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

Relative algebroids and Cartan realization problems

Rui Loja Fernandes Department of Mathematics, University of Illinois at Urbana-Champaign, 1409 W. Green Street, Urbana, IL 61801 USA ruiloja@illinois.edu  and  Wilmer Smilde Department of Mathematics, University of Illinois at Urbana-Champaign, 1409 W. Green Street, Urbana, IL 61801 USA wsmilde2@illinois.edu
Abstract.

We develop a new framework of relative algebroids to address existence and classification problems of geometric structures subject to partial differential equations.

This work was partially supported by NSF grant DMS-2303586 and UIUC Campus Research Board Award RB25014.

Introduction

The role of algebroids in certain classification problems was first explicitly recognized by Bryant in his work on the classification of Bochner-Kähler metrics [7]. Struchiner and the first author built on this insight, establishing precise connections between existence and classification problems and the integrability of the underlying Lie algebroid [17, 18, 19]. However, their work is restricted to cases where the local classification is finite-dimensional in nature, which excludes most classification problems. In this paper, we initiate the study of classifications problems without this restriction. We introduce here the concept of relative Lie algebroid, which unifies the theories of algebroids and (formal) PDEs. This yields a powerful tool for understanding generic existence and classification problems for geometric structures.

The new notion of a relative algebroid has roots in the work of Bryant, particularly in his notes on Lie theory and Exterior Differential Systems [8]. There, he observed through various examples that many existence problems can be recast in a particular form—referred to here as Bryant’s equations (0.5)—which resembles an algebroid but is not quite one [6, 8]. Although Bryant did not identify this notion explicitly, he showed that techniques going back to Cartan can be applied to solve these equations in many interesting examples.

In the rest of this introduction, we will first explain what Bryant’s equations are, where they come from and how they are used in existence and classification problems. Then, we will briefly describe the contents of this paper.

Bryant’s equations and two examples

Let (M,g)(M,g) be a Riemannian manifold. Recall that its orthogonal frame bundle

π:PM.\pi\colon P\to M.

carries a coframe

(θ,ω):TPn𝔬(n).(\theta,\omega)\colon TP\to\mathbb{R}^{n}\oplus\mathfrak{o}(n).

whose components are the tautological form (also called the solder form) and the Levi-Civita connection 1-form. Together they satisfy Cartan’s structure equations

{dθ=ωθ,dω=R(θθ)ωω,\begin{cases}{\mathrm{d}}\theta=-\omega\wedge\theta,\\ {\mathrm{d}}\omega=R(\theta\wedge\theta)-\omega\wedge\omega,\end{cases} (0.1)

where R:PHom(2n,𝔬(n))R\colon P\to\operatorname{Hom}(\wedge^{2}\mathbb{R}^{n},\mathfrak{o}(n)) is a map inducing the Riemann curvature tensor when passing to the associate bundles.

The structure equations (0.1) completely characterize the orthonormal frame bundles of Riemannian manifolds, and allows to relate existence and classification problems for Riemannian structures to those of coframes, as observed by Cartan himself. In a classification problem one typically has restrictions on the curvature RR in the form of an algebraic or a differential equation (or both). Let us recall one simple example where one already sees the appearance of a Lie algebroid.

Example 0.1 (Space forms).

For the classification of Riemannian manifolds (M,g)(M,g) with constant sectional curvature, a.k.a. space forms, the Riemann curvature R:PHom(2n,𝔬(n))R\colon P\to\operatorname{Hom}(\wedge^{2}\mathbb{R}^{n},\mathfrak{o}(n)) takes the special form:

R(K)(u,v)w=K(w,uvw,vu),R(K)(u,v)w=K\left(\langle w,u\rangle v-\langle w,v\rangle u\right),

where KK is a constant (the scalar curvature), u,v,wnu,v,w\in\mathbb{R}^{n} and ,\langle\cdot,\cdot\rangle is the Euclidean inner product on n\mathbb{R}^{n}. So one considers Cartan’s structure equations with this specific shape of the tensor together with the equation that KK is constant:

{dθ=ωθdω=R(K)(θθ)ωωdK=0.\begin{cases}{\mathrm{d}}\theta=-\omega\wedge\theta\\ {\mathrm{d}}\omega=R(K)(\theta\wedge\theta)-\omega\wedge\omega\\ {\mathrm{d}}K=0.\end{cases} (0.2)

These are Bryant’s equations for this classification problem.

Equations (0.2) already show the appearance of a Lie algebroid behind this classification problem of space forms. We ”forget” about the underlying manifold these equations were derived from and look at the system “as is”. More precisely, we consider the trivial vector bundle A=n𝔬(n)¯A=\underline{\mathbb{R}^{n}\oplus\mathfrak{o}(n)}\to\mathbb{R}, we interpret KK has a coordinate function the base, and θ\theta and ω\omega as generating sections of the dual vector bundle. Equations (0.2) then define a linear operator

D:Ω(A)Ω+1(A),where Ωk(A):=Γ(kA).{\mathrm{D}}\colon\Omega^{\bullet}(A)\to\Omega^{\bullet+1}(A),\quad\text{where }\Omega^{k}(A):=\Gamma(\wedge^{k}A^{*}).

This is a graded derivation satisfying D2=0{\mathrm{D}}^{2}=0, hence it defines a Lie algebroid structure on AA. In [18] it is explained how integration of this Lie algebroid leads to the well-known classification of space forms.

In the previous example the “moduli space” of solutions is one dimensional: there is only one invariant, the scalar curvature. For such classification problems of finite-type a complete theory was developed in [17, 18, 19]. However, in practice, such problems are rarely encountered. Our next example illustrates a more typical occurrence.

Example 0.2 (Surfaces with |K|=1|\nabla K|=1, [8, §5.1]).

Consider the problem of classifying Riemann surfaces (M,g)(M,g) whose scalar curvature KK satisfies the differential equation |K|=1|\nabla K|=1.

Let (P,θ,ω)(P,\theta,\omega) be the orthonormal frame bundle of such a surface and identify 2𝔬(1)3\mathbb{R}^{2}\oplus\mathfrak{o}(1)\cong\mathbb{R}^{3}. Writing (θ,ω)=(θ1,θ2,θ3)(\theta,\omega)=(\theta^{1},\theta^{2},\theta^{3}). The structure equations become

{dθ1=θ3θ1,dθ2=θ3θ1,dθ3=Kθ1θ2,\begin{cases}{\mathrm{d}}\theta^{1}=-\theta^{3}\wedge\theta^{1},\\ {\mathrm{d}}\theta^{2}=\theta^{3}\wedge\theta^{1},\\ {\mathrm{d}}\theta^{3}=K\theta^{1}\wedge\theta^{2},\end{cases}

where KK is the Gauss curvature. Since KK is an O(2)O(2)-invariant function on PP it’s derivative can be written in the form

dK=K1θ1+K2θ2.{\mathrm{d}}K=K_{1}\theta^{1}+K_{2}\theta^{2}.

The equation |K|=1|\nabla K|=1 becomes K12+K22=1K_{1}^{2}+K_{2}^{2}=1, so the components KiK_{i} are related through a single coordinate φ\varphi on the circle. The structure equations augmented by these conditions on the curvature take the form:

{dθ1=θ3θ1,dθ2=θ3θ1,dθ3=Kθ1θ2,dK=cos(φ)θ1+sin(φ)θ2.\begin{cases}{\mathrm{d}}\theta^{1}=-\theta^{3}\wedge\theta^{1},\\ {\mathrm{d}}\theta^{2}=\theta^{3}\wedge\theta^{1},\\ {\mathrm{d}}\theta^{3}=K\theta^{1}\wedge\theta^{2},\\ {\mathrm{d}}K=\cos(\varphi)\theta^{1}+\sin(\varphi)\theta^{2}.\end{cases} (0.3)

These are Bryant’s equations for this classification problem.

As in the previous example, the next step is to ”forget” about the manifold PP and look at the equations “as they are”. So, as in that example, we consider the trivial vector bundle A=3¯A=\underline{\mathbb{R}^{3}}\to\mathbb{R}, where \mathbb{R} has coordinate KK, and think of the θi\theta^{i}’s as generating sections of the dual vector bundle. But now we run into a problem: equations (0.3) define a degree 1 linear operator D{\mathrm{D}} whose values depend on the extra variable φ\varphi. To solve this issue we consider the vector bundle B=3¯SS1×B=\underline{\mathbb{R}^{3}}\to\SS^{1}\times\mathbb{R}, where SS1×\SS^{1}\times\mathbb{R} has coordinates (φ,K)(\varphi,K), so that now

D:Ω(A)Ω+1(B){\mathrm{D}}\colon\Omega^{\bullet}(A)\to\Omega^{\bullet+1}(B)

If p:SS1×p\colon\SS^{1}\times\mathbb{R}\to\mathbb{R} is the projection, so that B=pAB=p^{*}A, it makes sense to call D{\mathrm{D}} a derivation relative to pp. This does not define a Lie algebroid anymore since, for instance, the equation D2=0{\mathrm{D}}^{2}=0 doesn’t make sense at this point. So we cannot directly use integration techniques for Lie algebroids to obtain solutions.

One could try to look for a derivation DB{\mathrm{D}}_{B} on BB that extends D{\mathrm{D}} and does square to zero by adding the extra equation DB2K=0{\mathrm{D}}_{B}^{2}K=0. This implies

0=DB2K\displaystyle 0={\mathrm{D}}_{B}^{2}K =DB(cos(φ)θ1+sin(φ)θ2)\displaystyle={\mathrm{D}}_{B}(\cos(\varphi)\theta^{1}+\sin(\varphi)\theta^{2})
=sin(φ)DBφθ1cos(φ)θ3θ2+cos(φ)DBφθ2+sin(φ)θ3θ1.\displaystyle=-\sin(\varphi){\mathrm{D}}_{B}\varphi\wedge\theta^{1}-\cos(\varphi)\theta^{3}\wedge\theta^{2}+\cos(\varphi){\mathrm{D}}_{B}\varphi\wedge\theta^{2}+\sin(\varphi)\theta^{3}\wedge\theta^{1}.

It follows that, for this equation to hold, we must have

DBφ=θ3+c1(sin(φ)θ1+cos(φ)θ2),{\mathrm{D}}_{B}\varphi=\theta^{3}+c_{1}(-\sin(\varphi)\theta^{1}+\cos(\varphi)\theta^{2}), (0.4)

where c1c_{1} is a new independent variable. So, again DB{\mathrm{D}}_{B} is not an actual derivation but a derivation relative to the projection ×SS1×SS1×\mathbb{R}\times\SS^{1}\times\mathbb{R}\to\SS^{1}\times\mathbb{R}, (c1,φ,K)(φ,K)(c_{1},\varphi,K)\mapsto(\varphi,K).

This process never stops, and for this reason there is no finite dimensional Lie algebroid governing this classification problem. We will return to this example in Section 6.

These, and many other examples, led Bryant to observe that many classification problems amount to solve a problem involving data consisting of a principal bundle PP with a coframe θ:TPn\theta\colon TP\to\mathbb{R}^{n}, together with functions (aμ,bϱ):Ps×r(a^{\mu},b^{\varrho})\colon P\to\mathbb{R}^{s}\times\mathbb{R}^{r}, satisfying Bryant’s equations:

{dθi=12cjki(a)θjθk,daμ=Fiμ(a,b)θi,\begin{cases}{\mathrm{d}}\theta^{i}=-\frac{1}{2}c^{i}_{jk}(a)\,\theta^{j}\wedge\theta^{k},\\ {\mathrm{d}}a^{\mu}=F^{\mu}_{i}(a,b)\,\theta^{i},\end{cases} (0.5)

for some given functions cjki:sc^{i}_{jk}\colon\mathbb{R}^{s}\to\mathbb{R} and Fiμ:s×rF^{\mu}_{i}\colon\mathbb{R}^{s}\times\mathbb{R}^{r}\to\mathbb{R}.

As in the two problems above, “forgetting” about the underlying bundle PP, treating (a,b)(a,b) as independent coordinates on s×r\mathbb{R}^{s}\times\mathbb{R}^{r} and θ\theta as sections of the bundle dual to the trivial vector bundle A:=n¯r×sA:=\underline{\mathbb{R}^{n}}\to\mathbb{R}^{r}\times\mathbb{R}^{s}, Bryant’s equations define a degree 1, linear operator

D:Ω(A)Ω+1(B),with B=pA,{\mathrm{D}}\colon\Omega^{\bullet}(A)\to\Omega^{\bullet+1}(B),\quad\text{with }B=p^{*}A,

where p:s×rsp\colon\mathbb{R}^{s}\times\mathbb{R}^{r}\to\mathbb{R}^{s} is the projection.

PDEs and derivations

Degree 1 differential operators similar to the ones in the previous example, are well-known in the formal theory of PDEs. For example, there one considers the so-called total exterior differential. In its simplest form, one has a function in jet space, say f=f(x,y,u)f=f(x,y,u), where x,yx,y are independent variables and uu is a dependent variable, and then its total differential is given by

Df:=(Dxf)dx+(Dyf)dy=(fx+fuux)dx+(fy+fuuy)dy,{\mathrm{D}}f:=({\mathrm{D}}_{x}f)\,{\mathrm{d}}x+({\mathrm{D}}_{y}f)\,{\mathrm{d}}y=(f_{x}+f_{u}u_{x}){\mathrm{d}}x+(f_{y}+f_{u}u_{y}){\mathrm{d}}y,

Similarly, for a 1-form α=a(x,y,u)dx+b(x,y,u)dy\alpha=a(x,y,u){\mathrm{d}}x+b(x,y,u){\mathrm{d}}y its total differential is defined by

Dα:=(DxbDya)dxdy=(bx+buuxayauuy)dxdy.{\mathrm{D}}\alpha:=\big{(}{\mathrm{D}}_{x}b-{\mathrm{D}}_{y}a\big{)}{\mathrm{d}}x\wedge{\mathrm{d}}y=\big{(}b_{x}+b_{u}u_{x}-a_{y}-a_{u}u_{y}\big{)}{\mathrm{d}}x\wedge{\mathrm{d}}y.

If q:32q\colon\mathbb{R}^{3}\to\mathbb{R}^{2} is the submersion (x,y,u)(x,y)(x,y,u)\mapsto(x,y) and we denote by J1qJ^{1}q its 1st jet bundle with coordinates (x,y,u,ux,uy)(x,y,u,u_{x},u_{y}), then one can think of D{\mathrm{D}} as a degree 1, linear operator

D:Ω(A)Ω+1(B),where A:=qT2,B:=pA.{\mathrm{D}}\colon\Omega^{\bullet}(A)\to\Omega^{\bullet+1}(B),\quad\text{where }A:=q^{*}T\mathbb{R}^{2},\ B:=p^{*}A.

Relative derivations, relative algebroids and the contents of this paper

The common geometric structure underlying the previous examples is captured by a relative derivation. More precisely, given a vector bundle ANA\to N and a map p:MNp\colon M\to N, a degree 1 derivation relative to pp is a linear operator

D:Ω(A)Ω+1(pA),{\mathrm{D}}\colon\Omega^{\bullet}(A)\to\Omega^{\bullet+1}(p^{*}A),

satisfying

D(αβ)=(Dα)pβ+(1)|α|pαDβ,{\mathrm{D}}(\alpha\wedge\beta)=({\mathrm{D}}\alpha)\wedge p^{*}\beta+(-1)^{|\alpha|}p^{*}\alpha\wedge{\mathrm{D}}\beta,

for homogeneous elements α,βΩ(A)\alpha,\beta\in\Omega^{\bullet}(A). The resemblance with Lie algebroids now becomes clear: a Lie algebroid is a degree 1 derivation D{\mathrm{D}} on a vector bundle ANA\to N relative to the identity map, which additionally satisfies D2=0{\mathrm{D}}^{2}=0. For a general relative derivation D{\mathrm{D}} cannot be squared since domain and codomain are distinct. Hence, the equation D2=0{\mathrm{D}}^{2}=0 does not seem to make sense. In this paper, we will develop the necessary theory to make sense of this in a natural way.

In Section 1, we develop the theory of relative degree kk derivations, for arbitrary kk. Just like ordinary degree kk derivations which can be viewed, dually, as kk-nary brackets, relative derivations can be viewed, dually, as relative kk-brackets. Besides analyzing the structure of relative multiderivarions and relative multibrackets, we will extend several notions from the theory of (formal) PDEs to this setting. These include

  • a notion of a tableau (bundle) of derivations,

  • a Spencer complex for tableaux of derivations,

  • involutivity and Cartan characters for relative derivations.

The computations and examples in [8] suggest that Bryant may have already been aware of such notions. We intend to place his computations within a natural and robust framework.

In Section 2, we introduce and study the central concept in this paper, namely the notion of relative algebroid. This is a triple (A,p,D)(A,p,{\mathrm{D}}), where ANA\to N is a vector bundle and D{\mathrm{D}} is a degree 1 derivation relative to a map p:MNp\colon M\to N. As we already pointed out, these are the geometric objects underlying Bryant’s equations. Similar to the theory developed in [7, 17, 18], solutions to Bryant’s equations are translated to the notion of realization of a relative algebroid, so one speaks of Cartan’s realization problem. In the language of derivations, a realization allows to realize a relative derivation as a manifold with the de Rham differential. More precisely, given a relative algebroid (A,p,D)(A,p,{\mathrm{D}}) a realization is a manifold PP together with a bundle map θ:TPpA\theta\colon TP\to p^{*}A that is fiberwise an isomorphism and is also a morphism of relative algebroids, i.e., satisfies

dθp=θD.{\mathrm{d}}\circ\theta^{*}\circ p^{*}=\theta^{*}\circ{\mathrm{D}}.

In the case of finite-type problems, realizations naturally appear as fibers of an integrating Lie groupoid, making it possible to apply the power of Lie theory to solve classification problems (see [17, 18]). In the general case, however, the situation is much more complex. As we will see, the theory of relative derivations unifies Lie algebroids and PDEs.

We will show first that the formal theory of the existence of realizations parallels that of partial differential equations, as developed by Spencer and his school, and later formalized by Goldschmidt in [20]. Specifically, we will show that first-order and second order obstructions to the existence of realizations are captured by Spencer cohomology classes, called the torsion class and the curvature class.

In Section 3, we will see that the vanishing of the torsion class leads to the notion of prolongation, which seeks to complete the derivation of a relative algebroid to an operator that squares to zero. More concretely, a prolongation of a relative algebroid (A,p,D)(A,p,{\mathrm{D}}) is another relative algebroid (B,p1,D(1))(B,p_{1},{\mathrm{D}}^{(1)}), with projection p1:M(1)Mp_{1}\colon M^{(1)}\to M. Here, B=pAB=p^{*}A and the derivation is a map D(1):Ω(B)Ω+1(B(1)){\mathrm{D}}^{(1)}\colon\Omega^{\bullet}(B)\to\Omega^{\bullet+1}(B^{(1)}), where B(1):=p1BB^{(1)}:=p_{1}^{*}B, such that D(1)D=0{\mathrm{D}}^{(1)}\circ{\mathrm{D}}=0. Schematically, this is described by the diagram

B(1)=p1B\textstyle{B^{(1)}=p_{1}^{*}B\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B=pA\textstyle{B=p^{*}A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D(1)\scriptstyle{{\mathrm{D}}^{(1)}}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D\scriptstyle{{\mathrm{D}}}M(1)\textstyle{M^{(1)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p1\scriptstyle{p_{1}}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}N\textstyle{N}

The existence of the prolongation is contingent on the vanishing of the torsion class. Higher prolongations are obtained by iterating the first prolongations and are indexed by kk. A relative algebroid is called kk-integrable if all prolongations up to order kk exist, and is called formally integrable if it is kk-integrable for all k1k\geq 1. Foundational results from the theory of formal PDEs, such as Goldschmidt’s formal integrability theorem [20, Thm. 8.1], have natural generalizations to the theory of relative algebroids, namely we will prove the following analogue (or rather, extension) of that result for relative algebroids.

Theorem 1 (Theorem 3.10).

Let (A,p,D)(A,p,{\mathrm{D}}) be a relative algebroid with tableau of derivations τ\tau. Suppose that:

  1. (i)

    (A,p,D)(A,p,{\mathrm{D}}) is 1-integrable;

  2. (ii)

    the Spencer cohomology groups Hk,2(τ)H^{k,2}(\tau) vanish for all k0k\geq 0.

Then (A,p,D)(A,p,{\mathrm{D}}) is formally integrable.

If, for some kk, the map pk:M(k)M(k1)p_{k}\colon M^{(k)}\to M^{(k-1)} is a diffeomorphism, the derivation D(k){\mathrm{D}}^{(k)} actually defines the structure of a Lie algebroid, and we are in the finite-type case. However, in general – as in Example 0.2 – this process does not stop, and the full prolongation tower defines a profinite Lie algebroid. For this reason, we call a relative algebroid a relative Lie algebroid if all prolongations exist. Finding solutions to the realization problem in this more general case is much more challenging. The only general statement we can make is in the analytic setting, where an existence result, essentially due to Cartan and Kähler – see Bryant [8, Thm. 3 and Thm. 4] – holds. We state it as follows:

Theorem 2 (Bryant-Cartan, Theorem 3.20).

Let (A,p,D)(A,p,{\mathrm{D}}) be an analytic relative Lie algebroid. For each kk and each xM(k)x\in M^{(k)}, there exists a realization through xx.

