{2}_{M}). Indeed, Cω=0\sharp^{C}\circ\omega=0 implies Cωsym=0\sharp^{C}\circ{\omega}^{\mathrm{sym}}=0 and from δsym=ωsym+12(β(s1,[s2,s])+β(s2,[s1,s])){\delta}^{\mathrm{sym}}={\omega}^{\mathrm{sym}}+\frac{1}{2}(\beta(s_{1},[s_{2},s])+\beta(s_{2},[s_{1},s])) we find that

Cδsyms(s1,s2)=12C(β(s1,[s2,s])+β(s2,[s1,s]))=(3.2.4)([[s,s1],s2]+[[s,s2],s1])=(δsymT2M)s(s1,s2).\sharp^{C}\circ{\delta}^{\mathrm{sym}}_{s}(s_{1},s_{2})=\frac{1}{2}\sharp^{C}\circ\left(\beta(s_{1},[s_{2},s])+\beta(s_{2},[s_{1},s])\right)\stackrel{{\scriptstyle\eqref{i:AL_beta}}}{{=}}\sharp([[s,s_{1}],s_{2}]+[[s,s_{2}],s_{1}])=\left({\delta}^{\mathrm{sym}}_{\mathrm{T}^{2}M}\right)_{\sharp s}(\sharp s_{1},\sharp s_{2}).

We shall show that (4.17) gives (4.13e). We shall work with an adapted coordinate system (xa,yi,wμ)(x^{a},y^{i},w^{\mu}) for (E2,κ2)(E^{2},\kappa^{2}) (see Definition 3.7), so Qμ(ij)=0Q^{\mu}_{(ij)}=0. The general idea is to express δsymek(ei,ej){\delta}^{\mathrm{sym}}_{e_{k}}(e_{i},e_{j}) entirely in terms of the structure functions QaiQ^{a}_{i} and its derivatives and then compare with (δsymT2M)ek(ei,ej)\left({\delta}^{\mathrm{sym}}_{\mathrm{T}^{2}M}\right)_{\sharp e_{k}}(\sharp e_{i},\sharp e_{j}), which is easily seen to be of this form. From the expression for Q~μijk\tilde{Q}^{\mu}_{ijk} in (3.21) we find that

δsymek(ei,ej)=12(Qμij,kQlikQμljQljkQμli)cμ,{\delta}^{\mathrm{sym}}_{e_{k}}(e_{i},e_{j})=\frac{1}{2}\left(Q^{\mu}_{ij,k}-Q^{l}_{ik}Q^{\mu}_{lj}-Q^{l}_{jk}Q^{\mu}_{li}\right)c_{\mu},

hence

2Cδsymek(ei,ej)=Qaμ(Qμij,kQlikQμljQljkQμli)xa.2\cdot\sharp^{C}\circ{\delta}^{\mathrm{sym}}_{e_{k}}(e_{i},e_{j})=Q^{a}_{\mu}\left(Q^{\mu}_{ij,k}-Q^{l}_{ik}Q^{\mu}_{lj}-Q^{l}_{jk}Q^{\mu}_{li}\right)\partial_{x^{a}}.

We replace QaμQljkQμliQ^{a}_{\mu}Q^{l}_{jk}Q^{\mu}_{li} with

QaμQljkQμli=(4.13b)Qljk(Qali+2QblQaixb)=(4.13c)QljkQali+2QwidecheckbjkQaixbQ^{a}_{\mu}Q^{l}_{jk}Q^{\mu}_{li}\stackrel{{\scriptstyle\eqref{e:Qa_kj}}}{{=}}Q^{l}_{jk}(-Q^{a}_{li}+2Q^{b}_{l}\frac{\partial Q^{a}_{i}}{\partial x^{b}})\stackrel{{\scriptstyle\eqref{e:QakQk_ij}}}{{=}}-Q^{l}_{jk}Q^{a}_{li}+2\widecheck{Q}^{b}_{jk}\frac{\partial Q^{a}_{i}}{\partial x^{b}}

and similarly for QaμQlikQμljQ^{a}_{\mu}Q^{l}_{ik}Q^{\mu}_{lj} and get

2Cδsymek(ei,ej)=(QaμQμij,k+QljkQali+QlikQalj2(QwidecheckbjkQaixb+QwidecheckbikQajxb))xa2\cdot\sharp^{C}\circ{\delta}^{\mathrm{sym}}_{e_{k}}(e_{i},e_{j})=(Q^{a}_{\mu}Q^{\mu}_{ij,k}+Q^{l}_{jk}Q^{a}_{li}+Q^{l}_{ik}Q^{a}_{lj}-2(\widecheck{Q}^{b}_{jk}\frac{\partial Q^{a}_{i}}{\partial x^{b}}+\widecheck{Q}^{b}_{ik}\frac{\partial Q^{a}_{j}}{\partial x^{b}}))\partial_{x^{a}} (4.18)

As Qaij=QwidehataijQ^{a}_{ij}=\widehat{Q}^{a}_{ij} by (4.16b) the condition (4.13e) can be equivalently written as

2Cδsymek(ei,ej)=(22QakxbxcQbiQcj+QwidehatbjiQakxbQbkQwidehataijxb2(QwidecheckbjkQaixb+QwidecheckbikQajxb))xa,2\cdot\sharp^{C}\circ{\delta}^{\mathrm{sym}}_{e_{k}}(e_{i},e_{j})=\left(2\frac{\partial^{2}Q^{a}_{k}}{\partial x^{b}\partial x^{c}}Q^{b}_{i}Q^{c}_{j}+\widehat{Q}^{b}_{ji}\frac{\partial Q^{a}_{k}}{\partial x^{b}}-Q^{b}_{k}\frac{\partial\widehat{Q}^{a}_{ij}}{\partial x^{b}}-2(\widecheck{Q}^{b}_{jk}\frac{\partial Q^{a}_{i}}{\partial x^{b}}+\widecheck{Q}^{b}_{ik}\frac{\partial Q^{a}_{j}}{\partial x^{b}})\right)\partial_{x^{a}},

It remains to show that the last expressions coincides with [ei,[ej,ek]]+[ej,[ei,ek]][\sharp e_{i},[\sharp e_{j},\sharp e_{k}]]+[\sharp e_{j},[\sharp e_{i},\sharp e_{k}]]. This a direct calculation of the brackets of vector fields. Namely, from (4.15) we get

[ei,[ej,ek]]=(QbiQwidecheckajkxbQwidecheckbjkQaixb)xa.[\sharp e_{i},[\sharp e_{j},\sharp e_{k}]]=\left(Q^{b}_{i}\frac{\partial\widecheck{Q}^{a}_{jk}}{\partial x^{b}}-\widecheck{Q}^{b}_{jk}\frac{\partial Q^{a}_{i}}{\partial x^{b}}\right)\partial_{x^{a}}.

On the other hand, the following identity holds

22QakxbxcQbiQcj+QwidehatbjiQakxbQbkQwidehataijxb=QbiQwidecheckajkxb+QwidecheckbjkQaixb+QbjQwidecheckaikxb+QwidecheckbikQajxb2\frac{\partial^{2}Q^{a}_{k}}{\partial x^{b}\partial x^{c}}Q^{b}_{i}Q^{c}_{j}+\widehat{Q}^{b}_{ji}\frac{\partial Q^{a}_{k}}{\partial x^{b}}-Q^{b}_{k}\frac{\partial\widehat{Q}^{a}_{ij}}{\partial x^{b}}=Q^{b}_{i}\frac{\partial\widecheck{Q}^{a}_{jk}}{\partial x^{b}}+\widecheck{Q}^{b}_{jk}\frac{\partial Q^{a}_{i}}{\partial x^{b}}+Q^{b}_{j}\frac{\partial\widecheck{Q}^{a}_{ik}}{\partial x^{b}}+\widecheck{Q}^{b}_{ik}\frac{\partial Q^{a}_{j}}{\partial x^{b}}

which can be verified by expanding Qwidehataij\widehat{Q}^{a}_{ij} and Qwidecheckajk\widecheck{Q}^{a}_{jk} using (3.7) and (4.14), and then grouping and cancelling similar terms. On the LHS is the part of the expression (4.18) involving second derivatives. After plugging the RHS to (4.18) we shall easily recognize the desired formula. ∎

Lie HAs.

For completeness we provide a system of equations ensuring that a given AL HA is Lie. They are obtained by examining the conditions listed in Theorem 3.20 on local frames (ei)(e_{i}) and (cμ)(c_{\mu}) of the VBs AA and CC, respectively, see Remark 3.22. The can be also obtained by examining the condition given in Remark 2.12.

cyclic i,j,kQijlQlkm=0,\displaystyle\sum_{\text{cyclic }i,j,k}Q_{ij}^{l}Q_{lk}^{m}=0, (4.19a)
QμνkQνi=QμlQlik+QakQμixa,\displaystyle Q^{\mu}_{\nu k}Q^{\nu}_{i}=Q^{\mu}_{l}Q^{l}_{ik}+Q^{a}_{k}\frac{\partial Q^{\mu}_{i}}{\partial x^{a}}, (4.19b)
Qμ[ik]=QμjQjik,\displaystyle Q^{\mu}_{[ik]}=Q^{\mu}_{j}Q^{j}_{ik}, (4.19c)
Qμii,k=QjikQμjiQνiiQμνkQjkiQμij+QakQμiixa,\displaystyle Q^{\mu}_{{i^{\prime}}i,k}=Q^{j}_{ik}Q^{\mu}_{j{i^{\prime}}}-Q^{\nu}_{i{i^{\prime}}}Q^{\mu}_{\nu k}-Q^{j}_{k{i^{\prime}}}Q^{\mu}_{ij}+Q^{a}_{k}\frac{\partial Q^{\mu}_{i{i^{\prime}}}}{\partial x^{a}}, (4.19d)
QjkkQμνj+QjkkxaQaνQμj=QakQμνkxa+QρνkQμρkQakQμνkxa+QρνkQμρk\displaystyle Q^{j}_{k{k^{\prime}}}Q^{\mu}_{\nu j}+\frac{\partial Q^{j}_{k{k^{\prime}}}}{\partial x^{a}}Q^{a}_{\nu}Q^{\mu}_{j}=Q^{a}_{k}\frac{\partial Q^{\mu}_{\nu{k^{\prime}}}}{\partial x^{a}}+Q^{\rho}_{\nu k}Q^{\mu}_{\rho{k^{\prime}}}-Q^{a}_{{k^{\prime}}}\frac{\partial Q^{\mu}_{\nu k}}{\partial x^{a}}+Q^{\rho}_{\nu{k^{\prime}}}Q^{\mu}_{\rho k} (4.19e)