In Section 4, we consider several important constructions with relative algebroids. First, given some vector bundle ANA\to N we construct a universal relative algebroid (A,p1,D˘)(A,p_{1},\breve{\mathrm{D}}), which is formally integrable, with the property that every algebroid (A,p,D)(A,p,{\mathrm{D}}) relative to a submersion p:MNp\colon M\to N is a pullback of the universal one via a classifying map. Then we discuss the operation of restriction to subspaces and how this operation interacts with prolongations and realizations. Finally, we introduce the notion of systatic foliation of a relative algebroid, which captures the directions in which the tableau map is zero. We show that these directions are essentially “redundant” from the perspective of the realization problem, since the original almost relative algebroid descends to a reduced almost relative algebroid, which has essentially the same realizations and the same integrability properties.

In Section 5, our discussion comes to a full circle by showing that any partial differential equation can be recast as a relative algebroid in such a way that the formal theory of prolongations of PDEs coincides with the prolongation theory for relative algebroids. For this, we first interpret the Cartan distribution on a jet space as a relative connection, and then we show that any relative connection has an associated relative algebroid. The derivation corresponding to the Cartan distribution is nothing more than the horizontal differential in the first row of the variational bicomplex. Then, given a kk-th order PDE EE viewed as a submanifold in the jet space JkqJ^{k}q of a submersion q:NXq\colon N\to X, by pulling back the Cartan derivation one obtains the associated relative algebroid. We will prove the following result:

Theorem 3 (Theorem 5.12).

Let EJkqE\subset J^{k}q be a PDE. Then, germs of solutions to EE are in one-to-one correspondence with germs of realizations of the associated relative algebroid, modulo diffeomorphisms. Moreover:

  1. (i)

    EE is a 1-integrable PDE if and only if the associated relative algebroid is 1-integrable.

  2. (ii)

    If EE is a 1-integrable PDE, then the relative algebroid corresponding to its first prolongation E(1)Jk+1qE^{(1)}\subset J^{k+1}q is the prolongation of the relative algebroid associated with EE.

In particular, a PDE is formally integrable if and only if its associated relative algebroid is.

In Section 6, we return to Example 0.2 and discuss its solutions (i.e., realizations) in light of the theory developed in the previous sections.

The reader will notice that this paper lays only the foundations of the theory. There are many promising directions to explore, often involving connections to existing literature and related fields of mathematics. We conclude the paper with an outlook on future work in Section 7.

Acknowledgements

We would like to thank Luca Accornero, Francesco Cattafi, Marius Crainic, Ivan Struchiner, and Luca Vitagliano for many discussions that helped us shape the ideas presented here. We especially thank Ori Yudilevich, who several years ago coauthored some foundational work that led to this paper but whose professional path has since taken him elsewhere.

1. Relative derivations

In this section, we develop the theory of relative derivations. While after this section, we will only encounter relative derivations of degree 1 and 2, we consider here relative derivations of arbitrary degree. We believe that higher-degree relative derivations, like their ordinary counterparts, will prove useful in other applications.

1.1. Derivations and brackets

We recall here some basic facts about kk-derivations and kk-brackets. For more details and proofs we refer the reader to [15].

Let VNV\to N be a vector bundle and set Ω(V):=Γ(V)\Omega^{\bullet}(V):=\Gamma(\wedge^{\bullet}V^{*}). A kk-derivation on VV is a graded derivation of Ω(V)\Omega^{\bullet}(V) of degree kk, i.e., a linear map D:Ω(V)Ω+k(V){\mathrm{D}}\colon\Omega^{\bullet}(V)\to\Omega^{\bullet+k}(V) satisfying

D(αβ)=(Dα)β+(1)|α|kα(Dβ).{\mathrm{D}}(\alpha\wedge\beta)=\left({\mathrm{D}}\alpha\right)\wedge\beta+(-1)^{|\alpha|k}\alpha\wedge\left({\mathrm{D}}\beta\right).

for homogeneous α,βΩ(V)\alpha,\beta\in\Omega^{\bullet}(V). The symbol of a kk-derivation MM is the bundle map σ(D):TNkV\sigma({\mathrm{D}})\colon T^{*}N\to\wedge^{k}V^{*} defined by

σ(D)(df)=D(f), for fC(N).\sigma({\mathrm{D}})({\mathrm{d}}f)={\mathrm{D}}(f),\mbox{ for $f\in C^{\infty}(N)$}.

Any derivation is determined by its symbol and its action on a generating set (over C(N)C^{\infty}(N)) of Ω1(V)\Omega^{1}(V).

A dual point of view to multi-derivations is via multi-brackets. A kk-bracket on a vector bundle VNV\to N is a skew-symmetric kk-linear map

[,,]:Γ(V)××Γ(V)(k+1)-timesΓ(V)[\cdot,\dots,\cdot]\colon\underbrace{\Gamma(V)\times\cdots\times\Gamma(V)}_{\text{$(k+1)$-times}}\to\Gamma(V)

together with an anchor ρ:kVTN\rho\colon\wedge^{k}V\to TN satisfying the Leibniz rule:

[v0,,fvk]=f[v0,,vk]+ρ(v0,,vk1)(f)vk,[v_{0},\cdots,fv_{k}]=f[v_{0},\cdots,v_{k}]+\rho(v_{0},\dots,v_{k-1})(f)v_{k},

for v0,,vkΓ(V),fC(N)v_{0},\dots,v_{k}\in\Gamma(V),f\in C^{\infty}(N).

The notions of kk-brackets and kk-derivations are in duality through the Koszul formula (see [15, §2.5]):

σ(D)(df),v1vk)\displaystyle\langle\sigma({\mathrm{D}})({\mathrm{d}}f),v_{1}\wedge\dots\wedge v_{k}) =df,ρ(v1,,vk)\displaystyle=\langle{\mathrm{d}}f,\rho(v_{1},\dots,v_{k})\rangle
Dα(v0,,vk)\displaystyle{\mathrm{D}}\alpha(v_{0},\dots,v_{k}) =i=0k(1)iρ(v0,,vi^,,vk)(α(vi))α([v0,,vk]),\displaystyle=\sum_{i=0}^{k}(-1)^{i}\rho(v_{0},\dots,\widehat{v_{i}},\dots,v_{k})(\alpha(v_{i}))-\alpha([v_{0},\dots,v_{k}]),

for fC(N)f\in C^{\infty}(N), αΩ1(V)\alpha\in\Omega^{1}(V) and v0,,vkΓ(V)v_{0},\dots,v_{k}\in\Gamma(V).

Proposition 1.1.

The space of kk-derivations on VV, denoted Derk(V)\operatorname{Der}^{k}(V), is the space of sections of a vector bundle 𝒟VkN\mathcal{D}_{V}^{k}\to N, and the symbol induces a short exact sequence

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(V,k+1V)\textstyle{\operatorname{Hom}(V^{*},\wedge^{k+1}V^{*})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟Vk\textstyle{\mathcal{D}_{V}^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}Hom(TN,kV)\textstyle{\operatorname{Hom}(T^{*}N,\wedge^{k}V^{*})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

Dually, using the canonical identification Hom(W,V)Hom(V,W)\operatorname{Hom}(W^{*},V^{*})\cong\operatorname{Hom}(V,W), one has

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(k+1V,V)\textstyle{\operatorname{Hom}(\wedge^{k+1}V,V)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟Vk\textstyle{\mathcal{D}_{V}^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}Hom(kV,TN)\textstyle{\operatorname{Hom}(\wedge^{k}V,TN)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0,\textstyle{0,} (1.1)

which is the sequence that we will use in practice.

A choice of connection \nabla on VV determines a splitting 𝒟VkHom(k+1V,V)\mathcal{D}_{V}^{k}\to\operatorname{Hom}(\wedge^{k+1}V,V) of this sequence. Namely, for any kk-bracket, the expression

[v0,,vk]:=[v0,,vk]+(1)k+1i=0k(1)iρ(v0,,vi^,,vk)vi,[v_{0},\dots,v_{k}]_{\nabla}:=[v_{0},\dots,v_{k}]+(-1)^{k+1}\sum_{i=0}^{k}(-1)^{i}\nabla_{\rho(v_{0},\dots,\widehat{v_{i}},\dots,v_{k})}v_{i}, (1.2)

is C(M)C^{\infty}(M)-multilinear.

Example 1.2.

A derivation D:Ω(V)Ω(V){\mathrm{D}}\colon\Omega^{\bullet}(V)\to\Omega^{\bullet}(V) of degree 0 corresponds to a linear vector field on VV. The flow φt\varphi_{t} of this linear vector field is a 1-parameter family of vector bundle maps that solves the ODE

{ddtφtα=φt(Dα),φ0=idV.\begin{cases}\frac{{\mathrm{d}}}{{\mathrm{d}}t}\varphi_{t}^{*}\alpha=\varphi^{*}_{t}({\mathrm{D}}\alpha),\\ \varphi_{0}=\operatorname{id}_{V}.\end{cases}
Example 1.3.

A Lie algebroid structure on a vector bundle AMA\to M is 1-derivation D:Ω(A)Ω+1(A){\mathrm{D}}\colon\Omega^{\bullet}(A)\to\Omega^{\bullet+1}(A) such that D2=0{\mathrm{D}}^{2}=0.

1.2. Relative derivations

Fix two vector bundles WMW\to M and VNV\to N together with bundle map (φ,p):WV(\varphi,p)\colon W\to V covering a smooth map p:MNp\colon M\to N.

Definition 1.4.

A kk-derivation relative to φ\varphi is a map

D:Ω(V)Ω+k(W){\mathrm{D}}\colon\Omega^{\bullet}(V)\to\Omega^{\bullet+k}(W)

satisfying the Leibniz rule

D(αβ)=(Dα)(φβ)+(1)|α|k(φα)(Dβ).{\mathrm{D}}(\alpha\wedge\beta)=({\mathrm{D}}\alpha)\wedge(\varphi^{*}\beta)+(-1)^{|\alpha|k}(\varphi^{*}\alpha)\wedge({\mathrm{D}}\beta).

The space of kk-derivations relative to φ\varphi is denoted by Derk(φ)\operatorname{Der}^{k}(\varphi).

Lemma 1.5.

The space Derk(φ)\operatorname{Der}^{k}(\varphi) arises as the space of sections of a vector bundle 𝒟φk\mathcal{D}^{k}_{\varphi} over MM.

Proof.

The space Derk(φ)\operatorname{Der}^{k}(\varphi) is a C(M)C^{\infty}(M)-module. In local trivializations and coordinates, it is clear that this module is also locally finitely generated. The lemma follows from the Serre-Swan theorem. ∎

Note that ordinary derivations are derivations relative to the identity map, so that Derk(V)=Derk(idV)\operatorname{Der}^{k}(V)=\operatorname{Der}^{k}(\operatorname{id}_{V}) and 𝒟Vk=𝒟idVk\mathcal{D}^{k}_{V}=\mathcal{D}^{k}_{\operatorname{id}_{V}}.

Definition 1.6.

The symbol of a derivation of a kk-derivation relative to (φ,p)(\varphi,p) is the map σ(D):pTNkW\sigma({\mathrm{D}})\colon p^{*}T^{*}N\to\wedge^{k}W^{*} defined by

σ(D)x(dp(x)f)=D(f)xkV|x, for fC(N),xM.\sigma({\mathrm{D}})_{x}({\mathrm{d}}_{p(x)}f)={\mathrm{D}}(f)_{x}\in\wedge^{k}V^{*}|_{x},\quad\text{ for }f\in C^{\infty}(N),\ x\in M.

The dual of the symbol, arising from the canonical identification Hom(W,V)Hom(W,V)\operatorname{Hom}(W,V)\cong\operatorname{Hom}(W^{*},V^{*}), is also denoted by σ(D):kWpTN\sigma({\mathrm{D}})\colon\wedge^{k}W\to p^{*}TN. As for ordinary derivations, the symbol induces an exact sequence

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(k+1W,pV)\textstyle{\operatorname{Hom}(\wedge^{k+1}W,p^{*}V)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟φk\textstyle{\mathcal{D}^{k}_{\varphi}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}Hom(kW,pTN)\textstyle{\operatorname{Hom}(\wedge^{k}W,p^{*}TN)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.} (1.3)
Example 1.7.

If D1Derk(W){\mathrm{D}}_{1}\in\operatorname{Der}^{k}(W) and D2Derk(V){\mathrm{D}}_{2}\in\operatorname{Der}^{k}(V) are kk-derivations on WW and VV, respectively, then φD2\varphi^{*}\circ{\mathrm{D}}_{2} and D1φ{\mathrm{D}}_{1}\circ\varphi^{*} are both derivations relative to φ\varphi.

Example 1.8.

Let VNV\to N be a vector bundle and fix xNx\in N. Any element D(𝒟Vk)x{\mathrm{D}}\in(\mathcal{D}^{k}_{V})_{x} determines a kk-derivation relative to the inclusion ιx:VxV\iota_{x}\colon V_{x}\hookrightarrow V:

D:Ω(V)+kVx{\mathrm{D}}\colon\Omega^{\bullet}(V)\to\wedge^{\bullet+k}V_{x}^{*}

It can be defined as follows. Pick any section D~Γ(𝒟Vk)\tilde{{\mathrm{D}}}\in\Gamma(\mathcal{D}^{k}_{V}) with D~x=D\tilde{{\mathrm{D}}}_{x}={\mathrm{D}} and set

Dα=(ιxD~)α.{\mathrm{D}}\alpha=\left(\iota_{x}^{*}\circ\tilde{{\mathrm{D}}}\right)\alpha.

This gives a canonical identification between the space of kk-derivations relative to the inclusion ιx:WxW\iota_{x}\colon W_{x}\hookrightarrow W and the fiber (𝒟Wk)x(\mathcal{D}^{k}_{W})_{x}.

The most important case relevant to Bryant’s equations occurs when VV is a vector bundle over NN, W=pVW=p^{*}V is the pullback of WW along a map p:MNp\colon M\to N and φ=p:pVV\varphi=p_{*}\colon p^{*}V\to V is the canonical projection – see also Remark 1.11 below. In this case, the short exact sequence (1.3) becomes

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pHom(k+1V,V)\textstyle{p^{*}\operatorname{Hom}(\wedge^{k+1}V,V)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟φk\textstyle{\mathcal{D}^{k}_{\varphi}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}pHom(kV,TN)\textstyle{p^{*}\operatorname{Hom}(\wedge^{k}V,TN)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.} (1.4)

The previous sequence suggests the following result.

Lemma 1.9.

Let p:MNp\colon M\to N be a map, VNV\to N a vector bundle and p:pVWp_{*}\colon p_{*}V\to W the projection. Then there is a canonical isomorphism

𝒟pkp𝒟Vk.\mathcal{D}^{k}_{p_{*}}\cong p^{*}\mathcal{D}^{k}_{V}.
Proof.

This follows from the observation that

Derk(p)C(M)C(N)Derk(V)\operatorname{Der}^{k}(p_{*})\cong C^{\infty}(M)\otimes_{C^{\infty}(N)}\operatorname{Der}^{k}(V)

as a C(M)C^{\infty}(M)-module. ∎

1.3. Relative brackets

Relative derivations are in duality with relative brackets, a concept that we will introduce here.

Definition 1.10.

Let VNV\to N be a vector bundle and p:MNp\colon M\to N any smooth map. A kk-bracket relative to pp is a skew-symmetric \mathbb{R}-multilinear map

[,,]:Γ(V)××Γ(V)(k+1)-timesΓ(pV)[\cdot,\dots,\cdot]\colon\underbrace{\Gamma(V)\times\cdots\times\Gamma(V)}_{\text{$(k+1)$-times}}\to\Gamma(p^{*}V)

and a relative anchor ρ:kpVpTN\rho\colon\wedge^{k}p^{*}V\to p^{*}TN satisfying the Leibniz rule

[v0,,fvk]=(pf)[v0,,vk]+ρ(pv0,,pvk1)(f)pvk[v_{0},\dots,fv_{k}]=(p^{*}f)[v_{0},\dots,v_{k}]+\rho(p^{*}v_{0},\dots,p^{*}v_{k-1})(f)p^{*}v_{k}

for all fC(N)f\in C^{\infty}(N).

Remark 1.11.

There seems to be no canonical way to define a bracket relative to an arbitrary bundle map (φ,p):WV(\varphi,p)\colon W\to V because there is no canonical map Γ(V)Γ(W)\Gamma(V)\to\Gamma(W), unless φ\varphi is a fiberwise isomorphism, in which case WW is isomorphic to pVp^{*}V.

Similar to the case of ordinary multiderivations, we have the following correspondence between relative multiderivations and relative multibrackets.

Lemma 1.12.

Let p:MNp\colon M\to N be a map and VNV\to N a vector bundle. There is a 1:1 correspondence between

  1. (i)

    kk-derivations relative to p:pVVp_{*}\colon p^{*}V\to V and

  2. (ii)

    kk-brackets relative to pp.

Proof.

This follows from a modified version of the Koszul formula:

(Df)(pv1,,pvk)\displaystyle({\mathrm{D}}f)(p^{*}v_{1},\dots,p^{*}v_{k}) =ρ(pv1,,pvk)(f),\displaystyle=\rho(p^{*}v_{1},\dots,p^{*}v_{k})(f),
(Dα)(pv0,,pvk)\displaystyle({\mathrm{D}}\alpha)(p^{*}v_{0},\dots,p^{*}v_{k}) =i=0k(1)i+kρ(pv0,,pvi^,,pvk)(α,vi)\displaystyle=\sum_{i=0}^{k}(-1)^{i+k}\rho(p^{*}v_{0},\dots,\widehat{p^{*}v_{i}},\dots,p^{*}v_{k})\left(\langle\alpha,v_{i}\rangle\right)
pα([v0,,vk])\displaystyle\phantom{=}-p^{*}\alpha([v_{0},\dots,v_{k}])

for fC(N)f\in C^{\infty}(N), αΩ1(V)\alpha\in\Omega^{1}(V) and v0,,vkΓ(V)v_{0},\dots,v_{k}\in\Gamma(V). ∎

Remark 1.13.

Thinking of multiderivations relative to pp as multibrackets relative to pp, one sees that each choice of connection \nabla for VV gives a splitting of the sequence (1.4) associated with the anchor. Namely, we have the analogue of formula (1.2): given a relative kk-bracket the expression

[pv0,,pvk]:=[v0,,vk]+(1)k+1p(i=0k(1)iρ(pv0,,pvi^,,pvk)vi),[p^{*}v_{0},\dots,p^{*}v_{k}]_{\nabla}:=[v_{0},\dots,v_{k}]+(-1)^{k+1}p^{*}\Big{(}\sum_{i=0}^{k}(-1)^{i}\nabla_{\rho(p^{*}v_{0},\dots,\widehat{p^{*}v_{i}},\dots,p^{*}v_{k})}v_{i}\Big{)},

defines a unique element of Hom(k+1pV,pV)=pHom(k+1V,V)\operatorname{Hom}(\wedge^{k+1}p^{*}V,p^{*}V)=p^{*}\operatorname{Hom}(\wedge^{k+1}V,V).

1.4. Brackets and derivations relative to foliations

In practice, relative derivations often appear relative to a submersion p:MNp\colon M\to N. In some cases, however, we encounter derivations that are only locally relative to a submersion, i.e., they are relative to a foliation \mathcal{F} of MM. In order to define them properly, we need to recall first some basic notions from foliation theory (see, e.g., [25]).

1.4.1. Foliated flat vector bundles

Henceforth, we will identify a foliation \mathcal{F} on a manifold MM with an involutive subbundle TM\mathcal{F}\subset TM. Associated to \mathcal{F} one has the sheaf of basic functions given by

Cbas(U)={fC(U):X(f)=0 for all XΓ(|U)}.C^{\infty}_{\text{{bas}}}(U)=\{f\in C^{\infty}(U):X(f)=0\text{ for all $X\in\Gamma(\mathcal{F}|_{U})$}\}.

Given a vector bundle WMW\to M, we will write Ωk(;W)\Omega^{k}(\mathcal{F};W) for the foliated kk-forms with values in WW, i.e., sections of kW\wedge^{k}\mathcal{F}^{*}\otimes W. Recall that a \mathcal{F}-connection on WW is a \mathbb{R}-bilinear map ¯:Γ()×Γ(W)Γ(W)\overline{\nabla}\colon\Gamma(\mathcal{F})\times\Gamma(W)\to\Gamma(W) satisfying:

¯fXw=f¯X,¯X(fw)=f¯xw+X(f)w,\overline{\nabla}_{fX}w=f\overline{\nabla}_{X},\quad\overline{\nabla}_{X}(fw)=f\overline{\nabla}_{x}w+X(f)w,

for XΓ()X\in\Gamma(\mathcal{F}), fC(M)f\in C^{\infty}(M) and wΓ(W)w\in\Gamma(W). It can be interpreted as a differential operator ¯:Γ(W)Ω1(;W)\overline{\nabla}\colon\Gamma(W)\to\Omega^{1}(\mathcal{F};W). A section wΓ(W)w\in\Gamma(W) is ¯\overline{\nabla}-parallel when ¯w=0\overline{\nabla}w=0. This gives rise to a sheaf of modules over CbasC^{\infty}_{\text{{bas}}} given by

Γ(W,¯)(U)={wΓW(U):¯w=0}.\Gamma_{(W,\overline{\nabla})}(U)=\{w\in\Gamma_{W}(U):\overline{\nabla}w=0\}.