The equation (4.19a) corresponds to the Jacobi identity, while (4.19b), (4.19c), (4.19d), and (4.19e) correspond to (Lie\mathrm{Lie_{\partial}}), (Lieβ\mathrm{Lie_{\beta}}), (Lieω\mathrm{Lie_{\omega}}), and (Lie\mathrm{Lie_{\Box}}), respectively.

Proof of Conjecture 3.5 in the case k=2k=2.

Let (ei)(e_{i}) be a local basis of sections of σ:AM\sigma:A\rightarrow M and denote Ei[α]:=eiακ[2]E_{i}^{[{\alpha}]}:=e_{i}^{\langle{\alpha}\rangle_{\kappa^{[2]}}} – the (A[2],κ[2])(A^{[2]},\kappa^{[2]})–algebroid lifts, α=0,1,2\alpha=0,-1,2. Using (3.18) and Example 3.8 we find that

{Ek[0]=Qakxa+Qjkiyjyi+(Qmnky˙n+12Qwidehatij,kmyiyj)y˙m,Ek[1]=yk+Qikmyiy˙m,Ek[2]=2y˙k\begin{cases}E_{k}^{[0]}&=Q^{a}_{k}\partial_{x^{a}}+Q_{jk}^{i}y^{j}\partial_{y^{i}}+(Q^{m}_{nk}\,\dot{y}^{n}+\frac{1}{2}\,\widehat{Q}_{ij,k}^{m}\,y^{i}y^{j})\,\partial_{{\dot{y}}^{m}},\\ E_{k}^{[-1]}&=\partial_{y^{k}}+Q_{ik}^{m}\,y^{i}\,\partial_{{\dot{y}}^{m}},\\ E_{k}^{[-2]}&=2\partial_{{\dot{y}}^{k}}\end{cases} (4.20)

It remains to show that the vector fields Ek[α]𝔛α(A[2])E_{k}^{[\alpha]}\in\mathfrak{X}_{\alpha}(A^{[2]}) and ekα𝔛α(E2){e_{k}}^{\langle{\alpha}\rangle}\in\mathfrak{X}_{\alpha}(E^{2}), given in (3.18), are Θ2\mathrm{\Theta}^{2}-related. As Θ1\mathrm{\Theta}^{1} is the identity on AA we only need to show that

(Θ2)ekα(zμ)=Ek[α]((Θ2)zμ)=Ek[α](Qμiy˙i+12Qμ(ij)yiyj)(\mathrm{\Theta}^{2})^{\ast}{e_{k}}^{\langle{\alpha}\rangle}(z^{\mu})=E_{k}^{[\alpha]}((\mathrm{\Theta}^{2})^{\ast}z^{\mu})=E_{k}^{[\alpha]}(Q^{\mu}_{i}\dot{y}^{i}+\frac{1}{2}Q^{\mu}_{(ij)}y^{i}y^{j}) (4.21)

for α=0,1,2\alpha=0,-1,-2 where Θ2:A[2]E2\mathrm{\Theta}^{2}:A^{[2]}\rightarrow E^{2} is given in (3.3). For α=2\alpha=-2, this equation is satisfied automatically. For α=1\alpha=-1, it results in equation (4.19c). For α=0\alpha=0, the equation (4.21) can be expressed as a combination of (4.19b) and (4.19d). ∎

4.4 Representations up to homotopy of Lie algebroids

We shall review some calculus and sign conventions concerning representations up to homotopy of Lie algebroids. We follow the presentation given in [AC12].

Let (σ:AM,[,],)(\sigma:A\rightarrow M,[\cdot,\cdot],\sharp) be a Lie algebroid. Then Ω(A)=Γ(A)\Omega(A)=\operatorname{\Gamma}(\bigwedge^{\cdot}A^{\ast}) is known as the algebra of AA-differential forms. In the case the tangent algebroid, A=TMA=\mathrm{T}M, it is simply the algebra of differential forms on the manifold MM. There is an algebroid de Rham differential, called AA-differential dAd_{A}, which is a derivation of Ω(A)\Omega(A) such that

  1. (i)

    dA(f):s(s)(f)d_{A}(f):s\mapsto(\sharp s)(f), for sΓ(A)s\in\operatorname{\Gamma}(A), f𝒞(M)=Ω0(A)f\in\mathcal{C}^{\infty}(M)=\Omega^{0}(A);

  2. (ii)

    dA(ω):(s1,s2)ω([s1,s2])(s1)(ω,s2)+(s2)(ω,s1)d_{A}(\omega):(s_{1},s_{2})\mapsto\omega([s_{1},s_{2}])-(\sharp s_{1})\left(\langle\omega,s_{2}\rangle\right)+(\sharp s_{2})\left(\langle\omega,s_{1}\rangle\right), where s1,s2Γ(A)s_{1},s_{2}\in\operatorname{\Gamma}(A), ωΓ(A)=Ω1(A)\omega\in\operatorname{\Gamma}(A^{\ast})=\Omega^{1}(A).