Local existence of parallel sections is controlled by the curvature of ¯{\overline{\nabla}} which is the the 2-form R¯Ω2(;End(W))R_{\overline{\nabla}}\in\Omega^{2}(\mathcal{F};\operatorname{End}(W)) given by

R¯(X,Y)=[¯X,¯Y]¯[X,Y].R_{\overline{\nabla}}(X,Y)=\left[\overline{\nabla}_{X},\overline{\nabla}_{Y}\right]-\overline{\nabla}_{[X,Y]}.

For a flat \mathcal{F}-connection (i.e. R¯=0R_{\overline{\nabla}}=0), there exist a parallel local section through every point wWw\in W. However, non-zero global parallel sections may not exist.

Definition 1.14.

A foliated vector bundle (W,¯)(W,\overline{\nabla}) is a vector bundle WW over a manifold with foliation (M,)(M,\mathcal{F}) together with an \mathcal{F}-connection ¯\overline{\nabla}. The foliated vector bundle is called flat when the \mathcal{F}-connection is flat.

The dual WW^{*} of a foliated vector vector bundle (W,¯)(W,\overline{\nabla}) carries a canonical \mathcal{F}-connection ¯\overline{\nabla} determined by

Xα,w=¯Xα,w+α,¯Xw,\mathcal{L}_{X}\langle\alpha,w\rangle=\left\langle\overline{\nabla}_{X}\alpha,w\right\rangle+\left\langle\alpha,\overline{\nabla}_{X}w\right\rangle,

for XΓX\in\Gamma_{\mathcal{F}}, αΓW\alpha\in\Gamma_{W^{*}} and wΓWw\in\Gamma_{W}. Note that (W,¯)(W^{*},\overline{\nabla}) is flat iff (W,¯)(W,\overline{\nabla}) is flat.

Next, by a map of foliations p:(M1,1)(M2,2)p\colon(M_{1},\mathcal{F}_{1})\to(M_{2},\mathcal{F}_{2}) we mean a map such that dp(1)2{\mathrm{d}}p(\mathcal{F}_{1})\subset\mathcal{F}_{2}. For example, the identity id:MM\operatorname{id}\colon M\to M maps the trivial foliation to any foliation M\mathcal{F}_{M}. A map of foliated vector bundles φ:(W1,¯1)(W2,¯2)\varphi\colon(W_{1},{\overline{\nabla}}^{1})\to(W_{2},{\overline{\nabla}}^{2}) covering a map of foliations p:(M1,1)(M2,2)p\colon(M_{1},\mathcal{F}_{1})\to(M_{2},\mathcal{F}_{2}) is a bundle map φ:W1W2\varphi\colon W_{1}\to W_{2} which satisfies

¯X1(φα))=φ(¯p(X)2α),for all XΓ1,αΩ1(W2).{\overline{\nabla}}^{1}_{X}(\varphi^{*}\alpha))=\varphi^{*}({\overline{\nabla}}^{2}_{p_{*}(X)}\alpha),\quad\text{for all }X\in\Gamma_{\mathcal{F}_{1}},\alpha\in\Omega^{1}(W_{2}).

In the case of flat foliated bundles, this condition ensures that pullback maps flat forms to flat forms φ:Ω(W2,¯2)Ω(W1,¯1)\varphi^{*}\colon\Omega^{\bullet}_{(W_{2},{\overline{\nabla}}^{2})}\to\Omega^{\bullet}_{(W_{1},{\overline{\nabla}}^{1})}.

Remark 1.15.

When φ\varphi is a fiberwise isomorphism, we obtain also a pullback of sections which preserves flatness:

φ:Γ(W2,¯2)Γ(W1,¯1),(φw)(x):=φ1(w(p(x))).\varphi^{*}\colon\Gamma_{(W_{2},{\overline{\nabla}}^{2})}\to\Gamma_{(W_{1},{\overline{\nabla}}^{1})},\quad(\varphi^{*}w)(x):=\varphi^{-1}(w(p(x))).

There is a version of the Serre-Swan theorem for flat foliated vector bundles.

Proposition 1.16.

Let (M,)(M,\mathcal{F}) be a manifold with foliation. There is, up to a natural isomorphism, a one-to-one correspondence

{ locally finitely generatedand locally free Cbas-modules }1:1{flat foliated vectorbundles over (M,)}\left\{\begin{subarray}{c}\mbox{ locally finitely generated}\\ \\ \mbox{and locally free $C^{\infty}_{\mathrm{bas}}$-modules }\\ \end{subarray}\right\}\overset{1:1}{\longleftrightarrow}\left\{\begin{subarray}{c}\mbox{flat foliated vector}\\ \\ \mbox{bundles over $(M,\mathcal{F})$}\\ \end{subarray}\right\}
Proof.

Given (B,¯)M(B,\overline{\nabla})\to M, we just saw how to construct a locally finitely generated and locally free CbasC^{\infty}_{\text{{bas}}}-module Γ(B,¯)\Gamma_{(B,\overline{\nabla})}. For the converse, given such a module Γ\Gamma_{\parallel}, set

ΓB:=ΓCbasC\Gamma_{B}:=\Gamma_{\parallel}\otimes_{C^{\infty}_{\text{{bas}}}}C^{\infty}

which is a locally finitely generated, locally free CC^{\infty}-module. By the Serre-Swan theorem, ΓB\Gamma_{B} is the sheaf of sections of a vector bundle BMB\to M. The flat \mathcal{F}-connection is determined by requiring its local flat sections to satisfy

¯Xb=0, for all XΓbΓ.\overline{\nabla}_{X}b=0,\quad\text{ for all $X\in\Gamma_{\mathcal{F}}$, $b\in\Gamma_{\parallel}$}.

It is well-defined because Γ\Gamma_{\parallel} is a CbasC^{\infty}_{\text{{bas}}}-module. These constructions are inverses of each other. Naturality follows from naturality of the original Serre-Swan theorem, observing that the resulting maps preserve the parallel sections, and so are maps of flat foliated vector bundles. ∎

Let (W,¯)(W,\overline{\nabla}) be flat foliated vector bundle over a foliated manifold (M,)(M,\mathcal{F}). The holonomy representation of a leaf LL of \mathcal{F} is denoted

h:π1(L,x)GL(Wx)h\colon\pi_{1}(L,x)\to\operatorname{GL}(W_{x})

and is obtained, as usual, by parallel transport along loops based at xx. More generally, parallel transport along paths, gives the groupoid representation

Π1()GL(W),\Pi_{1}(\mathcal{F})\to\operatorname{GL}(W),

where Π1()M\Pi_{1}(\mathcal{F})\rightrightarrows M is the Lie groupoid whose arrows are the leafwise homotopy classes of paths in \mathcal{F}, and GL(W)M\operatorname{GL}(W)\rightrightarrows M is the Lie groupoid whose arrows are the linear isomorphisms between the fibers of WW. Conversely, any such Lie groupoid representation defines a flat \mathcal{F}-connection on BB.

If ~\tilde{\mathcal{F}}\subset\mathcal{F} is subfoliation, a flat \mathcal{F}-connection determines by restriction a flat ~\tilde{\mathcal{F}}-connection and the holonomy representations are related by

Π1(~)\textstyle{\Pi_{1}(\tilde{\mathcal{F}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h~\scriptstyle{\tilde{h}}i\scriptstyle{i_{*}}Π1()\textstyle{\Pi_{1}(\mathcal{F})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h\scriptstyle{h}GL(W)\textstyle{\operatorname{GL}(W)}
Proposition 1.17.

Let (M,)(M,\mathcal{F}) be a foliated manifold and p:MNp\colon M\to N a surjective submersion with connected fibers such that ker(dp)\ker({\mathrm{d}}p)\subset\mathcal{F}. Then there is a foliation N\mathcal{F}_{N} on NN such that =(dp)1(N)\mathcal{F}=({\mathrm{d}}p)^{-1}(\mathcal{F}_{N}) and p:(M,)(N,N)p\colon(M,\mathcal{F})\to(N,\mathcal{F}_{N}) is a map of foliated manifolds, for which there is, up to a natural isomorphism, a one-to-one correspondence between

{vector bundles over N withflat N-connection}1:1{vector bundles over M withflat -connection havingtrivial holonomy along ker(dp)}\left\{\begin{subarray}{c}\mbox{vector bundles over $N$ with}\\ \\ \mbox{flat $\mathcal{F}_{N}$-connection}\end{subarray}\right\}\xleftrightarrow{1:1}\left\{\begin{subarray}{c}\mbox{vector bundles over $M$ with}\\ \\ \mbox{flat $\mathcal{F}$-connection having}\\ \\ \mbox{trivial holonomy along $\ker({\mathrm{d}}p)$}\end{subarray}\right\}
Proof.

Since pp is a surjective submersion and ker(dp)\ker({\mathrm{d}}p)\subset\mathcal{F}, it follows that

N:=dp()\mathcal{F}_{N}:={\mathrm{d}}p(\mathcal{F})

is a well-defined involutive subbundle of TNTN. Since the fibers of pp are connected and contained in \mathcal{F}, one has that =(dp)1(N)\mathcal{F}=({\mathrm{d}}p)^{-1}(\mathcal{F}_{N}).

One direction of the correspondence is clear: since pp is a map of foliations, if (WN,¯N)(N,N)(W_{N},\overline{\nabla}_{N})\to(N,\mathcal{F}_{N}) is a flat foliated vector bundle, then (pWN,p¯N)(M,)(p^{*}W_{N},p^{*}\overline{\nabla}_{N})\to(M,\mathcal{F}) is a vector bundle with flat \mathcal{F}-connection with trivial holonomy along ker(dp)\ker({\mathrm{d}}p).

For the converse, suppose we are given a vector bundle with a flat \mathcal{F}-connection (W,¯)M(W,\overline{\nabla})\to M having trivial holonomy along p:=ker(dp)\mathcal{F}_{p}:=\ker({\mathrm{d}}p). Then the holonomy representation of p\mathcal{F}_{p} factors via the submersion groupoid

Π1(p)\textstyle{\Pi_{1}(\mathcal{F}_{p})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M×pM\textstyle{M\times_{p}M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}GL(W)\textstyle{\operatorname{GL}(W)}

The resulting linear action of M×pMMM\times_{p}M\rightrightarrows M on WMW\to M is free and proper. The quotient is a vector bundle WNNW_{N}\to N and there is a canonical isomorphism

pWNW.p^{*}W_{N}\cong W.

Moreover, the submersion p:MNp\colon M\to N induces a surjective groupoid morphism

p:Π1()Π1(N)p_{*}\colon\Pi_{1}(\mathcal{F})\to\Pi_{1}(\mathcal{F}_{N})

whose kernel contains Π1(p)\Pi_{1}(\mathcal{F}_{p}). It follows that the holonomy representation descends to a representation of Π1(N)\Pi_{1}(\mathcal{F}_{N}) making the following diagram commute.

Π1()\textstyle{\Pi_{1}(\mathcal{F})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h\scriptstyle{h}p\scriptstyle{p_{*}}GL(B)\textstyle{\operatorname{GL}(B)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Π1(N)\textstyle{\Pi_{1}(\mathcal{F}_{N})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}hN\scriptstyle{h_{N}}GL(WN)\textstyle{\operatorname{GL}(W_{N})}

Hence, there is a unique flat N\mathcal{F}_{N}-connection ¯N{\overline{\nabla}}_{N} on WNW_{N} whose holonomy representation is hNh_{N}. For such a connection one has

p¯N=.p^{*}{\overline{\nabla}}_{N}=\nabla.

The previous two constructions are inverse to each other so the result follows. ∎

Example 1.18.

Let \mathcal{F} be any foliation on MM and denote by ν()=TM/\nu(\mathcal{F})=TM/\mathcal{F} the normal bundle of \mathcal{F}. This carries a flat \mathcal{F}-connection, namely the Bott connection

¯XBottY¯:=[X,Y]¯, for XΓ()Y¯ν().\overline{\nabla}^{\text{Bott}}_{X}\overline{Y}:=\overline{[X,Y]},\quad\mbox{ for $X\in\Gamma(\mathcal{F})$, $\overline{Y}\in\nu(\mathcal{F})$}.

When \mathcal{F} is a simple foliation, so that N=M/N=M/\mathcal{F} is a manifold, the resulting foliation on NN is the trivial one: N=0N\mathcal{F}_{N}=0_{N}. The corresponding vector bundle over NN is the tangent bundle TNTN (a flat foliated vector bundle for the trivial foliation).

More generally, any vector bundle VNV\to N is a flat foliated vector bundle for the trivial foliation, so the pullback W:=pVW:=p^{*}V carries a canonical flat \mathcal{F}-connection ¯{\overline{\nabla}}. It is the connection whose local flat sections are the sections of the form pvp^{*}v, with vv any a local section of VV.

1.4.2. Derivations relative to a foliation

Throughout the rest of this section, we fix a flat foliated vector bundle (W,¯)(W,{\overline{\nabla}}) over (M,)(M,\mathcal{F}). The sheaf of sections of WM\wedge^{\bullet}W^{*}\to M (the “WW-forms”) will be denoted by ΩW\Omega^{\bullet}_{W}, whereas the sheaf Ω(W,¯)\Omega^{\bullet}_{(W,\overline{\nabla})} of ¯{\overline{\nabla}}-parallel WW-forms is given by

Ω(W,¯)(U):={αΩW(U):¯α=0}.\Omega^{\bullet}_{(W,\overline{\nabla})}(U):=\{\alpha\in\Omega^{\bullet}_{W}(U):\overline{\nabla}\alpha=0\}.
Definition 1.19.

A parallel kk-derivation of on a flat foliated vector bundle (W,¯)(W,\overline{\nabla}) is a map of sheaves

D:Ω(W,¯)Ω(W,¯)+k{\mathrm{D}}\colon\Omega^{\bullet}_{(W,\overline{\nabla})}\to\Omega^{\bullet+k}_{(W,\overline{\nabla})}

satisfying, for any homogeneous α,βΩ(W,¯)\alpha,\beta\in\Omega^{\bullet}_{(W,\overline{\nabla})}, the Leibniz rule

D(αβ)=(Dα)β+(1)|α|kα(Dβ).{\mathrm{D}}(\alpha\wedge\beta)=({\mathrm{D}}\alpha)\wedge\beta+(-1)^{|\alpha|k}\alpha\wedge({\mathrm{D}}\beta).

The sheaf of parallel kk-derivations is denoted by Derk(W,¯)\operatorname{Der}^{k}_{\parallel}(W,\overline{\nabla}).

Lemma 1.20.

Derk(W,¯)\operatorname{Der}^{k}_{\parallel}(W,\overline{\nabla}) is the sheaf of flat sections of a flat foliated vector bundle, denoted (𝒟(W,¯)k,¯)(\mathcal{D}^{k}_{(W,\overline{\nabla})},\overline{\nabla}).

Proof.

Derk(W,¯)\operatorname{Der}^{k}_{\parallel}(W,\overline{\nabla}) is a locally finitely generated, locally free CbasC^{\infty}_{\text{{bas}}}-module. Thus, by Proposition 1.16, it is the space of parallel sections of a flat foliated vector bundle (𝒟(W,¯)k,¯)(\mathcal{D}^{k}_{(W,\overline{\nabla})},\overline{\nabla}). ∎

We are now ready to introduce the following generalization of derivations relative to submersions to derivations relative to foliations.

Definition 1.21.

A kk-derivation on (W,¯)(W,\overline{\nabla}) relative to \mathcal{F} is a map of sheaves

D:Ω(W,¯)ΩW+k,{\mathrm{D}}\colon\Omega^{\bullet}_{(W,\overline{\nabla})}\to\Omega^{\bullet+k}_{W},

satisfying for homogeneous α,βΩ(W,¯)\alpha,\beta\in\Omega^{\bullet}_{(W,\overline{\nabla})} the Leibniz rule

D(αβ)=(Dα)β+(1)|α|kα(Dβ).{\mathrm{D}}(\alpha\wedge\beta)=({\mathrm{D}}\alpha)\wedge\beta+(-1)^{|\alpha|k}\alpha\wedge({\mathrm{D}}\beta).

Its symbol is the map σ(D)Hom(ν(),kW)\sigma({\mathrm{D}})\in\operatorname{Hom}(\nu^{*}(\mathcal{F}),\wedge^{k}W^{*}) given by

σ(D)(df)=D(f),for fCbas.\sigma({\mathrm{D}})({\mathrm{d}}f)={\mathrm{D}}(f),\quad\text{for $f\in C^{\infty}_{\text{{bas}}}$.}

In practice, we will think of the symbol as a map σ(D):kWν()\sigma({\mathrm{D}})\colon\wedge^{k}W\to\nu(\mathcal{F}) via the usual canonical identification.

Remark 1.22.

The relative kk-derivations just defined are canonically identified with the global sections of the vector bundle 𝒟(W,¯)k\mathcal{D}^{k}_{(W,\overline{\nabla})}, while parallel kk-derivations correspond to parallel sections of 𝒟(W,¯)k\mathcal{D}^{k}_{(W,{\overline{\nabla}})}. The flat \mathcal{F}-connection on sections of 𝒟(W,¯)k\mathcal{D}^{k}_{(W,{\overline{\nabla}})} is given by

(¯XD)α=¯X(Dα), for αΩ(W,¯).({\overline{\nabla}}_{X}{\mathrm{D}})\alpha={\overline{\nabla}}_{X}({\mathrm{D}}\alpha),\quad\mbox{ for $\alpha\in\Omega^{\bullet}_{(W,{\overline{\nabla}})}$}.

Unlike global parallel kk-derivations, which may or may not exist, relative kk-derivations always exist in abundance.

The duality between brackets and derivations relative to submersions extends to foliations.

Definition 1.23.

A kk-bracket on (W,¯)(W,\overline{\nabla}) relative to \mathcal{F} is a skew-symmetric \mathbb{R}-multilinear map of sheaves

[,,]:Γ(W,¯)××ΓW,¯)(k+1)-timesΓW[\cdot,\dots,\cdot]\colon\underbrace{\Gamma_{(W,\overline{\nabla})}\times\cdots\times\Gamma_{W,\overline{\nabla})}}_{\text{$(k+1)$-times}}\to\Gamma_{W}

with a relative anchor ρ:kWν()\rho\colon\wedge^{k}W\to\nu(\mathcal{F}) satisfying the Leibniz rule

[w0,,fwk]=f[w0,,wk]+ρ(w0,,wk1)(f)wk[w_{0},\dots,fw_{k}]=f[w_{0},\dots,w_{k}]+\rho(w_{0},\dots,w_{k-1})(f)w_{k}

for any local flat sections wiw_{i} and basic function fCbasf\in C^{\infty}_{\text{{bas}}}.

In a manner entirely similar to Lemma 1.12, we find:

Lemma 1.24.

Let (W,¯)(W,{\overline{\nabla}}) be a flat foliated vector bundle over (M,)(M,\mathcal{F}). There is a 1:1 correspondence between

  1. (i)

    kk-derivations relative the \mathcal{F} and

  2. (ii)

    kk-brackets relative to \mathcal{F}.

Notice that in the case of a submersion p:MNp\colon M\to N and a vector bundle VNV\to N – see Example 1.18 – these definitions and results specialize to the previous notions of kk-derivation and kk-bracket relative to p:MNp\colon M\to N.

1.4.3. The structure of relative derivations

We start by giving the analogue of the short exact sequence (1.4). The proof is immediate.

Lemma 1.25.

Given a foliated flat bundle (W,¯)(M,)(W,{\overline{\nabla}})\to(M,\mathcal{F}), the symbol map induces a short exact sequence

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(k+1W,W)\textstyle{\operatorname{Hom}(\wedge^{k+1}W,W)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟(W,¯)k\textstyle{\mathcal{D}^{k}_{(W,\overline{\nabla})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}Hom(kW,ν())\textstyle{\operatorname{Hom}(\wedge^{k}W,\nu(\mathcal{F}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0,\textstyle{0,} (1.5)

If we equip these bundles with the induced connections ¯{\overline{\nabla}} and ¯¯Bott\overline{\nabla}\otimes\overline{\nabla}^{\mathrm{Bott}} (for the last term), this is a sequence of flat foliated vector bundles.

An extension of the \mathcal{F}-connection ¯\overline{\nabla} to an ordinary connection \nabla is a connection on BB such that

Xw=¯Xw,for all XΓ(),wΓ(W).\nabla_{X}w=\overline{\nabla}_{X}w,\quad\text{for all }X\in\Gamma(\mathcal{F}),\ w\in\Gamma(W).

It induces a well-defined map

:Γν()×Γ(W,¯)ΓW,[X]w:=Xw.\nabla\colon\Gamma_{\nu(\mathcal{F})}\times\Gamma_{(W,\overline{\nabla})}\to\Gamma_{W},\quad\nabla_{[X]}w:=\nabla_{X}w.

The extension ¯\overline{\nabla} is called \mathcal{F}-parallel when this map takes values in Γ(W,¯)\Gamma_{(W,\overline{\nabla})}. While extensions always exist, parallel extensions are only guaranteed to exist locally.

Extensions of ¯\overline{\nabla} yield splittings of the previous short exact sequence.

Lemma 1.26.

Let (W,¯)(M,)(W,{\overline{\nabla}})\to(M,\mathcal{F}) be flat foliated bundle. An extension \nabla of ¯\overline{\nabla} induces a splitting of the short exact sequence (1.5) so that

𝒟(W,¯)kHom(k+1W,W)Hom(kW,ν()).\mathcal{D}^{k}_{(W,\overline{\nabla})}\cong\operatorname{Hom}(\wedge^{k+1}W,W)\oplus\operatorname{Hom}(\wedge^{k}W,\nu(\mathcal{F})).

If the extension is \mathcal{F}-parallel, then this splitting is an isomorphism of flat foliated vector bundles.

Proof.

For any extension \nabla we can define a splitting 𝒟(W,¯)kHom(k+1W,W)\mathcal{D}^{k}_{(W,\overline{\nabla})}\to\operatorname{Hom}(\wedge^{k+1}W,W) of (1.5) similar to Remark 1.13: given a relative bracket in 𝒟(W,¯)k\mathcal{D}^{k}_{(W,\overline{\nabla})} the expression

[w0,,wk]:=[w0,,wk]+(1)k+1i=0k(1)iρ(w0,,wi^,,wk)wi,[w_{0},\dots,w_{k}]_{\nabla}:=[w_{0},\dots,w_{k}]+(-1)^{k+1}\sum_{i=0}^{k}(-1)^{i}\nabla_{\rho(w_{0},\dots,\widehat{w_{i}},\dots,w_{k})}w_{i},

which is defined for local flat sections wiΓ(W,¯)w_{i}\in\Gamma_{(W,{\overline{\nabla}})}, extends to a unique CC^{\infty}-linear map, determining a bundle map k+1WW\wedge^{k+1}W\to W. When \nabla is an \mathcal{F}-parallel extension, this bundle map sends \mathcal{F}-flat sections to \mathcal{F}-flat sections, so it is map of foliated flat bundles. ∎

Any kk-derivation DΓ(𝒟Wk){\mathrm{D}}\in\Gamma(\mathcal{D}^{k}_{W}) can be restricted to Ω(W,¯)\Omega^{\bullet}_{(W,\overline{\nabla})} to yield a derivation Π(D)\Pi({\mathrm{D}}) relative to \mathcal{F}. This restriction map is CC^{\infty}-linear, so it is induced from a bundle map Π:𝒟Wk𝒟(W,¯)k\Pi\colon\mathcal{D}^{k}_{W}\to\mathcal{D}^{k}_{(W,\overline{\nabla})}.

Lemma 1.27.

There is a short exact sequence

0Hom(kW,)ι𝒟WkΠ𝒟(W,¯)k0.\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 5.5pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&&&\crcr}}}\ignorespaces{\hbox{\kern-5.5pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 29.5pt\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 29.5pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\operatorname{Hom}(\wedge^{k}W,\mathcal{F})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 100.10303pt\raise 4.50694pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.50694pt\hbox{$\scriptstyle{\iota}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 116.3418pt\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 116.3418pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathcal{D}^{k}_{W}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 139.44736pt\raise 5.39166pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.39166pt\hbox{$\scriptstyle{\Pi}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 157.07236pt\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 157.07236pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathcal{D}^{k}_{(W,\overline{\nabla})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 197.80292pt\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 197.80292pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{0}$}}}}}}}\ignorespaces}}}}\ignorespaces.

where the inclusion ι\iota is defined at the level of sections by

ι(ρ)(α)(w1,,wk+l)=1k!l!σSk+l(1)σ(¯ρ(wσ(1),,wσ(k))α)(wσ(k+1),,wσ(k+l)).\iota(\rho)(\alpha)(w_{1},\dots,w_{k+l})=\frac{1}{k!l!}\sum_{\sigma\in S_{k+l}}(-1)^{\sigma}\left(\overline{\nabla}_{\rho(w_{\sigma(1)},\dots,w_{\sigma(k)})}\alpha\right)(w_{\sigma(k+1)},\dots,w_{\sigma(k+l)}).

for αΩl(W)\alpha\in\Omega^{l}(W) and w1,,wk+lΓ(B)w_{1},\dots,w_{k+l}\in\Gamma(B).

Remark 1.28.

The map ι:Hom(kW,)𝒟Wk\iota\colon\operatorname{Hom}(\wedge^{k}W,\mathcal{F})\to\mathcal{D}^{k}_{W} in the previous statement is the unique linear map satisfying

{ι(ρ)(f)=ρ(f), if fC(M),ι(ρ)(α)=0, if αΩ(W,¯).\begin{cases}\iota(\rho)(f)=\rho(f),\quad&\text{ if }f\in C^{\infty}(M),\\ \iota(\rho)(\alpha)=0,\quad&\text{ if }\alpha\in\Omega^{\bullet}_{(W,\overline{\nabla})}.\end{cases}
Proof.

Restriction induces a surjective map of the short exact sequences of the symbols, resulting in the following commutative diagram:

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}K\textstyle{K\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ι\scriptstyle{\iota}Hom(kW,)\textstyle{\operatorname{Hom}(\wedge^{k}W,\mathcal{F})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(k+1W,W)\textstyle{{\operatorname{Hom}(\wedge^{k+1}W,W)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟Wk\textstyle{\mathcal{D}^{k}_{W}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}Π\scriptstyle{\Pi}Hom(kW,TM)\textstyle{\operatorname{Hom}(\wedge^{k}W,TM)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(k+1W,W)\textstyle{\operatorname{Hom}(\wedge^{k+1}W,W)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟(W,¯)k\textstyle{\mathcal{D}^{k}_{(W,\overline{\nabla})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}Hom(kW,ν())\textstyle{\operatorname{Hom}(\wedge^{k}W,\nu(\mathcal{F}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}0\textstyle{0}0\textstyle{0}0\textstyle{0}

It follows that K=Hom(kW,)K=\operatorname{Hom}(\wedge^{k}W,\mathcal{F}) and that ι\iota is as in the statement. ∎

1.5. Morphisms and extensions of relative derivations

In this section, we discuss morphisms of relative derivations, extensions and what it means to “compose” two relative degree 1 derivations.

Definition 1.29.

Let (φ,p):(W,¯M)(V,¯N)(\varphi,p)\colon(W,{\overline{\nabla}}^{M})\to(V,{\overline{\nabla}}^{N}) a map of flat foliated vector bundles covering p:(M,M)(N,N)p\colon(M,\mathcal{F}_{M})\to(N,\mathcal{F}_{N}). We say that a derivation DWΓ(𝒟(W,¯M)k){\mathrm{D}}_{W}\in\Gamma(\mathcal{D}^{k}_{(W,{\overline{\nabla}}^{M})}) is (φ,p)(\varphi,p)-related to a derivation DVΓ(𝒟(V,¯N)k)D_{V}\in\Gamma(\mathcal{D}^{k}_{(V,{\overline{\nabla}}^{N})}) if

DWφ=φDV.{\mathrm{D}}_{W}\circ\varphi^{*}=\varphi^{*}\circ{\mathrm{D}}_{V}. (1.6)

Note that (1.6) implies that the anchors of DW{\mathrm{D}}_{W} and DV{\mathrm{D}}_{V} are related by

ρVkφ=dφρW.\rho_{V}\circ\wedge^{k}\varphi={\mathrm{d}}\varphi\circ\rho_{W}.

The relation between the associated kk-brackets is more complicated to express since, in general, there is no map relating sections of WW and sections of VV. However, when φ\varphi is a fiberwise isomorphism, as in Remark 1.15, one can express relation (1.6) in terms of kk-brackets as

[φv0,,φvk]W=φ[v0,,vk]V, for all viΓ(V,¯V).[\varphi^{*}v_{0},\dots,\varphi^{*}v_{k}]_{W}=\varphi^{*}[v_{0},\dots,v_{k}]_{V},\quad\text{ for all }v_{i}\in\Gamma_{(V,{\overline{\nabla}}^{V})}.

Actually, when φ\varphi is a fiberwise isomorphism, it induces a bundle map

φ:𝒟(W,¯W)k𝒟(V,¯V)k\varphi_{*}\colon\mathcal{D}^{k}_{(W,{\overline{\nabla}}^{W})}\to\mathcal{D}^{k}_{(V,{\overline{\nabla}}^{V})}

as follows. Recall that an element Dx{\mathrm{D}}_{x} in the fiber (𝒟(W,¯W)k)x(\mathcal{D}^{k}_{(W,{\overline{\nabla}}^{W})})_{x} can be regarded as a derivation Dx:Ω(W,¯W)(+kW)x{\mathrm{D}}_{x}\colon\Omega^{\bullet}_{(W,{\overline{\nabla}}^{W})}\to\big{(}\wedge^{\bullet+k}W^{*}\big{)}_{x} relative to the inclusion (as a map of foliated vector bundles). We define φ(Dx)\varphi_{*}({\mathrm{D}}_{x}) by

φ(Dx):=φDxφ:Ω(V,¯V)(+kV)p(x),\varphi_{*}({\mathrm{D}}_{x}):=\varphi_{*}\circ{\mathrm{D}}_{x}\circ\varphi^{*}\colon\Omega^{\bullet}_{(V,{\overline{\nabla}}^{V})}\to\left(\wedge^{\bullet+k}V^{*}\right)_{p(x)},

where φ=(φ)1:(W)x(V)p(x)\varphi_{*}=(\varphi^{*})^{-1}\colon\big{(}\wedge^{\bullet}W^{*}\big{)}_{x}\to\big{(}\wedge^{\bullet}V^{*}\big{)}_{p(x)}. Then, a derivation DW{\mathrm{D}}_{W} is (φ,p)(\varphi,p)-related to a derivation DV{\mathrm{D}}_{V} if and only if the outside square in the following diagram commutes.

𝒟(W,¯W)k\textstyle{\mathcal{D}^{k}_{(W,{\overline{\nabla}}^{W})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}φ\scriptstyle{\varphi_{*}}𝒟(V,¯V)k\textstyle{\mathcal{D}^{k}_{(V,{\overline{\nabla}}^{V})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}DW\scriptstyle{{\mathrm{D}}_{W}}N\textstyle{N\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}DV\scriptstyle{{\mathrm{D}}_{V}}
Example 1.30 (Extensions of relative derivations).

If DW{\mathrm{D}}_{W} is (φ,p)(\varphi,p)-related to D{\mathrm{D}} and the base map pp is a submersion, we say that DW{\mathrm{D}}_{W} is an extension of DV{\mathrm{D}}_{V}. In particular, we have the following:

  1. (1)

    An extension of a derivation DΓ(𝒟(V,¯)k){\mathrm{D}}\in\Gamma(\mathcal{D}^{k}_{(V,{\overline{\nabla}})}) to a derivation D1Γ(𝒟Vk){\mathrm{D}}_{1}\in\Gamma(\mathcal{D}^{k}_{V}) is an ordinary derivation D1{\mathrm{D}}_{1} that agrees with D{\mathrm{D}} on the flat forms:

    D1α=Dα, for all αΩ(V,¯).{\mathrm{D}}_{1}\alpha={\mathrm{D}}\alpha,\quad\text{ for all }\alpha\in\Omega^{\bullet}_{(V,{\overline{\nabla}})}.

    This fits into the setting of Definition 1.29 by viewing the identity map id:(N,0N)(N,)\operatorname{id}\colon(N,0_{N})\to(N,\mathcal{F}) as a map of foliated manifolds.

  2. (2)

    More generally, given a a submersion p:MNp\colon M\to N, an extension of a derivation DΓ(𝒟(V,¯)k){\mathrm{D}}\in\Gamma(\mathcal{D}^{k}_{(V,{\overline{\nabla}})}) to a derivation D1Γ(𝒟pk){\mathrm{D}}_{1}\in\Gamma(\mathcal{D}^{k}_{p}), is a derivation D1{\mathrm{D}}_{1} on VV relative to p:MNp\colon M\to N satisfying:

    D1α=p(Dα), for all αΩ(V,¯).{\mathrm{D}}_{1}\alpha=p^{*}({\mathrm{D}}\alpha),\quad\text{ for all }\alpha\in\Omega^{\bullet}_{(V,{\overline{\nabla}})}.

Extensions will be important in the theory of relative algebroids, developed in Section 2.

Given two ordinary derivations, their composition usually fails to be a derivation. An important exception occurs in the case of a 1-derivation DΓ(𝒟V1){\mathrm{D}}\in\Gamma(\mathcal{D}^{1}_{V}): the square D2:=DD{\mathrm{D}}^{2}:=D\circ D is a 2-derivation. It is easy to see that if D{\mathrm{D}} has symbol ρ\rho and associated 1-bracket [,][\cdot,\cdot] then D2{\mathrm{D}}^{2} has symbol

ρD2(v1,v2)(f)=ρ(v1)(ρ(v2)(f))ρ(v2)(ρ(v1)(f))ρ([v1,v2])(f),\rho_{{\mathrm{D}}^{2}}(v_{1},v_{2})(f)=\rho(v_{1})\big{(}\rho(v_{2})(f)\big{)}-\rho(v_{2})\big{(}\rho(v_{1})(f)\big{)}-\rho([v_{1},v_{2}])(f),

while the associated 2-bracket is the Jacobiator

[v1,v2,v3]D2=[v1,[v2,v3]]+[v2,[v3,v1]]+[v3,[v1,v2]].[v_{1},v_{2},v_{3}]_{{\mathrm{D}}^{2}}=[v_{1},[v_{2},v_{3}]]+[v_{2},[v_{3},v_{1}]]+[v_{3},[v_{1},v_{2}]].

This generalizes to relative 1-derivations which are extensions, as in Example 1.30. For the statement, note that under the assumptions of that example, the foliation \mathcal{F} pulls back to a foliation pp^{*}\mathcal{F} on MM, and the bundle pVp^{*}V inherits a flat pp^{*}\mathcal{F}-connection p¯p^{*}{\overline{\nabla}} from (V,¯)(V,\overline{\nabla}). Pullback gives a canonical isomorphism between the space of flat forms

Ω(pV,¯)pΩ(V,¯)\Omega^{\bullet}_{(p^{*}V,\overline{\nabla})}\cong p^{*}\Omega^{\bullet}_{(V,\overline{\nabla})}
Lemma 1.31.

If a derivation D1Γ(𝒟(pV,¯)1){\mathrm{D}}_{1}\in\Gamma(\mathcal{D}^{1}_{(p^{*}V,{\overline{\nabla}})}) extends a derivation D0Γ(𝒟(V,¯)1){\mathrm{D}}_{0}\in\Gamma(\mathcal{D}^{1}_{(V,{\overline{\nabla}})}), their composition

D1D0:Ω(pV,¯)Ω+2(pV),{\mathrm{D}}_{1}\circ{\mathrm{D}}_{0}\colon\Omega^{\bullet}_{(p^{*}V,{\overline{\nabla}})}\to\Omega^{\bullet+2}(p^{*}V),

is a relative 2-derivation with symbol and 2-bracket given by

ρD1D0(pv1,pv2)(f)\displaystyle\rho_{{\mathrm{D}}_{1}\circ{\mathrm{D}}_{0}}(p^{*}v_{1},p^{*}v_{2})(f) =ρ1(pv1)(ρ0(v2)(f))ρ1(pv2)(ρ0(v1)(f))\displaystyle=\rho_{1}(p^{*}v_{1})\big{(}\rho_{0}(v_{2})(f)\big{)}-\rho_{1}(p^{*}v_{2})\big{(}\rho_{0}(v_{1})(f)\big{)} (1.7)
ρ1(p[v1,v2]0)(f),\displaystyle\phantom{=}-\rho_{1}(p^{*}[v_{1},v_{2}]_{0})(f),
[pv1,pv2,pv3]D1D0\displaystyle[p^{*}v_{1},p^{*}v_{2},p^{*}v_{3}]_{{\mathrm{D}}_{1}\circ{\mathrm{D}}_{0}} =[v1,[v2,v3]0]1+[v2,[v3,v1]0]1+[v3,[v1,v2]0]1,\displaystyle=[v_{1},[v_{2},v_{3}]_{0}]_{1}+[v_{2},[v_{3},v_{1}]_{0}]_{1}+[v_{3},[v_{1},v_{2}]_{0}]_{1},

for any v1,v2,v3Γ(V,¯)v_{1},v_{2},v_{3}\in\Gamma_{(V,{\overline{\nabla}})} and fCbasf\in C^{\infty}_{\text{{bas}}}.

Proof.

The fact that D1D0{\mathrm{D}}_{1}\circ{\mathrm{D}}_{0} is a 22-derivation follows from

D1D0(αβ)\displaystyle{\mathrm{D}}_{1}\circ{\mathrm{D}}_{0}(\alpha\wedge\beta) =D1(D0αβ+(1)|α|αD0β)\displaystyle={\mathrm{D}}_{1}({\mathrm{D}}_{0}\alpha\wedge\beta+(-1)^{|\alpha|}\alpha\wedge{\mathrm{D}}_{0}\beta)
=D1(D0α)pβ+(1)|α|+1pD0αD1β+\displaystyle={\mathrm{D}}_{1}({\mathrm{D}}_{0}\alpha)\wedge p^{*}\beta+(-1)^{|\alpha|+1}p^{*}{\mathrm{D}}_{0}\alpha\wedge{\mathrm{D}}_{1}\beta+
+(1)|α|D1αpD0β+(1)2|α|pαD1(D0β)\displaystyle\phantom{=}+(-1)^{|\alpha|}{\mathrm{D}}_{1}\alpha\wedge p^{*}{\mathrm{D}}_{0}\beta+(-1)^{2|\alpha|}p^{*}\alpha\wedge{\mathrm{D}}_{1}({\mathrm{D}}_{0}\beta)
=(D1D0)(α)pβ+pα(D1D0)(β),\displaystyle=({\mathrm{D}}_{1}\circ{\mathrm{D}}_{0})(\alpha)\wedge p^{*}\beta+p^{*}\alpha\wedge({\mathrm{D}}_{1}\circ{\mathrm{D}}_{0})(\beta),

where we used the extension property D1α=p(D0α){\mathrm{D}}_{1}\alpha=p^{*}({\mathrm{D}}_{0}\alpha) to cancel two terms. The expressions for the symbol and 2-bracket follow from straightforward computations using the formulas for the duality (see the proof of Lemma 1.12). ∎

1.6. Tableaux of derivations

Tableaux are extremely useful gadgets in the theory of PDEs [9, 8, 30]. They codify higher order consequences of a set of partial differential equations and, when a tableau is involutive, provides a measure of the size of the space of local solutions.

Let W,VW,V be finite-dimensional vector spaces. Classically, a tableau is a subspace 𝒯Hom(W,V)\mathcal{T}\subset\operatorname{Hom}(W,V) or, more generally, a linear map τ:𝒯Hom(W,V)\tau\colon\mathcal{T}\to\operatorname{Hom}(W,V). In practice, a PDE does not come with a single tableau but with a family of tableaux depending on coordinates, i.e., a vector bundle. The theory of tableaux is pointwise in nature, and thus naturally carries over to vector bundles (possibly under some extra constant rank assumption). Associated to a classical tableau, one has the notions of prolongation, the Spencer complex, involutivity and of Cartan characters. We refer to [8, 9, 30] for more details.

It turns out that tableaux associated to classification problems as in the Introduction (or the applications of Theorem 4 in Bryant [8]) are not tableaux in the classical sense. Bryant was certainly aware of this, as is evident from his computations in [8] of the Cartan characters and prolongations. However, to the authors’ knowledge, this is not formalized or mentioned anywhere in the literature. In this section, we will extend the classical notion of tableau by formalizing the notion of a tableau of derivations.

1.6.1. Definition of tableau of derivations

In what follows, we fix vector bundles WMW\to M and VNV\to N and a vector bundle map (φ,p):WV(\varphi,p)\colon W\to V.

Definition 1.32.

A tableau of kk-derivations relative to a vector bundle map φ\varphi is a vector subbundle 𝒯𝒟φk\mathcal{T}\subset\mathcal{D}^{k}_{\varphi}.

Remark 1.33 (Tableaux for derivations relative to foliations).

All the definitions and results that follow apply equally well to bundles 𝒟(W,¯)k\mathcal{D}^{k}_{(W,\overline{\nabla})} of derivations relative to foliations. This is due to the pointwise nature of the operations and notions related to tableaux and the fact that 𝒟(W,¯)k\mathcal{D}^{k}_{(W,\overline{\nabla})} is locally isomorphic to a space of derivations relative to a bundle map.

The theory of prolongations and involutivity of tableaux of derivations, rests upon the following definition.

Definition 1.34.

The Spencer differential δ:Hom(lW,𝒟φk)𝒟φk+l\delta\colon\operatorname{Hom}(\wedge^{l}W,\mathcal{D}^{k}_{\varphi})\to\mathcal{D}^{k+l}_{\varphi} is the unique C(M)C^{\infty}(M)-linear map such that

δ(ωD)=ωD,\delta(\omega\otimes{\mathrm{D}})=\omega\wedge{\mathrm{D}},

for any ωΩl(W)\omega\in\Omega^{l}(W) and DΓ(𝒟φk){\mathrm{D}}\in\Gamma(\mathcal{D}^{k}_{\varphi}).

Remark 1.35.