It is well known that a Lie algebroid can be equivalently described by means of dAd_{A} — a degree 1 derivation on Ω(A)\Omega(A), see [Vai97].

Let FF be a vector bundle over the same base MM. An AA-connection on FF is a mapping :Γ(A)×Γ(F)Γ(F)\nabla:\operatorname{\Gamma}(A)\times\operatorname{\Gamma}(F)\rightarrow\operatorname{\Gamma}(F), (s,v)sv(s,v)\mapsto\nabla_{s}v such that

fsv=fsv,s(fv)=fsv+(s)(f)(v)\nabla_{fs}v=f\nabla_{s}v,\quad\nabla_{s}(fv)=f\nabla_{s}v+(\sharp s)(f)(v)

for f𝒞(M),vΓ(F),sΓ(A)f\in\mathcal{C}^{\infty}(M),v\in\operatorname{\Gamma}(F),s\in\operatorname{\Gamma}(A). Recall that the curvature of an AA-connection \nabla on FF is the tensor given by

curv(s1,s2)(v)=s1s2vs2s1v[s1,s2]v,\mathrm{curv}_{\nabla}(s_{1},s_{2})(v)=\nabla_{s_{1}}\nabla_{s_{2}}v-\nabla_{s_{2}}\nabla_{s_{1}}v-\nabla_{[s_{1},s_{2}]}v, (4.22)

where s1,s2Γ(A)s_{1},s_{2}\in\operatorname{\Gamma}(A), vΓ(F)v\in\operatorname{\Gamma}(F).

The space of FF-valued AA-differential forms is defined as Ω(A;F)=Γ(AF)\Omega(A;F)=\operatorname{\Gamma}(\bigwedge A^{\ast}\otimes F). In the setting of representations u.t.h., the vector bundle FF is \mathbb{Z}-graded, i.e., F=iFiF=\bigoplus_{i\in\mathbb{Z}}F^{i}, where FiF^{i} is, so called, the vector bundle of homogenous vectors of degree ii. Given graded vector bundles FF, GG over the same manifold MM, let Hom¯(F,G)=kHom¯k(F,G)\underline{\operatorname{Hom}}(F,G)=\bigoplus_{k\in\mathbb{Z}}\underline{\operatorname{Hom}}^{k}(F,G), where Hom¯k(F,G)\underline{\operatorname{Hom}}^{k}(F,G) denotes the bundle homomorphism from FF to GG that increase the degree by kk. In other words, the fiber (Hom¯k(F,G))x\left(\underline{\operatorname{Hom}}^{k}(F,G)\right)_{x} over xMx\in M is a collection of linear maps Ti:FixGi+kxT_{i}:F^{i}_{x}\rightarrow G^{i+k}_{x}. In the special case F=GF=G we write End¯(F)\underline{\operatorname{End}}(F) for Hom¯(F,F)\underline{\operatorname{Hom}}(F,F). An element ωΩi(A;Fj)\omega\in\Omega^{i}(A;F^{j}) is said to be of total degree |ω|=i+j|\omega|=i+j. There is an important operation, called the wedge product

Ωp(A;Ei)Ωq(A;Fj)Ωp+q(A;Gi+j),(α,β)αhβ,\Omega^{p}(A;E^{i})\otimes\Omega^{q}(A;F^{j})\to\Omega^{p+q}(A;G^{i+j}),\quad(\alpha,\beta)\mapsto\alpha\wedge_{h}\beta,

associated with a degree preserving graded vector bundle morphism h:EFGh:E\otimes F\to G. It is given by

(αhβ)(s1,s2,,sp+q)=σ(1)qisgn(σ)h(α(sσ(1),,sσp),β(sσ(k+1),,sσ(p+q))),(\alpha\wedge_{h}\beta)(s_{1},s_{2},\ldots,s_{p+q})=\sum_{\sigma}(-1)^{qi}\operatorname{sgn}(\sigma)h(\alpha(s_{\sigma(1)},\ldots,s_{\sigma_{p}}),\beta(s_{\sigma(k+1)},\ldots,s_{\sigma(p+q)})), (4.23)

where the summation is over all (p+q)(p+q)-shuffles, s1,s2,,sp+qΓ(A)s_{1},s_{2},\ldots,s_{p+q}\in\operatorname{\Gamma}(A). The left Ω(A)\Omega(A)-module structure on Ω(A;F)\Omega(A;F) is given by the wedge product associated with the isomorphism hL:FFh_{L}:\mathbb{R}\otimes F\xrightarrow{\simeq}F and is denoted by ω.η:=ωhLη\omega.\eta:=\omega\wedge_{h_{L}}\eta. On the other hand, the isomorphism hR:FFh_{R}:F\otimes\mathbb{R}\xrightarrow{\simeq}F gives rise to the right Ω(A)\Omega(A)-module structure on Ω(A;F)\Omega(A;F), η.ω=ηhRω\eta.\omega=\eta\wedge_{h_{R}}\omega that makes Ω(A;F)\Omega(A;F) a symmetric Ω(A)\Omega(A)-bimodule,