Sometimes, instead of a subbundle 𝒯𝒟φk\mathcal{T}\subset\mathcal{D}^{k}_{\varphi}, we need to consider a vector bundle map τ:𝒯𝒟φk\tau\colon\mathcal{T}\to\mathcal{D}^{k}_{\varphi} that is not necessarily injective. For such a bundle map, the Spencer differential δτ:Hom(lW,𝒯)𝒟φk+l\delta_{\tau}\colon\operatorname{Hom}(\wedge^{l}W,\mathcal{T})\to\mathcal{D}^{k+l}_{\varphi} is defined by requiring

δτ(ωD)=ωτ(D),\delta_{\tau}(\omega\otimes{\mathrm{D}})=\omega\wedge\tau({\mathrm{D}}),

for ωDΓ(lW𝒯)\omega\otimes{\mathrm{D}}\in\Gamma(\wedge^{l}W^{*}\otimes\mathcal{T}). For a classical tableau, such objects were considered in [30] under the name “generalized tableaux”. In this paper, we will refer to the bundle map τ\tau as a tableau map.

An element in Γ(Hom(lW,𝒟φk))\Gamma\left(\operatorname{Hom}(\wedge^{l}W,\mathcal{D}^{k}_{\varphi})\right) can be viewed as WW-form of degree ll with values in 𝒟φk\mathcal{D}^{k}_{\varphi}, so it is determined by its action on Ω1(V)\Omega^{1}(V). For ξΓ(Hom(lW,𝒟φk))\xi\in\Gamma\left(\operatorname{Hom}(\wedge^{l}W,\mathcal{D}^{k}_{\varphi})\right), the Spencer differential δξ\delta\xi is the (k+l)(k+l)-derivation relative to φ\varphi acting on αΩ1(V)\alpha\in\Omega^{1}(V) as

(δξ)(α)(w1,,wk+l+1)==1(k+1)!l!σSk+l+1(1)σ(ξ(α)(wσ(1),,wσ(l)))(wσ(l+1),,wσ(l+k+1))\begin{split}\left(\delta\xi\right)&(\alpha)(w_{1},\dots,w_{k+l+1})=\\ &=\frac{1}{(k+1)!l!}\sum_{\sigma\in S_{k+l+1}}(-1)^{\sigma}\left(\xi(\alpha)(w_{\sigma(1)},\dots,w_{\sigma(l)})\right)(w_{\sigma(l+1)},\dots,w_{\sigma(l+k+1)})\end{split} (1.8)

where w1,,wk+l+1Γ(W)w_{1},\dots,w_{k+l+1}\in\Gamma(W). Its symbol acts on a function fC(N)f\in C^{\infty}(N) by

(δξ)(f)(w1,,wk+l)=1k!l!σSk+l(1)σ(ξ(f)(wσ(1),,wσ(l)))(wσ(l+1),,wσ(l+k))\begin{split}&\left(\delta\xi\right)(f)(w_{1},\dots,w_{k+l})\\ &=\frac{1}{k!l!}\sum_{\sigma\in S_{k+l}}(-1)^{\sigma}\left(\xi(f)(w_{\sigma(1)},\dots,w_{\sigma(l)})\right)(w_{\sigma(l+1)},\dots,w_{\sigma(l+k)})\end{split} (1.9)
Example 1.36.

There is more than one way of interpreting a classical tableau bundle as a tableau of derivations.

One direct way is through 0-derivations. Let φ:WV\varphi\colon W\to V be any bundle map covering the identity on MM (e.g., it can be the zero map). From the short exact sequence (1.3), it follows that Hom(W,V)𝒟φ0\operatorname{Hom}(W,V)\subset\mathcal{D}^{0}_{\varphi}. Hence, a classical tableau bundle 𝒯Hom(W,V)\mathcal{T}\subset\operatorname{Hom}(W,V) can be seen as a 0-tableau in 𝒟φ0\mathcal{D}^{0}_{\varphi}.

Another interpretation of a classical tableau as a bundle of derivations is through relative 1-derivations. Assume that the bundle map (φ,p):WV(\varphi,p)\colon W\to V is a fiberwise isomorphism. By Lemma 1.27, there is a short exact sequence

0Hom(W,ker(dp))𝒟W1φp𝒟V10.\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 5.5pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&&&\crcr}}}\ignorespaces{\hbox{\kern-5.5pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 29.5pt\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 29.5pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\operatorname{Hom}(W,\ker({\mathrm{d}}p))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 133.44803pt\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 133.44803pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathcal{D}^{1}_{W}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 157.55516pt\raise 5.1875pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-0.8264pt\hbox{$\scriptstyle{\varphi_{*}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 173.88693pt\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 173.88693pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{p^{*}\mathcal{D}^{1}_{V}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 222.15707pt\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 222.15707pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{0}$}}}}}}}\ignorespaces}}}}\ignorespaces.

A subbundle 𝒯Hom(W,ker(dp))\mathcal{T}\subset\operatorname{Hom}(W,\ker({\mathrm{d}}p)) is both a tableau in the classical sense as well as a tableau of 1-derivations.

In both cases, the Spencer differential on Hom(W,𝒯)\operatorname{Hom}(W,\mathcal{T}), interpreted as a tableau of derivations, corresponds to the classical Spencer differential for 𝒯\mathcal{T} as a classical tableau. Therefore, both interpretations of 𝒯\mathcal{T} as a tableau of derivations recover the usual prolongation and Spencer cohomology theory.

Remark 1.37.

There is a warning: while naïvely the subbundle 𝒯Hom(2W,V)\mathcal{T}\subset\operatorname{Hom}(\wedge^{2}W,V) is a classical tableau (in the sense that it is a subbundle of a Hom\operatorname{Hom}-space), it really depends on the context whether it should be treated as such. Usually, the exterior power 2W\wedge^{2}W indicates the presence of brackets. Interpreting Hom(2W,V)\operatorname{Hom}(\wedge^{2}W,V) as a classical tableau leads to different Cartan characters and prolongations. For instance, to compute the characters of 𝒯\mathcal{T} as a classical tableau, a flag of 2W\wedge^{2}W must be used. However, as we shall see, in the derivation picture one only requires a flag of WW.

1.6.2. Spencer cohomology

Given a tableau of derivations we define its prolongation as follows.

Definition 1.38.

The first prolongation of a tableau 𝒯𝒟φk\mathcal{T}\subset\mathcal{D}^{k}_{\varphi} is defined as

𝒯(1)=kerδ|Hom(W,𝒯).\mathcal{T}^{(1)}=\ker\delta\big{|}_{\operatorname{Hom}(W,\mathcal{T})}.

The prolongation of a tableau 𝒯𝒟φk\mathcal{T}\subset\mathcal{D}^{k}_{\varphi} is a subbundle 𝒯(1)Hom(W,𝒯)\mathcal{T}^{(1)}\subset\operatorname{Hom}(W,\mathcal{T}) whenever it has constant rank. Hence, in this case, 𝒯(1)\mathcal{T}^{(1)} is a classical tableau and one defines the higher prolongations of 𝒯\mathcal{T}, when they exist, recursively as

𝒯(m):=(𝒯(m1)(1).\mathcal{T}^{(m)}:=(\mathcal{T}^{(m-1})^{(1)}.

The Spencer differentials

δ:Hom(lW,𝒯(m))Hom(l+1W,𝒯(m1))\delta\colon\operatorname{Hom}(\wedge^{l}W,\mathcal{T}^{(m)})\to\operatorname{Hom}(\wedge^{l+1}W,\mathcal{T}^{(m-1)})

are defined for m1m\geq 1 as in Definition 1.34 by restricting to Hom(lW,𝒯(m))\operatorname{Hom}(\wedge^{l}W,\mathcal{T}^{(m)}) (i.e., regarding 𝒯(m)\mathcal{T}^{(m)} as classical tableau, as in Example 1.36). It is clear from the definition that the Spencer differentials square to zero.

Definition 1.39.

The Spencer cohomologies of a tableau 𝒯𝒟φk\mathcal{T}\subset\mathcal{D}^{k}_{\varphi} are defined as

Hm,l(𝒯):=kerδ:Hom(lW,𝒯(m))Hom(l+1W,𝒯(m1))imδ:Hom(l1W,𝒯(m+1))Hom(lW,𝒯(m))H^{m,l}(\mathcal{T}):=\frac{\ker\delta\colon\operatorname{Hom}(\wedge^{l}W,\mathcal{T}^{(m)})\to\operatorname{Hom}(\wedge^{l+1}W,\mathcal{T}^{(m-1)})}{\operatorname{im}\delta\colon\operatorname{Hom}(\wedge^{l-1}W,\mathcal{T}^{(m+1)})\to\operatorname{Hom}(\wedge^{l}W,\mathcal{T}^{(m)})}

for m1m\geq 1. In the special case m=0m=0, one sets

H0,l(𝒯):=kerδ:Hom(lW,𝒯)𝒟φk+limδ:Hom(l1W,𝒯(1))Hom(lW,𝒯)H^{0,l}(\mathcal{T}):=\frac{\ker\delta\colon\operatorname{Hom}(\wedge^{l}W,\mathcal{T})\to\mathcal{D}^{k+l}_{\varphi}}{\operatorname{im}\delta\colon\operatorname{Hom}(\wedge^{l-1}W,\mathcal{T}^{(1)})\to\operatorname{Hom}(\wedge^{l}W,\mathcal{T})}

and for m=1m=-1, one defines

H1,l(𝒯):=𝒟φk+limδ:Hom(lW,𝒟φk)𝒟φk+l.H^{-1,l}(\mathcal{T}):=\frac{\mathcal{D}^{k+l}_{\varphi}}{\operatorname{im}\delta\colon\operatorname{Hom}(\wedge^{l}W,\mathcal{D}^{k}_{\varphi})\to\mathcal{D}^{k+l}_{\varphi}}.
Remark 1.40.

Note that Hm,0(𝒯)=Hm,1(𝒯)=0H^{m,0}(\mathcal{T})=H^{m,1}(\mathcal{T})=0 from the definition of the first prolongation.

Definition 1.41.

A tableau 𝒯𝒟φk\mathcal{T}\subset\mathcal{D}^{k}_{\varphi} is called involutive if all prolongations 𝒯(m)\mathcal{T}^{(m)}, m1m\geq 1, have locally constant rank and

Hm,l(𝒯)=0, for all m0,l1.H^{m,l}(\mathcal{T})=0,\quad\mbox{ for all $m\geq 0,l\geq 1$}.

1.6.3. Cartan characters and Cartan’s test

Heuristically, a system of differential equations is in involution when there are no higher order hidden consequences of the equations. These higher order consequences appear as cohomology classes in the Spencer cohomology groups which, in general, are hard to compute. A practical way of checking involutivity is through Cartan’s test. In this section, we make these notions precise for tableaux of derivations.

We continue to assume that (φ,p):WV(\varphi,p)\colon W\to V is a fixed morphism of vector bundles. By a flag of WW we mean sequence of vector subbundles of WW,

0=W0W1Wn=W, with rankWi=i.0=W_{0}\subset W_{1}\subset\dots\subset W_{n}=W,\quad\text{ with }\operatorname{rank}W_{i}=i.

Fix a flag (Wi)(W_{i}) for WW, denote by ιi:WiW\iota_{i}\colon W_{i}\hookrightarrow W the inclusion map and ιi:WWi\iota_{i}^{*}\colon\wedge^{\bullet}W^{*}\to\wedge^{\bullet}W_{i}^{*} the induced restrictions. A relative kk-derivation D:Ω(V)Ω+k(W){\mathrm{D}}\colon\Omega^{\bullet}(V)\to\Omega^{\bullet+k}(W) can be post-composed with the restriction map to yield an kk-derivation ιiD\iota_{i}^{*}\circ{\mathrm{D}} relative to φιi\varphi\circ\iota_{i} (cf. Example 1.7). In other words, there is restriction map

(ιi):𝒟φk𝒟φιlk.(\iota_{i})_{*}\colon\mathcal{D}^{k}_{\varphi}\to\mathcal{D}^{k}_{\varphi\circ\iota_{l}}.

If 𝒯𝒟φk\mathcal{T}\subset\mathcal{D}^{k}_{\varphi} is a tableau, we set

𝒯i:=ker(ιi)𝒯.\mathcal{T}_{i}:=\ker(\iota_{i})_{*}\cap\mathcal{T}.

Notice that 𝒯i𝒯i1\mathcal{T}_{i}\subseteq\mathcal{T}_{i-1} and that 𝒯0=𝒯\mathcal{T}_{0}=\mathcal{T}, 𝒯n=0\mathcal{T}_{n}=0 where n=rankEn=\operatorname{rank}E.

Definition 1.42.

Let 𝒯𝒟φk\mathcal{T}\subset\mathcal{D}^{k}_{\varphi} be a tableau and (Wi)(W_{i}) a flag for WW. The Cartan characters of 𝒯\mathcal{T} with respect to the flag are

si:=rank𝒯i1rank𝒯i,(i=1,,n=rankW).s_{i}:=\operatorname{rank}\mathcal{T}_{i-1}-\operatorname{rank}\mathcal{T}_{i},\quad(i=1,\dots,n=\operatorname{rank}W).

As for classical tableaux, the Cartan characters bound the rank of the prolongation and, moreover, provide a practical way to verify that a tableau of derivations is involutive through Cartan’s test. Here we present an extension of Cartan’s test to tableaux of derivations.

Theorem 1.43 (Cartan’s test).

Let 𝒯𝒟φk\mathcal{T}\subset\mathcal{D}^{k}_{\varphi} be a tableau of derivations.

  1. (i)

    (Cartan’s bound). If {si}\{s_{i}\} are the Cartan characters w.r.t. a flag (Wi)(W_{i}), the dimension of the prolongation is constrained by

    rank𝒯(1)s1+2s2++nsn.\operatorname{rank}\mathcal{T}^{(1)}\leq s_{1}+2s_{2}+\dots+ns_{n}.
  2. (ii)

    (Cartan’s test). If the Cartan characters sis_{i} are locally constant and Cartan’s bound is achieved, i.e.,

    rank𝒯(1)=s1+2s2++nsn,\operatorname{rank}\mathcal{T}^{(1)}=s_{1}+2s_{2}+\dots+ns_{n},

    then 𝒯\mathcal{T} is involutive.

A flag for which Cartan’s test holds is called a regular flag for 𝒯\mathcal{T}. From the proof below, it will be clear that if (Wi)(W_{i}) is a regular flag for 𝒯\mathcal{T}, then it is also a regular flag for the first prolongation 𝒯(1)\mathcal{T}^{(1)}.

Remark 1.44.

For a classical tableau, Cartan’s test provides an equivalent characterization of involutivity (see [31], Theorem 3.4). We suspect that it is also an equivalence for tableau of derivations, but it does not seem to follow from Cartan’s test applied to the prolongation as a classical tableau. The reason is that if 𝒯\mathcal{T} is a tableau, regular flags for the prolongation 𝒯(1)\mathcal{T}^{(1)} may not be regular for 𝒯\mathcal{T} itself. This poses no obstacle to our applications: computing the Cartan characters is typically much more practical than computing the Spencer cohomology groups, making the implication in Theorem 1.43 the most relevant one.

Before we give the proof of the previous theorem, we make the following observation. Since the first prolongation 𝒯(1)Hom(W,𝒯)\mathcal{T}^{(1)}\subset\operatorname{Hom}(W,\mathcal{T}) is also a tableau, it comes with spaces (𝒯(1))i(\mathcal{T}^{(1)})_{i}, associated to a choice of flag for WW. These are related to the prolongations of 𝒯i\mathcal{T}_{i} as follows.

Lemma 1.45.

Let 𝒯𝒟φk\mathcal{T}\subset\mathcal{D}^{k}_{\varphi} be a tableau of derivations and fix a flag (Wi)(W_{i}) for WW. Then:

(𝒯(1))i(𝒯i)(1),(i=1,,n).\left(\mathcal{T}^{(1)}\right)_{i}\subseteq(\mathcal{T}_{i})^{(1)},\quad(i=1,\dots,n).
Proof.

Let ξ(𝒯(1))i\xi\in(\mathcal{T}^{(1)})_{i}, so ξ(w)=0\xi(w)=0 for all wWiw\in W_{i}. Then, for any αΩ1(V)\alpha\in\Omega^{1}(V), uWu\in W and w0,,wkWiw_{0},\dots,w_{k}\in W_{i}, applying (1.8) one finds

0\displaystyle 0 =(δξ)(α)(u,w0,,wk)\displaystyle=(\delta\xi)(\alpha)(u,w_{0},\dots,w_{k})
=(ξ(u)α)(w0,,wk)i=0k(1)i(ξ(wi)α)(u,w0,,wi^,,wk)\displaystyle=\left(\xi(u)\alpha\right)(w_{0},\dots,w_{k})-\sum_{i=0}^{k}(-1)^{i}\left(\xi(w_{i})\alpha\right)(u,w_{0},\dots,\widehat{w_{i}},\dots,w_{k})
=(ξ(u)α)(w0,,wk).\displaystyle=\left(\xi(u)\alpha\right)(w_{0},\dots,w_{k}).

Therefore, ξ(u)𝒯i\xi(u)\in\mathcal{T}_{i} for all uWu\in W, and so ξ(𝒯i)(1)\xi\in\left(\mathcal{T}_{i}\right)^{(1)}. ∎

Remark 1.46.

For a tableaux 𝒯Hom(W,V)\mathcal{T}\subset\operatorname{Hom}(W,V) in the classical sense, the inclusion in Lemma 1.45 is an equality. However, in our more general case this may fail. Consider, for example, the case where W=VW=V is a 2-dimensional vector space, with flag W0W1W2W_{0}\subset W_{1}\subset W_{2}, and let

𝒯:=𝒟φ1=Hom(2W,W).\mathcal{T}:=\mathcal{D}^{1}_{\varphi}=\operatorname{Hom}(\wedge^{2}W,W).

Then 𝒯1=Hom(2W,W)=𝒯\mathcal{T}_{1}=\operatorname{Hom}(\wedge^{2}W,W)=\mathcal{T}, and so (𝒯1)(1)=𝒯(1)=Hom(W,Hom(2W,W))(\mathcal{T}_{1})^{(1)}=\mathcal{T}^{(1)}=\operatorname{Hom}(W,\operatorname{Hom}(\wedge^{2}W,W)). On the other hand, (𝒯(1))1=Hom(W1,Hom(2W,W))(𝒯1)(1)(\mathcal{T}^{(1)})_{1}=\operatorname{Hom}(W_{1},\operatorname{Hom}(\wedge^{2}W,W))\subsetneq(\mathcal{T}_{1})^{(1)}.

The proof of Theorem 1.43 rests on the following lemma.

Lemma 1.47.

If L:=ker(δ:Hom(lW,𝒯)𝒟φk+l)L:=\ker\left(\delta\colon\operatorname{Hom}(\wedge^{l}W,\mathcal{T})\to\mathcal{D}^{k+l}_{\varphi}\right), then

rankLi=1nsi((nl)(nil))\operatorname{rank}L\leq\sum_{i=1}^{n}s_{i}\left(\binom{n}{l}-\binom{n-i}{l}\right) (1.10)

for any flag (Wi)(W_{i}) of WW.

Proof.

Let (Wi)(W_{i}) be a flag of WW and set

Li:=LHom(lW,𝒯i).L_{i}:=L\cap\operatorname{Hom}(\wedge^{l}W,\mathcal{T}_{i}).

These define a filtration L=L0Li1LiLn=0L=L_{0}\supseteq\dots\supseteq L_{i-1}\supseteq L_{i}\supseteq\dots\supseteq L_{n}=0, and there are natural inclusions

Li1/LiHom(lW,𝒯i1/𝒯i).L_{i-1}/L_{i}\hookrightarrow\operatorname{Hom}\left(\wedge^{l}W,\mathcal{T}_{i-1}/\mathcal{T}_{i}\right).

If πi:WW/Wi\pi_{i}\colon W\to W/W_{i} is the projection, then pullback gives another set of inclusions

πi:Hom(l(W/Wi),𝒯i1/𝒯i)Hom(lW,𝒯i1/𝒯i).\pi_{i}^{*}\colon\operatorname{Hom}\left(\wedge^{l}(W/W_{i}),\mathcal{T}_{i-1}/\mathcal{T}_{i}\right)\to\operatorname{Hom}\left(\wedge^{l}W,\mathcal{T}_{i-1}/\mathcal{T}_{i}\right).
Claim.

(Li1/Li)imπi={0}\left(L_{i-1}/L_{i}\right)\cap\operatorname{im}\pi_{i}^{*}=\{0\}.

To prove this claim let πiξLi1πiHom(2(W/Wi),𝒯i1)\pi_{i}^{*}\xi\in L_{i-1}\cap\pi_{i}^{*}\operatorname{Hom}\left(\wedge^{2}\left(W/W_{i}\right),\mathcal{T}_{i-1}\right). If w1,wlWw_{1},\dots w_{l}\in W and wl+1,,wk+l+1Wiw_{l+1},\dots,w_{k+l+1}\in W_{i}, then by (1.8) for αΩ1(V)\alpha\in\Omega^{1}(V) we have

0\displaystyle 0 =δ(πiξ)(α)(w1,,wk+l+1)\displaystyle=\delta\left(\pi_{i}^{*}\xi\right)(\alpha)(w_{1},\dots,w_{k+l+1})
=1l!(k+1)!σSk+l+1(1)σ(ξ(α)(πi(vσ(1)),πi(wσ(l)))(wl+1,,wσ(k+l))\displaystyle=\frac{1}{l!(k+1)!}\sum_{\sigma\in S_{k+l+1}}(-1)^{\sigma}(\xi(\alpha)\left(\pi_{i}\left(v_{\sigma(1)}\right),\dots\pi_{i}\left(w_{\sigma(l)}\right)\right)\left(w_{l+1},\dots,w_{\sigma(k+l)}\right)
=((πiξ)(α)(w1,,wl))(wl+1,,wk+l+1).\displaystyle=((\pi_{i}^{*}\xi)(\alpha)(w_{1},\dots,w_{l}))(w_{l+1},\dots,w_{k+l+1}).