ω.η=(1)|ω||η|η.ω,\omega.\eta=(-1)^{|\omega||\eta|}\eta.\omega,

thanks to the sign (1)qi(-1)^{qi} in (4.23). We shall frequently encounter the case of the wedge product associated with the composition of homomorphisms :Hom(G,H)Hom(F,G)Hom(F,H)\circ:\operatorname{Hom}(G,H)\otimes\operatorname{Hom}(F,G)\rightarrow\operatorname{Hom}(F,H) which will be denoted by αβ:=αβ\alpha\bm{\wedge}\beta:=\alpha\wedge_{\circ}\beta. We have

(αβ)γ=α(βγ).(\alpha\bm{\wedge}\beta)\bm{\wedge}\gamma=\alpha\bm{\wedge}(\beta\bm{\wedge}\gamma). (4.24)

for Hom\operatorname{Hom}-valued AA-forms α\alpha, β\beta, γ\gamma. The operation \bm{\wedge} turns Ω(A;End(F)¯)\Omega(A;\underline{\operatorname{End}(F)}) into a graded associative algebra. The graded commutator on Ω(A;End¯(F))\Omega(A;\underline{\operatorname{End}}(F)) is defined by [α,β]=αβ(1)|α||β|βα[\alpha,\beta]=\alpha\bm{\wedge}\beta-(-1)^{|\alpha||\beta|}\beta\bm{\wedge}\alpha. Another case is the wedge product associated with the evaluation map ev:Hom(F,G)FG\operatorname{ev}:\operatorname{Hom}(F,G)\otimes{F}\rightarrow G which will be denoted in the same way as αη:=αevη\alpha\bm{\wedge}\eta:=\alpha\wedge_{\operatorname{ev}}\eta where αΩ(A,Hom¯(F,G))\alpha\in\Omega(A,\underline{\operatorname{Hom}}(F,G)), ηΩ(A;F)\eta\in\Omega(A;F), since the evaluation map is a special case of the composition of maps, thanks to the isomorphism FHom(,F)F\simeq\operatorname{Hom}(\mathbb{R},F).

Definition 4.6.

[AC12] A representation up to homotopy of a Lie algebroid AA consists of a \mathbb{Z}-graded vector bundle F=iFiF=\oplus_{i\in\mathbb{Z}}F^{i} and an operator, called the structure operator,

D:Ω(A;F)Ω(A;F)D:\Omega(A;F)\rightarrow\Omega(A;F)

of total degree one which satisfies DD=0D\circ D=0 and the graded derivation rule

D(ω.η)=dA(ω).η+(1)kω.D(η)D(\omega.\eta)=\mathrm{d}_{A}(\omega).\eta+(-1)^{k}\omega.D(\eta)

for ωΩk(A)\omega\in\Omega^{k}(A), ηΩ(A;F)\eta\in\Omega(A;F). A morphism Φ:(F,DF)(G,DG)\Phi:(F,D_{F})\to(G,D_{G}) linking two representations u.t.h. is a degree zero Ω(A)\Omega(A)-module map Φ:Ω(A;F)Ω(A;G)\Phi:\Omega(A;F)\to\Omega(A;G) which commutes with the structure operators DFD_{F} and DGD_{G}.

There is a one-to-one correspondence between AA-forms ωΩ(A;Hom¯(F,G))\omega\in\Omega(A;\underline{\operatorname{Hom}}(F,G)) of total degree nn and operators ωwidehat:Ω(A;F)Ω(A;G)\widehat{\omega}:\Omega(A;F)\to\Omega(A;G) of degree nn which are Ω(A)\Omega(A)-linear in the graded sense. The operator ωwidehat\widehat{\omega} is given by ωwidehat(η)=ωη\widehat{\omega}(\eta)=\omega\bm{\wedge}\eta. The equation (4.24) implies that αβwidehat=αwidehatβwidehat\widehat{\alpha\bm{\wedge}\beta}=\widehat{\alpha}\circ\widehat{\beta}, where βΩ(A;Hom¯(F,G))\beta\in\Omega(A;\underline{\operatorname{Hom}}(F,G)), αΩ(A;Hom¯(G,H))\alpha\in\Omega(A;\underline{\operatorname{Hom}}(G,H)).