It follows that πiξHom(lW,𝒯i)\pi_{i}^{*}\xi\in\operatorname{Hom}\left(\wedge^{l}W,\mathcal{T}_{i}\right), and the claim follows.

The estimate of the lemma now follows from this claim by telescoping:

rankL\displaystyle\operatorname{rank}L =i=1nrankLi1rankLi\displaystyle=\sum_{i=1}^{n}\operatorname{rank}L_{i-1}-\operatorname{rank}L_{i}
i=1nrankHom(lW,𝒯i1/𝒯i)rankHom(l(W/Wi),𝒯i1/Ti)\displaystyle\leq\sum_{i=1}^{n}\operatorname{rank}\operatorname{Hom}(\wedge^{l}W,\mathcal{T}_{i-1}/\mathcal{T}_{i})-\operatorname{rank}\operatorname{Hom}(\wedge^{l}(W/W_{i}),\mathcal{T}_{i-1}/T_{i})
=i=1nsi((nl)(nil)).\displaystyle=\sum_{i=1}^{n}s_{i}\left(\binom{n}{l}-\binom{n-i}{l}\right).\qed
Proof of Theorem 1.43.

Note that the case l=1l=1 in Lemma 1.47 is exactly item (i) (Cartan’s bound) in the theorem.

For the proof of item (ii), suppose that (Wi)(W_{i}) is a regular flag for 𝒯\mathcal{T}, with characters (si)(s_{i}), so that Cartan’s test on dim𝒯(1)\dim\mathcal{T}^{(1)} is satisfied. Since the Cartan characters are assumed locally constant, so is the rank of 𝒯(1)\mathcal{T}^{(1)}.

The proof goes in two steps:

Step 1:

Prove the inequality

rank(imδ)i=1nsi((nl)(nil))\operatorname{rank}(\operatorname{im}\delta)\geq\sum_{i=1}^{n}s_{i}\left(\binom{n}{l}-\binom{n-i}{l}\right) (1.11)

for l=2l=2 and that Cartan’s bound is achieved for rank𝒯(2)\operatorname{rank}\mathcal{T}^{(2)} – as the first prolongation of 𝒯(1)\mathcal{T}^{(1)} – for the same flag (Wi)(W_{i}). Then, by Lemma 1.47, it follows that

H0,2(𝒯)=0,H^{0,2}(\mathcal{T})=0,

and that – by Cartan’s test for ordinary tableaux ([31, Theorem 3.4]) – all prolongations 𝒯(m)\mathcal{T}^{(m)}, m1m\geq 1, have locally constant rank and are involutive, so that

Hm,l(𝒯)=Hm1,l(𝒯(1))=0,for m1,l0;H^{m,l}(\mathcal{T})=H^{m-1,l}(\mathcal{T}^{(1)})=0,\quad\text{for }m\geq 1,\,l\geq 0;
Step 2:

Prove inequality (1.11) for arbitrary ll so that, applying Lemma 1.47, one has

H0,l(𝒯)=0,for l0.H^{0,l}(\mathcal{T})=0,\quad\text{for }l\geq 0.

Step 1. First, we claim that the Cartan characters si(1)s^{(1)}_{i} of the first prolongation satisfy the bound:

si(1)si+si+1++sn.s^{(1)}_{i}\leq s_{i}+s_{i+1}+\dots+s_{n}. (1.12)

This follows from the Euler characteristic of the exact sequence

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(𝒯(1))i\textstyle{\left(\mathcal{T}^{(1)}\right)_{i}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(𝒯(1))i1\textstyle{\left(\mathcal{T}^{(1)}\right)_{i-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ιwi\scriptstyle{\iota_{w_{i}}}𝒯i1\textstyle{\mathcal{T}_{i-1}}

where ιwi\iota_{w_{i}} is interior contraction by a local frame (wi)(w_{i}) adapted to (Wi)(W_{i}), i.e., {w1,,wi}\{w_{1},\dots,w_{i}\} is a frame for EiE_{i}.

From the bound on si(1)s_{i}^{(1)} together with Cartan’s bound for rank𝒯(2)\operatorname{rank}\mathcal{T}^{(2)}, we see that

rank𝒯(2)i=1nisi(1)i=1n(i+12)si.\operatorname{rank}\mathcal{T}^{(2)}\leq\sum_{i=1}^{n}is^{(1)}_{i}\leq\sum_{i=1}^{n}\binom{i+1}{2}s_{i}.

Furthermore, from the exact row of the Spencer complex

0𝒯(2)Hom(W,𝒯(1))δHom(2W,𝒯)𝒟φk+2,\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 5.5pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&&&\crcr}}}\ignorespaces{\hbox{\kern-5.5pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 29.5pt\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 29.5pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathcal{T}^{(2)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 73.87778pt\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 73.87778pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\operatorname{Hom}(W,\mathcal{T}^{(1)})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 142.83345pt\raise 5.43056pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.43056pt\hbox{$\scriptstyle{\delta}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 158.81122pt\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 158.81122pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\operatorname{Hom}(\wedge^{2}W,\mathcal{T})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 246.05579pt\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 246.05579pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathcal{D}^{k+2}_{\varphi}}$}}}}}}}\ignorespaces}}}}\ignorespaces,

we conclude that

rank(imδ)=rank(Hom(W,𝒯(1)))rank𝒯(2)i=0nsi(ni(i+12))=i=1nsi((n2)(ni2)),\displaystyle\begin{split}\operatorname{rank}(\operatorname{im}\delta)&=\operatorname{rank}\left(\operatorname{Hom}(W,\mathcal{T}^{(1)})\right)-\operatorname{rank}\mathcal{T}^{(2)}\\ &\geq\sum_{i=0}^{n}s_{i}\left(ni-\binom{i+1}{2}\right)\\ &=\sum_{i=1}^{n}s_{i}\left(\binom{n}{2}-\binom{n-i}{2}\right),\end{split} (1.13)

where we used the assumption that Cartan’s bound is achieved for rank𝒯(1)\operatorname{rank}\mathcal{T}^{(1)}. This proves the inequality (1.11) for l=2l=2. But then by Lemma 1.47, the inequality (1.13) must be an equality, so that

rank𝒯(2)=i=1nsi(i+12).\operatorname{rank}\mathcal{T}^{(2)}=\sum_{i=1}^{n}s_{i}\binom{i+1}{2}.

Hence, Cartan’s bound is achieved for 𝒯(2)\mathcal{T}^{(2)} and the proof of step 1 is concluded. Incidentally, this also shows that one must have equality in (1.12).

Step 2. It remains to prove (1.11) for l3l\geq 3. This bound comes again from the Spencer complex:

0𝒯(l)Hom(W,𝒯(l1))Hom(2W,𝒯(l2))Hom(l1W,𝒯(1))δHom(lW,𝒯)𝒟φk+l\begin{split}\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 5.5pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&&&\crcr}}}\ignorespaces{\hbox{\kern-5.5pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 29.5pt\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 29.5pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathcal{T}^{(l)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 72.85892pt\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 72.85892pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\operatorname{Hom}(W,\mathcal{T}^{(l-1)})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 161.44012pt\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 161.44012pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\operatorname{Hom}(\wedge^{2}W,\mathcal{T}^{(l-2)})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 259.48804pt\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 259.48804pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\dots}$}}}}}}}\ignorespaces}}}}\ignorespaces\qquad\\ \qquad\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 6.75pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&&\crcr}}}\ignorespaces{\hbox{\kern-6.75pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\dots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 30.75pt\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 30.75pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\operatorname{Hom}(\wedge^{l-1}W,\mathcal{T}^{(1)})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 112.24237pt\raise 5.43056pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.43056pt\hbox{$\scriptstyle{\delta}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 128.79791pt\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 128.79791pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\operatorname{Hom}(\wedge^{l}W,\mathcal{T})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 215.02359pt\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 215.02359pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathcal{D}^{k+l}_{\varphi}}$}}}}}}}\ignorespaces}}}}\ignorespaces\end{split}

Because 𝒯(1)\mathcal{T}^{(1)} is involutive, this sequence is exact up until imδ\operatorname{im}\delta. It follows that

rank(imδ)=k=1l(1)k+1rank(Hom(lkW,𝒯(k))).\operatorname{rank}(\operatorname{im}\delta)=\sum_{k=1}^{l}(-1)^{k+1}\operatorname{rank}\left(\operatorname{Hom}(\wedge^{l-k}W,\mathcal{T}^{(k)})\right).

We already know that Cartan’s bound is achieved for all prolongations, so rank𝒯(k)\operatorname{rank}\mathcal{T}^{(k)} can be given entirely in terms of the Cartan characters of 𝒯\mathcal{T}. An induction argument using si(1)=si++sns^{(1)}_{i}=s_{i}+\dots+s_{n} gives rank𝒯(k)=i=1nsi(i+k1k)\operatorname{rank}\mathcal{T}^{(k)}=\sum_{i=1}^{n}s_{i}\binom{i+k-1}{k}, and therefore

rank(imδ)=i=1nsi(k=1l(1)k+1(nlk)(i+k1k)).\operatorname{rank}(\operatorname{im}\delta)=\sum_{i=1}^{n}s_{i}\left(\sum_{k=1}^{l}(-1)^{k+1}\binom{n}{l-k}\binom{i+k-1}{k}\right).

The result now follows from the combinatorial identity

k=1l(1)k+1(nlk)(i+k1k)=(nl)(nil).\sum_{k=1}^{l}(-1)^{k+1}\binom{n}{l-k}\binom{i+k-1}{k}=\binom{n}{l}-\binom{n-i}{l}.\qed

1.6.4. Symbol exact sequences of tableaux and involutivity

We continue to assume that (φ,p):WV(\varphi,p)\colon W\to V is a vector bundle map.

Proposition 1.48.

Both 𝒟φk\mathcal{D}^{k}_{\varphi} and Hom(k+1W,pV)𝒟φk\operatorname{Hom}(\wedge^{k+1}W,p^{*}V)\subset\mathcal{D}^{k}_{\varphi} are involutive tableaux for which every flag is regular.

In order to prove this proposition, will use the following lemma which relates the Spencer differential and the short exact sequence (1.3) induced by the symbol map.

Lemma 1.49.

The following diagram of short exact sequences commutes

Hom(W,Hom(k+1W,pV))\textstyle{{\operatorname{Hom}(W,\operatorname{Hom}(\wedge^{k+1}W,p^{*}V))}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(W,𝒟φk)\textstyle{{\operatorname{Hom}(W,\mathcal{D}^{k}_{\varphi})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δ\scriptstyle{\delta}Hom(W,Hom(kW,pTN))\textstyle{{\operatorname{Hom}(W,\operatorname{Hom}(\wedge^{k}W,p^{*}TN))}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(k+2W,pV)\textstyle{{\operatorname{Hom}(\wedge^{k+2}W,p^{*}V)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟φk+1\textstyle{\mathcal{D}^{k+1}_{\varphi}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(k+1W,pTN),\textstyle{{\operatorname{Hom}(\wedge^{k+1}W,p^{*}TN)},}

where the side vertical arrows are skew-symmetrization. In particular, the Spencer differential δ:Hom(E,𝒟φk)𝒟φk+1\delta\colon\operatorname{Hom}(E,\mathcal{D}^{k}_{\varphi})\to\mathcal{D}^{k+1}_{\varphi} is surjective.

Proof.

This is clear from (1.8) and (1.9). ∎

Proof of Proposition 1.48.

Consider first the tableau 𝒯:=Hom(kW,pV)\mathcal{T}:=\operatorname{Hom}(\wedge^{k}W,p^{*}V). In this case the Spencer differential is the skew-symmetrization map

δ:Hom(W,Hom(kW,pV))Hom(k+1W,pV)\delta\colon\operatorname{Hom}(W,\operatorname{Hom}(\wedge^{k}W,p^{*}V))\to\operatorname{Hom}(\wedge^{k+1}W,p^{*}V)

which is surjective. Setting n=rankWn=\operatorname{rank}W and m=rankVm=\operatorname{rank}V, we have

rank𝒯(1)=m(n(nk)(nk+1))=mk(n+1k+1).\operatorname{rank}\mathcal{T}^{(1)}=m\left({n\binom{n}{k}-\binom{n}{k+1}}\right)=m\,k\binom{n+1}{k+1}.

Now let (Wi)(W_{i}) be a (local) flag for WW. From the exact sequence

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒯i\textstyle{\mathcal{T}_{i}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(kW,pV)\textstyle{\operatorname{Hom}(\wedge^{k}W,p^{*}V)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(kWi,pV)\textstyle{\operatorname{Hom}(\wedge^{k}W_{i},p^{*}V)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

it follows that si=m((ik)(i1k))=m(i1k1)s_{i}=m\left(\binom{i}{k}-\binom{i-1}{k}\right)=m\binom{i-1}{k-1}. Hence, we find that

i=1nisi=mki=kn(ik)=mk(n+1k+1)=rank𝒯(1),\sum_{i=1}^{n}is_{i}=mk\sum_{i=k}^{n}\binom{i}{k}=m\,k\binom{n+1}{k+1}=\operatorname{rank}\mathcal{T}^{(1)},

so Cartan’s Test 1.43 holds and 𝒯\mathcal{T} is involutive.

Next, to show that 𝒟φk\mathcal{D}^{k}_{\varphi} is also involutive we apply Lemma 1.49. Setting 𝒟=𝒟φk\mathcal{D}=\mathcal{D}^{k}_{\varphi} and 𝒮=Hom(kW,pTN)\mathcal{S}=\operatorname{Hom}(\wedge^{k}W,p^{*}TN), the lemma gives a short exact sequence

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒯(1)\textstyle{\mathcal{T}^{(1)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟(1)\textstyle{\mathcal{D}^{(1)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒮(1)\textstyle{\mathcal{S}^{(1)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

Moreover, restriction gives a commutative diagram

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(k+1W,pV)\textstyle{{\operatorname{Hom}(\wedge^{k+1}W,p^{*}V)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟φk\textstyle{\mathcal{D}^{k}_{\varphi}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(kW,pTN)\textstyle{{\operatorname{Hom}(\wedge^{k}W,p^{*}TN)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(k+1Wi,pV)\textstyle{{\operatorname{Hom}(\wedge^{k+1}W_{i},p^{*}V)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟φιWik\textstyle{\mathcal{D}^{k}_{\varphi\circ\iota_{W_{i}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(kWi,pTN)\textstyle{{\operatorname{Hom}(\wedge^{k}W_{i},p^{*}TN)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

so we obtain that the restricted spaces also fit into a short exact sequence

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒯i\textstyle{\mathcal{T}_{i}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟i\textstyle{\mathcal{D}_{i}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒮i\textstyle{\mathcal{S}_{i}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

Since Cartan’s test is satisfied for both 𝒮\mathcal{S} and 𝒯\mathcal{T}, we conclude that it is satisfied for 𝒟\mathcal{D}, which then must also be involutive. ∎

2. Relative algebroids

2.1. Relative algebroids

With the formalism of relative derivations at hand, we are now ready to introduce the geometric objects underlying Cartan’s realization problem and Bryant’s equations (0.5).

Definition 2.1.

An almost Lie algebroid relative to a submersion or in short algebroid relative to a submersion p:MNp\colon M\to N consists of a vector bundle ANA\to N together with a 1-derivation

D:Ω(A)Ω+1(pA){\mathrm{D}}\colon\Omega^{\bullet}(A)\to\Omega^{\bullet+1}(p^{*}A)

relative to pp.

We will denote by (A,p,D)(A,p,{\mathrm{D}}) the entire structure of the algebroid relative to pp and often write BB for the pullback bundle pAMp^{*}A\to M. The manifolds MM and NN will be specified when those are not clear from the context. We like to graphically depict an algebroid relative to pp as

B\textstyle{B\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D\scriptstyle{{\mathrm{D}}}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}N\textstyle{N}

where the dotted arrow is not a map but indicates a derivation that is defined on Ω(A)\Omega^{\bullet}(A) and takes values in Ω+1(B)\Omega^{\bullet+1}(B). We use the letter BB to stand for Bryant, whose equations (0.5) inspired this definition.

According to Lemma 1.12, the structure of an algebroid relative to p:MNp\colon M\to N can also be encoded by a relative bracket and anchor

[,]:2Γ(A)Γ(B),ρ:BpTN[\cdot,\cdot]\colon\wedge^{2}\Gamma(A)\to\Gamma(B),\quad\rho\colon B\to p^{*}TN

subject to the Leibniz rule:

[a0,fa1]=pf[a0,a1]+ρ(pa0)(f)pa1.[a_{0},fa_{1}]=p^{*}f[a_{0},a_{1}]+\mathcal{L}_{\rho(p^{*}a_{0})}(f)p^{*}a_{1}.

The corresponding derivation is determined through the Koszul formula (see the proof of Lemma 1.12).

Example 2.2.

In the case that M=NM=N and p=idp=\operatorname{id}, we recover a notion of an almost Lie algebroid: a vector bundle ANA\to N together with an anchor and a bracket subject to the Leibniz rule, or, dually, with a degree 1 derivation DA{\mathrm{D}}_{A} on Ω(A)\Omega^{\bullet}(A). Such an almost Lie algebroid is a Lie algebroid when DA2=0{\mathrm{D}}_{A}^{2}=0. In particular, we can consider any (almost) Lie algebroid as an algebroid relative to the identity. A special case is the tangent bundle (TP,d)P(TP,{\mathrm{d}})\to P with the de Rham differential d{\mathrm{d}}.

At this point, when dealing with an arbitrary relative algebroid, it is not clear to make sense of “D2=0{\mathrm{D}}^{2}=0”. We will soon see how to handle this issue. This issue also influences our terminology: the objects in Definition 2.1 should properly be called almost relative Lie algebroid. This name is too long and so we refer to them simply as relative algebroids, removing “Lie” from its name rather than adding “almost”. Later, after we make sense of D2=0{\mathrm{D}}^{2}=0, we will be able to define relative Lie algebroids.

Example 2.3 (Relative algebroids in coordinates).

Let (U,(xμ,yϱ))(U,(x^{\mu},y^{\varrho})) be a coordinate system for MM and (V,xμ)(V,x^{\mu}) a coordinate system for NN such that

p(xμ,yϱ)=xμ.p(x^{\mu},y^{\varrho})=x^{\mu}.

Also, fix {ei}\{e_{i}\} a frame for A|VA|_{V}, and let {θi}\{\theta^{i}\} be the corresponding dual coframe. Then the derivation D{\mathrm{D}} is determined by

{Dθi=12cjki(xμ,yϱ)θjθk,Dxμ=Fiμ(xμ,yϱ)θi.\begin{cases}{\mathrm{D}}\theta^{i}=-\frac{1}{2}c^{i}_{jk}(x^{\mu},y^{\varrho})\,\theta^{j}\wedge\theta^{k},\\ {\mathrm{D}}x^{\mu}=F^{\mu}_{i}(x^{\mu},y^{\varrho})\,\theta^{i}.\end{cases} (2.1)

for functions cjki,FiμC(M)c^{i}_{jk},F^{\mu}_{i}\in C^{\infty}(M). Dually, the anchor and bracket are given by

{[ei,ej]=cijkek,ρ(ei)=Fiμxμ.\begin{cases}[e_{i},e_{j}]=c_{ij}^{k}e_{k},\\ \rho(e_{i})=F^{\mu}_{i}\partial_{x^{\mu}}.\end{cases}

The equations above look striking similar to Bryant’s equations (0.5) except that one has D{\mathrm{D}} instead of d{\mathrm{d}} and the cjkic^{i}_{jk} also depends on the “free derivatives” yϱy^{\varrho}.

A true globalization of Bryant’s equations requires a globalization of the projection. The language introduced in the previous section for derivations relative to foliations precisely captures this idea. This is not a significant generalization; as we will see in Section 3, whenever the first prolongation of a relative algebroid exists, it is always relative to a submersion.

Nevertheless, there are natural examples of relative algebroids that are relative to a foliations rather than a submersion. These include relative algebroids underlying Pfaffian fibrations [11] and those obtained through restrictions (Proposition 4.12).

Definition 2.4.

An algebroid relative to a foliation (M,)(M,\mathcal{F}) is a flat foliated vector bundle (M,,B,¯)(M,\mathcal{F},B,\overline{\nabla}) together with a 1-derivation D{\mathrm{D}} relative to \mathcal{F}:

D:Ω(B,¯)ΩB+1.{\mathrm{D}}\colon\Omega^{\bullet}_{(B,\overline{\nabla})}\to\Omega^{\bullet+1}_{B}.

We will write (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) to denote a relative algebroid over (M,)(M,\mathcal{F}). If there is ambiguity in what (M,)(M,\mathcal{F}) could be, we write (B,¯,D)(M,)(B,\overline{\nabla},{\mathrm{D}})\to(M,\mathcal{F}).

Example 2.5.

Let (A,p,D)(A,p,{\mathrm{D}}) be an algebroid relative to a submersion p:MNp\colon M\to N. Then it can be viewed as an algebroid relative to the foliation =ker(dp)\mathcal{F}=\ker({\mathrm{d}}p). Namely, the vector bundle B=pAB=p^{*}A carries a canonical flat \mathcal{F}-connection ¯{\overline{\nabla}} determined by requiring

¯pa=0, for all aΓ(A),\overline{\nabla}p^{*}a=0,\quad\text{ for all }a\in\Gamma(A),

and since Ω(B,¯)=pΩA\Omega^{\bullet}_{(B,\overline{\nabla})}=p^{*}\Omega^{\bullet}_{A}, we can view D{\mathrm{D}} as a derivation relative to \mathcal{F}.