A cochain complex (F,)(F,\partial) is a \mathbb{Z}-graded vector bundle FF equipped with an endomorphism End¯1(F)\partial\in\underline{\operatorname{End}}^{1}(F) such that =0\partial\circ\partial=0, i.e., a differential on FF. Such a differential can be consider as a 0-form with values in End¯(F)\underline{\operatorname{End}}(F), and gives rise to an operator widehat:Ωp(A;F)Ωp(A,F)\widehat{\partial}:\Omega^{p}(A;F)\to\Omega^{p}(A,F), widehat(η)=η\widehat{\partial}(\eta)=\partial\bm{\wedge}\eta. It satisfies widehatwidehat=0\widehat{\partial}\circ\widehat{\partial}=0 and

widehat(η)(x1,x2,,xp)=(1)p(η(x1,x2,,xp)),\widehat{\partial}(\eta)(x_{1},x_{2},\ldots,x_{p})=(-1)^{p}\partial(\eta(x_{1},x_{2},\ldots,x_{p})),

The sign (1)p(-1)^{p} comes from (4.23). Given two complexes (F,F)(F,\partial^{F}), (G,G)(G,\partial^{G}) and ωΩp(A;Hom¯i(F,G))\omega\in\Omega^{p}(A;\underline{\operatorname{Hom}}^{i}(F,G)) we get the induced pp-form Homω\partial^{\operatorname{Hom}}\omega defined as

Homω=Gω(1)|ω|ωF\partial^{\operatorname{Hom}}\omega=\partial^{G}\bm{\wedge}\omega-(-1)^{|\omega|}\omega\bm{\wedge}\partial^{F}

which takes values in Hom¯i+1(F,G)\underline{\operatorname{Hom}}^{i+1}(F,G), where |ω|=p+i|\omega|=p+i is the total degree of ω\omega. We have the complex (Hom¯(F,G),Hom)(\underline{\operatorname{Hom}}(F,G),\partial^{\operatorname{Hom}}) with the differential Hom\partial^{\operatorname{Hom}} obtained by specializing to the case p=0p=0 which reads as

(HomT)(v)=G(T(v))(1)|T|T(F(v)),(\partial^{\operatorname{Hom}}T)(v)=\partial^{G}(T(v))-(-1)^{|T|}T(\partial^{F}(v)),

where vFv\in F, THom¯(F,G)T\in\underline{\operatorname{Hom}}(F,G), and |T||T| stands for the degree of TT. It follows immediately from (4.24) that

Homωwidehat=Gwidehatωwidehat(1)|ω|ωwidehatFwidehat:Ω(A;F)Ω(A;G).\widehat{\partial^{\operatorname{Hom}}\omega}=\widehat{\partial^{G}}\circ\widehat{\omega}-(-1)^{|\omega|}\widehat{\omega}\circ\widehat{\partial^{F}}:\Omega(A;F)\rightarrow\Omega(A;G).

In the case F=GF=G we can write Homωwidehat=[Fwidehat,ωwidehat]\widehat{\partial^{\operatorname{Hom}}\omega}=[\widehat{\partial^{F}},\widehat{\omega}].

Besides, an AA-connection \nabla on a vector bundle FF induces an operator d\mathrm{d}_{\nabla} on Ω(A;F)\Omega(A;F) of degree one defined by means of Koszul formula. This formula can be derived from the conditions:

  1. (i)

    (dη)(s)=sη(\mathrm{d}_{\nabla}\eta)(s)=\nabla_{s}\eta for 0-forms ηΩ0(A;F)=Γ(F)\eta\in\Omega^{0}(A;F)=\operatorname{\Gamma}(F);

  2. (ii)

    the graded derivation rule: d(ω.η)=dA(ω).η+(1)|ω|ω.(dη)\mathrm{d}_{\nabla}(\omega.\eta)=\mathrm{d}_{A}(\omega).\eta+(-1)^{|\omega|}\omega.(\mathrm{d}_{\nabla}\eta).

Given AA-connections F\nabla^{F}, G\nabla^{G} on the graded vector bundles F,GF,G, respectively, we get an AA-connection on Hom¯(F,G)\underline{\operatorname{Hom}}(F,G) given by

(HomT)(v)=G(T(v))T(Fv)(\nabla^{\operatorname{Hom}}T)(v)=\nabla^{G}(T(v))-T(\nabla^{F}v) (4.25)

where vΓ(F)v\in\operatorname{\Gamma}(F), and TT is a section of Hom¯(F,G)\underline{\operatorname{Hom}}(F,G). The corresponding operator d\mathrm{d}_{\nabla} on Ω(A;Hom¯(F,G))\Omega(A;\underline{\operatorname{Hom}}(F,G)) is given by

dωwidehat=dGωwidehat(1)|ω|ωwidehatdF,ωΩ(A;Hom(F,G)).\widehat{\mathrm{d}_{\nabla}\omega}=\mathrm{d}_{\nabla^{G}}\circ\widehat{\omega}{-}(-1)^{|\omega|}\widehat{\omega}\circ\mathrm{d}_{\nabla^{F}},\quad\omega\in\Omega(A;\operatorname{Hom}(F,G)).

To prove it one shows that d\mathrm{d}_{\nabla} satisfies the graded derivation rule and that it reduces to the formula (4.25) when T=ωT=\omega is a 0-form.

The structure operator DD of a representation u.t.h. can be decomposed into a sequence of End¯(F)\underline{\operatorname{End}}(F)-valued AA-forms and an AA-connection giving an equivalent description. A precise statement is the following:

Proposition 4.7.