Conversely, any algebroid relative to a foliation is locally an algebroid relative to a submersion. It is globally an algebroid relative to a submersion precisely when \mathcal{F} is simple and ¯\overline{\nabla} has no holonomy along \mathcal{F} (see Corollary 1.17).

Suppose we choose foliation coordinates (U,(xμ,yϱ))(U,(x^{\mu},y^{\varrho})) on (M,)(M,\mathcal{F}), so that plaques of \mathcal{F} correspond to {yϱ=cϱ}\{y^{\varrho}=c^{\varrho}\}. Furthermore, let {ei}\{e_{i}\} be a local frame of flat sections of B|UB|_{U}, so the dual coframe {θi}\{\theta^{i}\} consists of local flat BB-forms in Ω(B,¯)1(U)\Omega^{1}_{(B,\overline{\nabla})}(U). Then the derivation D{\mathrm{D}} is still determined by the same equations (2.1). Also, we now see that we can retrieve Bryant’s equations, i.e., have the cijkc_{ij}^{k} not depend on the free variables, exactly when there is a local coframe (θi)(\theta^{i}) such that ¯(Dθi)=0\overline{\nabla}({\mathrm{D}}\theta^{i})=0. We will see later that this property is always satisfied for any prolongation of a relative algebroid (Definition 3.1).

Using the notion of (φ,p)(\varphi,p)-related derivations (see Definition 1.29, morphisms of relative algebroids can be defined as follows:

Definition 2.6.

A morphism (φ,p):(B1,¯1,D1)(B2,¯2,D2)(\varphi,p)\colon(B_{1},\overline{\nabla}_{1},{\mathrm{D}}_{1})\to(B_{2},\overline{\nabla}_{2},{\mathrm{D}}_{2}) of relative algebroids consists of

  1. (i)

    a map of foliations p:(M1,1)(M2,2)p\colon(M_{1},\mathcal{F}_{1})\to(M_{2},\mathcal{F}_{2}),

  2. (ii)

    a map of flat foliated vector bundles φ:(B1,¯1)(B2,¯2)\varphi\colon(B_{1},{\overline{\nabla}}_{1})\to(B_{2},{\overline{\nabla}}_{2}) covering pp,

such that D1{\mathrm{D}}_{1} is (φ,p)(\varphi,p)-related to D2{\mathrm{D}}_{2}:

D1φ=φD2, on Ω(B2,¯2).{\mathrm{D}}_{1}\circ\varphi^{*}=\varphi^{*}\circ{\mathrm{D}}_{2},\quad\text{ on $\Omega^{\bullet}_{(B_{2},{\overline{\nabla}}_{2})}$}.

2.2. Realizations

In this section, we will fix an algebroid (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) relative to a foliation (M,)(M,\mathcal{F}).

The tangent bundle of any manifold PP is a Lie algebroid with derivation d{\mathrm{d}}, the de Rham differential, so we will denote it by (TP,d)(TP,{\mathrm{d}}). A realization is an object that “realizes” an algebroid as the tangent bundle of a manifold.

Definition 2.7.

A realization (P,r,θ)(P,r,\theta) of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) is a morphism of relative algebroids from (TP,d)(TP,{\mathrm{d}}) to (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) that is fiberwise an isomorphism.

Explicitly, a realization consists of a manifold PP together with a bundle map (θ,r):TPB(\theta,r)\colon TP\to B that is fiberwise an isomorphism and satisfies

dθ=θD,on Ω(B,¯).{\mathrm{d}}\circ\theta^{*}=\theta^{*}\circ{\mathrm{D}},\quad\text{on }\Omega^{\bullet}_{(B,\overline{\nabla})}.
Remark 2.8.

In [16, 17], realizations of Lie algebroids are defined through Maurer-Cartan forms. In our situation, the bundle map (θ,r)(\theta,r) can be reinterpreted as a one-form θΩ1(P;rB)\theta\in\Omega^{1}(P;r^{*}B) that is fiberwise an isomorphism. Choosing any extension \nabla of ¯\overline{\nabla} on BB, Lemma 1.26 gives a splitting under which the relative derivation dθθDΓ(r𝒟(B,¯)1){\mathrm{d}}\theta^{*}-\theta^{*}{\mathrm{D}}\in\Gamma(r^{*}\mathcal{D}^{1}_{(B,\overline{\nabla})}) decomposes into two components: (MCθ,Πdrρθ)(\operatorname{MC}^{\nabla}_{\theta},\Pi\circ{\mathrm{d}}r-\rho\circ\theta), where Π:TMν()\Pi\colon TM\to\nu(\mathcal{F}) is the projection and the Maurer-Cartan form is given by

MCθ:=dθ+12[θ,θ].\operatorname{MC}_{\theta}^{\nabla}:={\mathrm{d}}^{\nabla}\theta+\frac{1}{2}[\theta,\theta]_{\nabla}.

So we conclude that in terms of anchors and brackets a bundle map (θ,r):TPB(\theta,r)\colon TP\to B is a realization of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) if and only if it is fiberwise an isomorphism and satisfies

{dθ=12[θ,θ],Πdr=ρθ.\begin{cases}{\mathrm{d}}^{\nabla}\theta=-\frac{1}{2}[\theta,\theta]_{\nabla},\\ \Pi\circ{\mathrm{d}}r=\rho\circ\theta.\end{cases} (2.2)

We conclude that:

  1. (i)

    If θ\theta is anchored (i.e. Πdr=ρθ)\Pi\circ{\mathrm{d}}r=\rho\circ\theta), then the Maurer-Cartan form is independent of \nabla.

  2. (ii)

    If θ\theta is anchored, it is a realization if and only if its Maurer-Cartan form vanishes.

In local coordinates (see Examples 2.3 and 2.5), writing θ=(θi):TPn\theta=(\theta^{i})\colon TP\to\mathbb{R}^{n} and r=(aμ,bϱ):Ps×rr=(a^{\mu},b^{\varrho})\colon P\to\mathbb{R}^{s}\times\mathbb{R}^{r}, equations (2.2) become

{dθi=12cjki(a,b)θjθk,daμ=Fiμ(a,b)θi,\begin{cases}{\mathrm{d}}\theta^{i}=-\frac{1}{2}c^{i}_{jk}(a,b)\,\theta^{j}\wedge\theta^{k},\\ {\mathrm{d}}a^{\mu}=F^{\mu}_{i}(a,b)\,\theta^{i},\end{cases}

and these are exactly Bryant’s equations (0.5) (see Example 2.5 for why the cjkic^{i}_{jk} may depend on the free derivatives bϱb^{\varrho}).

2.3. Tableaux

Let (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) be an algebroid relative to a foliation (M,)(M,\mathcal{F}). Recall from Lemma 1.20 that there is a canonical flat \mathcal{F}-connection on the vector bundle 𝒟(B,¯)1\mathcal{D}^{1}_{(B,\overline{\nabla})} also denoted by ¯\overline{\nabla}.

Definition 2.9.

The tableau map of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) is the bundle map

τ:𝒟(B,¯)1,X¯XD\tau\colon\mathcal{F}\to\mathcal{D}^{1}_{(B,\overline{\nabla})},\quad X\mapsto\overline{\nabla}_{X}{\mathrm{D}}

The relative algebroid is called standard or non-degenerate when its tableau map τ\tau is fiberwise injective.

The tableau map measures the dependence of the relative algebroid structure on the directions of \mathcal{F}.

Composing a tableau map with the symbol map

\textstyle{\mathcal{F}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ\scriptstyle{\tau}σ(τ)\scriptstyle{\sigma(\tau)}𝒟(B,¯)1\textstyle{\mathcal{D}^{1}_{(B,\overline{\nabla})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}Hom(B,ν())\textstyle{\operatorname{Hom}(B,\nu(\mathcal{F}))}

we see that our tableau map covers a classical tableau map

σ(τ):Hom(B,ν()),X¯Xρ\sigma(\tau)\colon\mathcal{F}\to\operatorname{Hom}(B,\nu(\mathcal{F})),\quad X\mapsto\overline{\nabla}_{X}\rho

where ρ=σ(D)\rho=\sigma({\mathrm{D}}) is the anchor of the relative algebroid. We call σ(τ)\sigma(\tau) the symbol tableau of the relative algebroid. Explicitly, it is given by

σ(τ)(X)(b)=¯XBottρ(b)ρ(¯Xb), for X and bΓ(B).\sigma(\tau)(X)(b)=\overline{\nabla}^{\mathrm{Bott}}_{X}\rho(b)-\rho\left(\overline{\nabla}_{X}b\right),\quad\text{ for $X\in\mathcal{F}$ and $b\in\Gamma(B)$}. (2.3)

On the other hand, using Lemma 1.12, one finds that the bracket [,]τ(X)[\cdot,\cdot]_{\tau(X)} associated to derivation τ(X)=¯XD\tau(X)={\overline{\nabla}}_{X}{\mathrm{D}} is given by

[b1,b2]τ(X)=¯X([b1,b2]), for b1,b2Γ(B,¯),[b_{1},b_{2}]_{\tau(X)}={\overline{\nabla}}_{X}([b_{1},b_{2}]),\quad\text{ for }b_{1},b_{2}\in\Gamma_{(B,{\overline{\nabla}})}, (2.4)

where [,][\cdot,\cdot] is the bracket associated to D{\mathrm{D}}.

2.4. Torsion

We will now discuss the first order obstructions to the existence of realizations of a relative algebroid. In the following discussion we fix a relative algebroid (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) over (M,)(M,\mathcal{F}), and we denote by [,][\cdot,\cdot] and ρ\rho the corresponding bracket and anchor. Recall from Lemma 1.27 that there is an exact sequence

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(B,)\textstyle{\operatorname{Hom}(B,\mathcal{F})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟B1\textstyle{\mathcal{D}^{1}_{B}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Π\scriptstyle{\Pi}𝒟(B,¯)1\textstyle{\mathcal{D}^{1}_{(B,\overline{\nabla})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

A pointwise lift or completion of D{\mathrm{D}} at mMm\in M is an element D~m(𝒟B1)m\tilde{{\mathrm{D}}}_{m}\in(\mathcal{D}^{1}_{B})_{m} such that Π(D~m)=Dm\Pi(\tilde{{\mathrm{D}}}_{m})={\mathrm{D}}_{m}. We let

L:={D~m𝒟B1|Π(D~m)=Dm for some mM }=Π1(D)L:=\{\tilde{{\mathrm{D}}}_{m}\in\mathcal{D}^{1}_{B}\ |\ \Pi(\tilde{{\mathrm{D}}}_{m})={\mathrm{D}}_{m}\text{ for some $m\in M$ }\}=\Pi^{-1}({\mathrm{D}})

be the space of completions of D{\mathrm{D}}, with projection p1:LMp_{1}\colon L\to M, D~mm\tilde{{\mathrm{D}}}_{m}\mapsto m. It is an affine bundle modeled on Hom(B,)\operatorname{Hom}(B,\mathcal{F}). The foliation \mathcal{F} pulls back to a foliation p1p_{1}^{*}\mathcal{F} on LL, and the bundle p1Bp_{1}^{*}B inherits a flat p1p_{1}^{*}\mathcal{F}-connection p1¯p_{1}^{*}\overline{\nabla} from (B,¯)(B,\overline{\nabla}).

Remark 2.10.

It follows from Lemma 1.27 that a pointwise lift D~m\tilde{{\mathrm{D}}}_{m} of D{\mathrm{D}} is completely determined by a pointwise lift ρ~m:BmTmM\tilde{\rho}_{m}\colon B_{m}\to T_{m}M of the symbol ρ\rho. In terms of the bracket [,]:Γ(B,¯)×Γ(B,¯)Γ(B)[\cdot,\cdot]:\Gamma_{(B,{\overline{\nabla}})}\times\Gamma_{(B,{\overline{\nabla}})}\to\Gamma(B), this means that such a lift determines a unique extension to a bracket [,]ρ~m:Γ(B)×Γ(B)Bm[\cdot,\cdot]_{\tilde{\rho}_{m}}:\Gamma(B)\times\Gamma(B)\to B_{m}, defined on all sections by requiring

[b1,fb2]ρ~m=f(m)[b1,b2](m)+ρ~(b1),dmfb2(m),[b_{1},fb_{2}]_{\tilde{\rho}_{m}}=f(m)[b_{1},b_{2}](m)+\langle\tilde{\rho}(b_{1}),{\mathrm{d}}_{m}f\rangle b_{2}(m),

for any b1,b2Γ(B,¯)b_{1},b_{2}\in\Gamma_{(B,{\overline{\nabla}})} and fC(M)f\in C^{\infty}(M).

For the next definition we recall that, according to Example 1.8, an element D~mL\tilde{{\mathrm{D}}}_{m}\in L can be regarded as a derivation D~:Ω(B)+1Bm\tilde{{\mathrm{D}}}\colon\Omega^{\bullet}(B)\to\wedge^{\bullet+1}B^{*}_{m}.

Definition 2.11.

The torsion of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) is the 2-derivation TΓ(𝒟(p1B,p1¯)2)T\in\Gamma(\mathcal{D}^{2}_{(p_{1}^{*}B,p_{1}^{*}\overline{\nabla})}) relative to p1p_{1}^{*}\mathcal{F} defined by

T(p1α)|D~m:=D~m(Dα),for p1αΩ(p1B,¯)p1Ω(B,¯).T(p^{*}_{1}\alpha)\big{|}_{\tilde{{\mathrm{D}}}_{m}}:=\tilde{{\mathrm{D}}}_{m}\left({\mathrm{D}}\alpha\right),\quad\text{for $p_{1}^{*}\alpha\in\Omega^{\bullet}_{(p_{1}^{*}B,\overline{\nabla})}\cong p_{1}^{*}\Omega^{\bullet}_{(B,\overline{\nabla})}$.}

The symbol σ(T)\sigma(T) is called the symbol torsion of D{\mathrm{D}}.

It follows from Lemma 1.31 that the torsion TT has associated 2-bracket given by

[p1b1,p1b2,p1b3]|D~m=[[b1,b2],b3]ρ~m+[[b2,b3],b1]ρ~m+[[b3,b1],b2]ρ~m,[p^{*}_{1}b_{1},p^{*}_{1}b_{2},p^{*}_{1}b_{3}]|_{\tilde{{\mathrm{D}}}_{m}}=[[b_{1},b_{2}],b_{3}]_{\tilde{\rho}_{m}}+[[b_{2},b_{3}],b_{1}]_{\tilde{\rho}_{m}}+[[b_{3},b_{1}],b_{2}]_{\tilde{\rho}_{m}}, (2.5)

for b1,b2,b3Γ(B,¯)b_{1},b_{2},b_{3}\in\Gamma_{(B,{\overline{\nabla}})}, where ρ~m\tilde{\rho}_{m} and [,]ρ~m[\cdot,\cdot]_{\tilde{\rho}_{m}} denote the symbol and the bracket of D~m\tilde{{\mathrm{D}}}_{m}, as in Remark 2.10. The symbol torsion, in turn, is given by

σ(T)(pb1,pb2)|D~m(f)=ρ~m(b1)(ρ(b2)(f))ρ~m(b2)(ρ(b1)(f))ρm([b1,b2])(f),\sigma(T)(p^{*}b_{1},p^{*}b_{2})|_{\tilde{{\mathrm{D}}}_{m}}(f)=\tilde{\rho}_{m}(b_{1})\left(\rho(b_{2})(f)\right)-\tilde{\rho}_{m}(b_{2})\left(\rho(b_{1})(f)\right)-\rho_{m}([b_{1},b_{2}])(f), (2.6)

for each b1,b2Γ(B,¯)b_{1},b_{2}\in\Gamma_{(B,\overline{\nabla})} and all fCbasf\in C^{\infty}_{\text{{bas}}}.

Let (P,r,θ)(P,r,\theta) be a realization of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}). For each pPp\in P, it determines the lift of ρ\rho at r(p)r(p) given by

ρ~r(p):=dprθp:Br(p)Tr(p)M.\tilde{\rho}_{r(p)}:={\mathrm{d}}_{p}r\circ\theta_{p}\colon B_{r(p)}\to T_{r(p)}M. (2.7)

Therefore, it determines also a lift D~r(p)\tilde{{\mathrm{D}}}_{r(p)} of D{\mathrm{D}}.

Proposition 2.12 (Existence of Realizations: first order necessary condition).

Let (P,r,θ)(P,r,\theta) be a realization of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}). Then, for each pPp\in P we have

T|D~r(p)=0.T\big{|}_{\tilde{{\mathrm{D}}}_{r(p)}}=0.
Proof.

By definition, dθ=θD{\mathrm{d}}\circ\theta^{*}=\theta^{*}\circ{\mathrm{D}} on Ω(B,¯)\Omega^{\bullet}_{(B,\overline{\nabla})}, so the derivation

(θp1)(dθ)|p:Ω(B)+1TpP+1Br(p)\left(\theta^{-1}_{p}\right)^{*}\circ\left({\mathrm{d}}\circ\theta^{*}\right)|_{p}\colon\Omega^{\bullet}(B)\to\wedge^{\bullet+1}T^{*}_{p}P\to\wedge^{\bullet+1}B^{*}_{r(p)}

is a lift of D{\mathrm{D}} at r(p)r(p). Since its symbol is (2.7), it must coincide with D~r(p)\tilde{{\mathrm{D}}}_{r(p)}. The proposition now follows from d2=0{\mathrm{d}}^{2}=0. ∎

Recall that any two pointwise lifts of Dm{\mathrm{D}}_{m} differ by an element ξHom(Bm,m)\xi\in\operatorname{Hom}(B_{m},\mathcal{F}_{m}), where ξ\xi acts as a derivation as described in the Lemma 1.27.

Proposition 2.13.

Let D~m\tilde{{\mathrm{D}}}_{m} be a pointwise lift of D{\mathrm{D}} at mm and ξHom(Bm,m)\xi\in\operatorname{Hom}(B_{m},\mathcal{F}_{m}). Then

T|D~m+ξT|D~m=δτξ, in Ω(p1B,¯)p1Ω(B,¯).T\big{|}_{\tilde{{\mathrm{D}}}_{m}+\xi}-T\big{|}_{\tilde{{\mathrm{D}}}_{m}}=\delta_{\tau}\xi,\quad\text{ in }\Omega^{\bullet}_{(p_{1}^{*}B,\overline{\nabla})}\cong p_{1}^{*}\Omega^{\bullet}_{(B,\overline{\nabla})}.
Proof.

Note that ¯X(Dα)=(¯XD)α=τ(X)α\overline{\nabla}_{X}(D\alpha)=(\overline{\nabla}_{X}D)\alpha=\tau(X)\alpha for αΩ(B,¯)\alpha\in\Omega^{\bullet}_{(B,\overline{\nabla})}. By Lemma 1.27, we find that for βΩ(B,¯)1\beta\in\Omega^{1}_{(B,\overline{\nabla})} we have

(T(β)|D~+ξT(β)|D~)(b1,b2,b3)\displaystyle\left(T(\beta)\big{|}_{\tilde{{\mathrm{D}}}+\xi}-T(\beta)\big{|}_{\tilde{{\mathrm{D}}}}\right)(b_{1},b_{2},b_{3}) =ι(ξ)(Dβ)(b1,b2,b3)\displaystyle=\iota(\xi)({\mathrm{D}}\beta)(b_{1},b_{2},b_{3})
=¯ξ(b1)(Dβ)(b2,b3)+c.p.\displaystyle=\overline{\nabla}_{\xi(b_{1})}({\mathrm{D}}\beta)(b_{2},b_{3})+\mbox{c.p.}
=τ(ξ(b1))β(b2,b3)+c.p.\displaystyle=\tau(\xi(b_{1}))\beta(b_{2},b_{3})+\mbox{c.p.}
=(δτξ)(β)(b1,b2,b3).\displaystyle=\left(\delta_{\tau}\xi\right)(\beta)(b_{1},b_{2},b_{3}).

Similarly, for fCbasf\in C^{\infty}_{\text{{bas}}}, we find

(T(f)|D~+ξT(f)|D~)(b1,b2)\displaystyle\left(T(f)\big{|}_{\tilde{{\mathrm{D}}}+\xi}-T(f)\big{|}_{\tilde{{\mathrm{D}}}}\right)(b_{1},b_{2}) =¯ξ(b1)(Df)(b2)¯ξ(b2)(Df)(b1)\displaystyle=\overline{\nabla}_{\xi(b_{1})}(Df)(b_{2})-\overline{\nabla}_{\xi(b_{2})}({\mathrm{D}}f)(b_{1})
=(δτξ)(f)(b1,b2),\displaystyle=\left(\delta_{\tau}\xi\right)(f)(b_{1},b_{2}),

which completes the proof. ∎

This justifies the following definition.

Definition 2.14.

The intrinsic torsion of a relative algebroid (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) is the section ΘΓ(H1,2(τ))\Theta\in\Gamma(H^{-1,2}(\tau)) defined by

Θm:=[T|D~m],\Theta_{m}:=[T\big{|}_{\tilde{{\mathrm{D}}}_{m}}],

for any D~mLm\tilde{{\mathrm{D}}}_{m}\in L_{m}.