[AC12, Proposition 3.2 and Definition 3.3] The structure operator DD on a \mathbb{Z}-graded vector bundle FF can be equivalently given by a series of maps:

  • A degree 11 operator \partial on FF making (F,)(F,\partial) a complex, i.e., =0\partial\circ\partial=0.

  • An AA-connection F\nabla^{F} on (F,)(F,\partial), i.e., (Fv)=F(v)\partial(\nabla^{F}v)=\nabla^{F}(\partial v) for vΓ(F)v\in\operatorname{\Gamma}(F).

  • A 2-form ω2Ω2(A;End¯1(F))\omega_{2}\in\Omega^{2}(A;\underline{\operatorname{End}}^{-1}(F)) such that Homω2+curvF=0\partial^{\operatorname{Hom}}\omega_{2}+\mathrm{curv}_{\nabla^{F}}=0.

  • A sequence (ω2,ω3,)(\omega_{2},\omega_{3},\ldots) of End¯(F)\underline{\operatorname{End}}(F)-valued AA-forms, ωiΩi(A;End¯1i(F))\omega_{i}\in\Omega^{i}(A;\underline{\operatorname{End}}^{1-i}(F))171717Note that ωi\omega_{i} has total degree 1such that for each n3n\geq 3

    0=Homωn+dωn1+i=2n2ωiωniΩn(A;End¯2n(F)).0=\partial^{\operatorname{Hom}}\omega_{n}+\mathrm{d}_{\nabla}\omega_{n-1}+\sum_{i=2}^{n-2}\omega_{i}\bm{\wedge}\omega_{n-i}\in\Omega^{n}(A;\underline{\operatorname{End}}^{2-n}(F)). (4.26)

A morphism Φ\Phi from (F,DF)(F,D_{F}) to (G,DG)(G,D_{G}) is given by a sequence of AA-forms ΦiΩi(A;Hom¯i(F,G))\Phi_{i}\in\Omega^{i}(A;\underline{\operatorname{Hom}}^{-i}(F,G)) such that Φ0\Phi_{0} is a map of complexes and for each n1n\geq 1

0=HomΦn+dΦn1+i=2n2(ωiGΦniΦniωiF)Ωn(A;Hom¯1n(F,G)).0=\partial^{\operatorname{Hom}}\Phi_{n}+\mathrm{d}_{\nabla}\Phi_{n-1}+\sum_{i=2}^{n-2}(\omega_{i}^{G}\bm{\wedge}\Phi_{n-i}-\Phi_{n-i}\bm{\wedge}\omega_{i}^{F})\ \in\Omega^{n}(A;\underline{\operatorname{Hom}}^{1-n}(F,G)). (4.27)
Remark 4.8.

Note that Homωiwidehat=Fwidehatωiwidehat+ωiwidehatFwidehat\widehat{\partial^{\operatorname{Hom}}\omega_{i}}=\widehat{\partial^{F}}\circ\widehat{\omega_{i}}+\widehat{\omega_{i}}\circ\widehat{\partial^{F}} and dωiwidehat=dFωiwidehat+ωiwidehatdF\widehat{\mathrm{d}_{\nabla}\omega_{i}}=\mathrm{d}_{\nabla^{F}}\circ\widehat{\omega_{i}}+\widehat{\omega_{i}}\circ\mathrm{d}_{\nabla^{F}} (as |ωi|=1|\omega_{i}|=1) and the structure equation (4.26) is equivalent to DD=0D\circ D=0 where

D=Fwidehat+dF+ω2widehat+:Ω(A;F)Ω(A;F).D=\widehat{\partial^{F}}+\mathrm{d}_{\nabla^{F}}+\widehat{\omega_{2}}+\ldots:\Omega(A;F)\rightarrow\Omega(A;F). (4.28)

Similarly, as |Φ|=0|\Phi|=0, we have HomΦnwidehat=GwidehatΦwidehatnΦwidehatnFwidehat\widehat{\partial^{\operatorname{Hom}}\Phi_{n}}=\widehat{\partial^{G}}\circ\widehat{\Phi}_{n}-\widehat{\Phi}_{n}\circ\widehat{\partial^{F}}, dΦn1widehat=dGΦwidehatn1Φwidehatn1dF\widehat{\mathrm{d}_{\nabla}\Phi_{n-1}}=\mathrm{d}_{\nabla^{G}}\circ\widehat{\Phi}_{n-1}-\widehat{\Phi}_{n-1}\circ\mathrm{d}_{\nabla^{F}} and the equation (4.27) means that the operator Φwidehat=Φ0widehat+Φ1widehat+\widehat{\Phi}=\widehat{\Phi_{0}}+\widehat{\Phi_{1}}+\ldots intertwines the structure operators DFD_{F} and DGD_{G}.

References

  • [AC12] C. A. Abad and M Crainic. Representations up to homotopy of Lie algebroids. Journal für die Reine und Angewandte Mathematik, 663:91–126, 2012.
  • [BGG15a] A. J. Bruce, K. Grabowska, and J. Grabowski. Graded bundles in the category of Lie groupoids. Symmetry, Integrability and Geometry: Methods and Applications, 11, 2015.
  • [BGG15b] A. J. Bruce, K. Grabowska, and J. Grabowski. Higher-order mechanics on graded bundles. J. Phys. A: Mathematical and Theoretical, 48(20):205203, 2015.
  • [BGG16] A. J. Bruce, K. Grabowska, and J. Grabowski. Linear duals of graded bundles and higher analogues of (Lie) algebroids. J. Geom. Phys., 101:71–99, 2016.
  • [BGR16] A. J. Bruce, J. Grabowski, and M. Rotkiewicz. Polarisation of graded bundles. SIGMA Symmetry Integrability Geom. Methods Appl., 12:106, 2016.
  • [BGV18] A. J. Bruce, J. Grabowski, and L. Vitagliano. Representations up to homotopy from weighted Lie algebroids. Journal of Lie Theory, 28(3), 2018.
  • [BO19] O. Brahic and C. Ortiz. Integration of 2-term representations up to homotopy via 2-functors. Transactions of the American Mathematical Society, 372(1):503–543, 2019.
  • [DJLO15] T. Drummond, M. Jotz Lean, and C. Ortiz. VB-algebroid morphisms and representations up to homotopy. Differential Geometry and its Applications, 40:332–357, 2015.
  • [EVT19] C. Esposito, L. Vitagliano, and A. Tortorella. Infinitesimal Automorphisms of VB-Groupoids and Algebroids. The Quarterly Journal of Mathematics, 70(3):1039–1089, 2019.
  • [GG08] K. Grabowska and J. Grabowski. Variational calculus with constraints on general algebroids. J. Phys. A, 41(17):175204, 2008.
  • [GR09] J. Grabowski and M. Rotkiewicz. Higher vector bundles and multi-graded symplectic manifolds. J. Geom. Phys., 59:1285–1305, 2009.
  • [GR11] J. Grabowski and M. Rotkiewicz. Graded bundles and homogeneity structures. J. Geom. Phys., 62:21–36, 2011.
  • [Gra12] J. Grabowski. Modular classes of skew algebroid relations. Transform. Groups, 17:989–1010, 2012.
  • [GSJLMM18] A. Gracia-Saz, M. Jotz Lean, K. C. H. Mackenzie, and R. A. Mehta. Double Lie algebroids and representations up to homotopy. Journal of Homotopy and Related Structures, 223:287–319, 2018.
  • [GSM10] A. Gracia-Saz and R. Mehta. Lie algebroid structures on double vector bundles and representation theory of Lie algebroids. Advances in Mathematics, 223:1236–1275, 2010.
  • [GU99] J. Grabowski and P. Urbański. Algebroids – general differential calculi on vector bundles. J. Geom. Phys., 31:111–141, 1999.
  • [JR13] M. Jóźwikowski and M. Rotkiewicz. Higher algebroids with applications to variational calculus. arXiv:1306.3379v1, 2013.
  • [JR15] M. Jóźwikowski and M. Rotkiewicz. Models for higher algebroids. J. Geom. Mech., 7:317–359, 2015.
  • [JR18] M. Jóźwikowski and M. Rotkiewicz. Higher-order analogs of Lie algebroids via vector bundle comorphisms. SIGMA Symmetry Integrability Geom. Methods Appl., 14:135, 2018. arXiv:1708.03174.
  • [KSM02] Y. Kosmann-Schwarzbach and K. C. H Mackenzie. Differential operators and actions of Lie algebroids, volume 315 of Contemporary Mathematics, pages 213–233. Amer. Math. Soc., 2002.
  • [Mac05] K. C. H. Mackenzie. General theory of Lie groupoids and Lie algebroids. Cambridge University Press, 2005.
  • [Mar01] E. Martínez. Geometric formulation of mechanics on Lie algebroids. In Proceedings of the VIII Fall Workshop on Geometry and Physics, volume 2, pages 209–222. Publicaciones de la RSME, 2001.
  • [Mar15] E. Martínez. Higher-order variational calculus on Lie algebroids. J. Geom. Mech., 7(1):81–108, 2015.
  • [Mor70] A. Morimoto. Liftings of tensor fields and connections to tangent bundles of higher-order. Nagoya Math. J., 40:99–120, 1970.
  • [Pop04] Paul Popescu. On higher order geometry on anchored vector bundles. Open Mathematics, 2(5):826–839, 2004.
  • [Pra75] J. Pradines. Fibrés vectoriels doubles et calcul des jets non holonomes. PhD thesis, 1975.
  • [Vai97] A. Yu. Vaintrob. Lie algebroids and homological vector fields. Russian Mathematical Surveys, 52(2):428, 1997.
  • [Vor02] Th. Th. Voronov. Graded manifolds and Drinfeld doubles for Lie bialgebroids, volume 315 of Contemporary Mathematics, pages 131–168. Amer. Math. Soc., 2002.
  • [Vor10] Th.Th. Voronov. Q-manifolds and higher analogs of Lie algebroids. AIP Conf. Proc., 1307:191–202, 2010.