Corollary 2.15.

If a relative algebroid (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) admits a realization through mm then Θm=0\Theta_{m}=0.

2.5. Curvature

Suppose we are given a section D~Γ(L)\tilde{{\mathrm{D}}}\in\Gamma(L). In this case, D~\tilde{{\mathrm{D}}} is a completion of the derivation D{\mathrm{D}} which is an actual 1-derivation on BB, so D~2\tilde{{\mathrm{D}}}^{2} defines a 2-derivation on BB. Recalling again from Lemma 1.27 that there is a short exact sequence

0Hom(2B,)𝒟B2Π𝒟(B,¯)20.\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 5.5pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&&&\crcr}}}\ignorespaces{\hbox{\kern-5.5pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 29.5pt\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 29.5pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\operatorname{Hom}(\wedge^{2}B,\mathcal{F})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 114.97028pt\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 114.97028pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathcal{D}^{2}_{B}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 137.78418pt\raise 5.39166pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.39166pt\hbox{$\scriptstyle{\Pi}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 155.40918pt\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 155.40918pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathcal{D}^{2}_{(B,\overline{\nabla})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 195.84808pt\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 195.84808pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{0}$}}}}}}}\ignorespaces}}}}\ignorespaces.

We conclude that:

Proposition 2.16.

Given a section D~Γ(L)\tilde{{\mathrm{D}}}\in\Gamma(L), the torsion of D{\mathrm{D}} along D~\tilde{{\mathrm{D}}} is

T|D~=Π(D~2).T|_{\tilde{{\mathrm{D}}}}=\Pi(\tilde{{\mathrm{D}}}^{2}).

If TT vanishes identically on the image of D~\tilde{{\mathrm{D}}}, we call D~\tilde{{\mathrm{D}}} a torsionless lift of D{\mathrm{D}}. In that case, according to the previous short exact sequence, the derivation D~2\tilde{{\mathrm{D}}}^{2} is a section of Hom(2B,)\operatorname{Hom}(\wedge^{2}B,\mathcal{F}).

Definition 2.17.

Let D~Γ(M(1))\tilde{{\mathrm{D}}}\in\Gamma(M^{(1)}) be a torsionless lift of D{\mathrm{D}}. The curvature of D~\tilde{{\mathrm{D}}} is the section

κD~=D~2Γ(Hom(2B,))Γ(𝒟B2).\kappa_{\tilde{{\mathrm{D}}}}=\tilde{{\mathrm{D}}}^{2}\in\Gamma(\operatorname{Hom}(\wedge^{2}B,\mathcal{F}))\subset\Gamma(\mathcal{D}^{2}_{B}).
Remark 2.18.

The curvature, as a 2-derivation, is uniquely determined by its symbol. By the discussion proceeding Lemma 1.31, if ρ~\tilde{\rho} and [,]D~[\cdot,\cdot]_{\tilde{{\mathrm{D}}}} are the anchor and bracket associated to D~\tilde{{\mathrm{D}}}, then the curvature is

κD~(b1,b2)=[ρ~(b1),ρ~(b2)]ρ~([b1,b2]D~),\kappa_{{\tilde{{\mathrm{D}}}}}(b_{1},b_{2})=[\tilde{\rho}(b_{1}),\tilde{\rho}(b_{2})]-\tilde{\rho}([b_{1},b_{2}]_{\tilde{{\mathrm{D}}}}),

for b1,b2Γ(B)b_{1},b_{2}\in\Gamma(B). Note that the curvature depends pointwise on the first jet of the section D~\tilde{{\mathrm{D}}} and that for any local sections b1,b2ΓBb_{1},b_{2}\in\Gamma_{B} and fCbasf\in C^{\infty}_{\text{{bas}}} one has

κD~(b1,b2)(f)=0.\kappa_{{\tilde{{\mathrm{D}}}}}(b_{1},b_{2})(f)=0.

Moreover, the fact that D~\tilde{{\mathrm{D}}} is a torsionless lift of D{\mathrm{D}} implies that for any local flat sections b1,b2,b3Γ(B,¯)b_{1},b_{2},b_{3}\in\Gamma_{(B,{\overline{\nabla}})} one also has

[[b1,b2]D~,b3]D~+[[b2,b3]D~,b1]D~+[[b3,b1]D~,b2]D~=0.[[b_{1},b_{2}]_{\tilde{{\mathrm{D}}}},b_{3}]_{\tilde{{\mathrm{D}}}}+[[b_{2},b_{3}]_{\tilde{{\mathrm{D}}}},b_{1}]_{\tilde{{\mathrm{D}}}}+[[b_{3},b_{1}]_{\tilde{{\mathrm{D}}}},b_{2}]_{\tilde{{\mathrm{D}}}}=0. (2.8)
Lemma 2.19.

The curvature κD~\kappa_{\tilde{{\mathrm{D}}}} is closed in the Spencer complex of τ\tau:

δτκD~=0.\delta_{\tau}\kappa_{\tilde{{\mathrm{D}}}}=0.
Proof.

Since the curvature is uniquely determined by it’s symbol, it is enough to show that the Spencer differential of the symbol tableau vanishes (Section 2.3). Since the result is a C(M)C^{\infty}(M)-linear, it is enough to show that it vanishes on parallel sections. For b1,b2,b3Γ(B,¯)b_{1},b_{2},b_{3}\in\Gamma_{(B,\overline{\nabla})} we find, combining (1.9) and (2.3),

δσ(τ)σ(κD~)(b1,b2,b3)\displaystyle\delta_{\sigma(\tau)}\sigma\left(\kappa_{\tilde{{\mathrm{D}}}}\right)(b_{1},b_{2},b_{3}) =¯κD~(b1,b2)Bottρ(b3)+c.p.\displaystyle=\overline{\nabla}^{\mathrm{Bott}}_{\kappa_{\tilde{{\mathrm{D}}}}(b_{1},b_{2})}\rho(b_{3})+\mathrm{c.p.}
=Π([κD~(b1,b2),ρ~(b3)])+c.p.\displaystyle=\Pi\left(\left[\kappa_{\tilde{{\mathrm{D}}}}(b_{1},b_{2}),\tilde{\rho}(b_{3})\right]\right)+\mathrm{c.p.}
=Π([ρ~([b1,b2]D~),ρ~(b3)])+c.p.\displaystyle=-\Pi\left(\left[\tilde{\rho}\left([b_{1},b_{2}]_{\tilde{{\mathrm{D}}}}\right),\tilde{\rho}(b_{3})\right]\right)+\mathrm{c.p.}
=Π(κD~([b1,b2]D~,b3))Πρ~([[b1,b2]D~,b3]D~)+c.p.\displaystyle=-\Pi\left(\kappa_{\tilde{{\mathrm{D}}}}\left([b_{1},b_{2}]_{\tilde{{\mathrm{D}}}},b_{3}\right)\right)-\Pi\circ\tilde{\rho}\left([[b_{1},b_{2}]_{\tilde{{\mathrm{D}}}},b_{3}]_{\tilde{{\mathrm{D}}}}\right)+\mathrm{c.p.}
=ρ([[b1,b2]D~,b3]D~+[[b2,b3]D~,b1]D~+[[b3,b1]D~,b2]D~)=0\displaystyle=-\rho([[b_{1},b_{2}]_{\tilde{{\mathrm{D}}}},b_{3}]_{\tilde{{\mathrm{D}}}}+[[b_{2},b_{3}]_{\tilde{{\mathrm{D}}}},b_{1}]_{\tilde{{\mathrm{D}}}}+[[b_{3},b_{1}]_{\tilde{{\mathrm{D}}}},b_{2}]_{\tilde{{\mathrm{D}}}})=0

where Π:TMν()\Pi\colon TM\to\nu(\mathcal{F}) is the projection and the last identity follows from (2.8). ∎

Definition 2.20.

The intrinsic curvature of the relative algebroid (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) is the section ΩΓ(p1H0,2(τ))\Omega\in\Gamma\left(p_{1}^{*}H^{0,2}(\tau)\right) defined along a torsionless lift D~Γ(M(1))\tilde{{\mathrm{D}}}\Gamma(M^{(1)}) as

Ω|D~:=[κD~].\Omega\big{|}_{\tilde{{\mathrm{D}}}}:=[\kappa_{\tilde{{\mathrm{D}}}}].
Remark 2.21.

The intrinsic curvature Ω\Omega depends affinely on the values of D~\tilde{{\mathrm{D}}} in M(1)M^{(1)}, and not on its first derivatives, in contrast to κD~\kappa_{\tilde{{\mathrm{D}}}}, which depends on the first jet of D~\tilde{{\mathrm{D}}}.

3. Prolongation and integrability

We shall now attempt to complete a relative algebroid to a true Lie algebroid.

3.1. Prolongations

Let (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) be an algebroid relative to a foliation \mathcal{F} on MM. There is no direct way to make sense of “D2=0{\mathrm{D}}^{2}=0”, because D{\mathrm{D}} is only defined on flat forms in Ω(B,¯)\Omega^{\bullet}_{(B,\overline{\nabla})}. To remedy this issue we consider extensions of D{\mathrm{D}} to ΩB\Omega^{\bullet}_{B} as in Section 1.5.

Definition 3.1.

A prolongation of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) is an algebroid (B,p1,D1)(B,p_{1},{\mathrm{D}}_{1}) relative to a submersion p:M1Mp\colon M_{1}\to M satisfying

  1. (i)

    (Extension) D1α=(p1)Dα{\mathrm{D}}_{1}\alpha=(p_{1})^{*}{\mathrm{D}}\alpha for αΩ(B,¯)\alpha\in\Omega^{\bullet}_{(B,\overline{\nabla})}, and

  2. (ii)

    (Completion) D1D=0{\mathrm{D}}_{1}\circ{\mathrm{D}}=0.

In this context, we write B1:=p1BB_{1}:=p_{1}^{*}B for the pullback bundle.

If (A,p,D)(A,p,D) is an algebroid relative to a submersion p:MNp\colon M\to N, we like to graphically depict a prolongation as

B1\textstyle{B_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B\textstyle{B\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D1\scriptstyle{{\mathrm{D}}_{1}}A\textstyle{A\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D\scriptstyle{{\mathrm{D}}}M1\textstyle{M_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p1\scriptstyle{p_{1}}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}N.\textstyle{N.}
Remark 3.2.

Note that for a relative algebroid the anchor and the bracket do not satisfy integrability conditions. For a prolongation, condition (ii) imposes a set of integrability conditions. By Lemma 1.31, in terms of the brackets of D{\mathrm{D}} and D1{\mathrm{D}}_{1}, this condition amounts to the Jacobi type identity

[b1,[b2,b3]]1+[b2,[b3,b1]]1+[b3,[b1,b2]]1=0,[b_{1},[b_{2},b_{3}]]_{1}+[b_{2},[b_{3},b_{1}]]_{1}+[b_{3},[b_{1},b_{2}]]_{1}=0,

together with the fact that the anchor almost preserve brackets:

ρ1(p[b1,b2])(f)=ρ1(pb1)(ρ0(b2)(f))ρ1(pb2)(ρ0(b1)(f)),\rho_{1}(p^{*}[b_{1},b_{2}])(f)=\rho_{1}(p^{*}b_{1})\big{(}\rho_{0}(b_{2})(f)\big{)}-\rho_{1}(p^{*}b_{2})\big{(}\rho_{0}(b_{1})(f)\big{)},

for any sections b1,b2,b2Γ(B,¯)b_{1},b_{2},b_{2}\in\Gamma_{(B,{\overline{\nabla}})} and fCbasf\in C^{\infty}_{\text{{bas}}}.

Prolongations may or may not exist, and its existence is contingent on vanishing of the intrinsic torsion. Thus, the space of torsionless lifts

p1:M(1)M,where M(1)={D~mL:T|D~m=0},p_{1}\colon M^{(1)}\to M,\quad\mbox{where }M^{(1)}=\{\tilde{{\mathrm{D}}}_{m}\in L\colon T\big{|}_{\tilde{{\mathrm{D}}}_{m}}=0\},

plays a special role. Actually, if this space is smooth then it yields a canonical prolongation!

Proposition 3.3.

Let (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) be an algebroid relative to a foliation (M,)(M,\mathcal{F}) and assume M(1)M^{(1)} has a smooth structure such that p1:M(1)Mp_{1}\colon M^{(1)}\to M is a submersion. Then the relative algebroid (B,p(1),D(1))(B,p^{(1)},{\mathrm{D}}^{(1)}), where D(1){\mathrm{D}}^{(1)} is defined by

(D(1)α)|D~m:=D~mα,for αΩB and D~mM(1).({\mathrm{D}}^{(1)}\alpha)\big{|}_{\tilde{{\mathrm{D}}}_{m}}:=\tilde{{\mathrm{D}}}_{m}\alpha,\quad\mbox{for $\alpha\in\Omega^{\bullet}_{B}$ and $\tilde{{\mathrm{D}}}_{m}\in M^{(1)}$.}

is a prolongation of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}).

Remark 3.4.

We will see later, in in Example 4.6, that the prolongation M(1)M^{(1)} is universal: any other prolongation must factor through M(1)M^{(1)}.

Proof.

Since p1p_{1} is a submersion, one checks immediately that for any αΩB\alpha\in\Omega^{\bullet}_{B}, D(1)αD^{(1)}\alpha is a smooth form. The derivation property then follows from the fact that each D~mM(1)\tilde{{\mathrm{D}}}_{m}\in M^{(1)} is a derivation relative to the inclusion. From the definition of D(1){\mathrm{D}}^{(1)} it follows that

D(1)α=(p1)Dα,for αΩ(B,¯).{\mathrm{D}}^{(1)}\alpha=(p_{1})^{*}{\mathrm{D}}\alpha,\quad\text{for $\alpha\in\Omega^{\bullet}_{(B,\overline{\nabla})}$}.

Finally, since every D~M(1)\tilde{{\mathrm{D}}}\in M^{(1)} is a torsionless lift of D{\mathrm{D}}, it follows that D(1)D|D~=TD~=0{\mathrm{D}}^{(1)}\circ{\mathrm{D}}\big{|}_{\tilde{{\mathrm{D}}}}=T_{\tilde{{\mathrm{D}}}}=0. Hence, (B,p(1),D(1))(B,p^{(1)},{\mathrm{D}}^{(1)}) is a prolongation of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}). ∎

Example 3.5 (1st prolongation in coordinates).

Let us assume that we have fixed local coordinates (U,(xμ,yϱ))(U,(x^{\mu},y^{\varrho})) and a frame {ei}\{e_{i}\} with dual coframe {θi}\{\theta^{i}\}, as in Examples 2.3 and 2.5. The the derivation D{\mathrm{D}} is determined by

{Dθi=12cjki(x,y)θjθk,Dxμ=Fiμ(x,y)θi,\begin{cases}{\mathrm{D}}\theta^{i}=-\frac{1}{2}c^{i}_{jk}(x,y)\,\theta^{j}\wedge\theta^{k},\\ {\mathrm{D}}x^{\mu}=F^{\mu}_{i}(x,y)\,\theta^{i},\end{cases} (3.1)

for some functions cjki,FiμC(M)c^{i}_{jk},F^{\mu}_{i}\in C^{\infty}(M). The completions of the anchor take the form

ρ~(ei)=Fiμ(x,y)xμ+uiϱyϱ,\tilde{\rho}(e_{i})=F^{\mu}_{i}(x,y)\partial_{x^{\mu}}+u_{i}^{\varrho}\partial_{y^{\varrho}},

where uiϱu_{i}^{\varrho} should be thought of as coordinates on the fibers of LML\to M. The corresponding completion of the derivation D~\tilde{{\mathrm{D}}} is determined by D~yϱ=uiϱθi\tilde{{\mathrm{D}}}y^{\varrho}=u^{\varrho}_{i}\theta^{i}.

The first prolongation space M(1)M^{(1)} consists of points (xμ,yϱ,ujς)(x^{\mu},y^{\varrho},u_{j}^{\varsigma}) for which D~D=0\tilde{{\mathrm{D}}}\circ{\mathrm{D}}=0. This yields a system of equations

D~Dxμ=0,D~Dθi=0.\tilde{{\mathrm{D}}}{\mathrm{D}}x^{\mu}=0,\qquad\tilde{{\mathrm{D}}}{\mathrm{D}}\theta^{i}=0.

The first set of equations corresponds to the vanishing of the symbol torsion associated with the extension – see (2.6):

ρ~(ei)(ρ(ej)(xμ)ρ~(ej)(ρ(ei)(xμ)ρ([ei,ej])(xμ)=0,\tilde{\rho}(e_{i})(\rho(e_{j})(x^{\mu})-\tilde{\rho}(e_{j})(\rho(e_{i})(x^{\mu})-\rho([e_{i},e_{j}])(x^{\mu})=0,

while the second set of equations amounts to the vanishing of the torsion on the frame {ei}\{e_{i}\} – see (2.5):

[[e1,e2],e3]ρ~m+[[e2,e3],e1]ρ~m+[[e3,e1],e2]ρ~m=0.[[e_{1},e_{2}],e_{3}]_{\tilde{\rho}_{m}}+[[e_{2},e_{3}],e_{1}]_{\tilde{\rho}_{m}}+[[e_{3},e_{1}],e_{2}]_{\tilde{\rho}_{m}}=0.

Together they express the vanishing of the total torsion, and they give rise to the system of equations

FiμyϱujϱFjμyϱuiϱ\displaystyle\frac{\partial F^{\mu}_{i}}{\partial y^{\varrho}}u_{j}^{\varrho}-\frac{\partial F^{\mu}_{j}}{\partial y^{\varrho}}u_{i}^{\varrho} =FiνFjμxνFjνFiμxνcijkFkμ,\displaystyle=F_{i}^{\nu}\frac{\partial F^{\mu}_{j}}{\partial x^{\nu}}-F_{j}^{\nu}\frac{\partial F^{\mu}_{i}}{\partial x^{\nu}}-c^{k}_{ij}F_{k}^{\mu},
cjkiyϱulϱ+ckliyϱujϱ+cljiyϱukϱ\displaystyle\frac{\partial c^{i}_{jk}}{\partial y^{\varrho}}u^{\varrho}_{l}+\frac{\partial c^{i}_{kl}}{\partial y^{\varrho}}u^{\varrho}_{j}+\frac{\partial c^{i}_{lj}}{\partial y^{\varrho}}u^{\varrho}_{k} =cmjicklm+cmkicljm+cmlicjkm\displaystyle=c^{i}_{mj}c^{m}_{kl}+c^{i}_{mk}c^{m}_{lj}+c^{i}_{ml}c^{m}_{jk}
FjμcklixμFkμcljixμFlμcjkixμ.\displaystyle\qquad-F^{\mu}_{j}\frac{\partial c^{i}_{kl}}{\partial x^{\mu}}-F^{\mu}_{k}\frac{\partial c^{i}_{lj}}{\partial x^{\mu}}-F^{\mu}_{l}\frac{\partial c^{i}_{jk}}{\partial x^{\mu}}.

Assuming that M(1)M^{(1)} is smooth, these equations will determine a subset of the variables uiαu_{i}^{\alpha} as functions of xμx^{\mu}, yϱy^{\varrho} and the remaining variables uiϱu_{i}^{\varrho}, call them vφv^{\varphi}. Hence, we have a set of local coordinates (xμ,yϱ,vφ)(x^{\mu},y^{\varrho},v^{\varphi}) for M(1)M^{(1)} and the derivation D(1){\mathrm{D}}^{(1)} is then defined by (3.1) together with

D(1)yϱ=uiϱ(xμ,yϱ,vφ)θi.{\mathrm{D}}^{(1)}y^{\varrho}=u^{\varrho}_{i}(x^{\mu},y^{\varrho},v^{\varphi})\,\theta^{i}.

In what follows, given a relative algebroid it will be convenient to identify the tableau of its 1st prolongation with the 1st prolongation of its tableau. The precise identification is as follows.

Lemma 3.6.

Let (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) be a relative algebroid. The tableau of its first prolongation is canonically isomorphic to p1τ(1)p1Hom(B,)p_{1}^{*}\tau^{(1)}\subset p_{1}^{*}\operatorname{Hom}(B,\mathcal{F}) (as a tableau), where τ(1)\tau^{(1)} is the first prolongation of the tableau τ\tau of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}).

Proof.

Notice that we have the following:

  1. (a)

    The 1st prolongation (B,p1,D(1))(B,p_{1},D^{(1)}) has a fiberwise injective tableau map

    kerdp1p1𝒟B1,X¯XD(1);\ker{\mathrm{d}}p_{1}\to p^{*}_{1}\mathcal{D}^{1}_{B},\quad X\mapsto{\overline{\nabla}}_{X}D^{(1)};
  2. (b)

    The 1st prolongation of τ\tau is the classical tableau

    τ(1):=ker(δ:Hom(B,)𝒟(B,¯)2);\tau^{(1)}:=\ker(\delta\colon\operatorname{Hom}(B,\mathcal{F})\to\mathcal{D}^{2}_{(B,{\overline{\nabla}})});

Now, the bundle p1:LMp_{1}\colon L\to M is an affine bundle modeled on Hom(B,)\operatorname{Hom}(B,\mathcal{F}), and by Proposition 2.13, the restriction p1:M(1)Mp_{1}\colon M^{(1)}\to M is an affine bundle modeled on τ(1)\tau^{(1)}. It follows that we have an isomorphism of tableaux

kerdp1p1τ(1).\ker{\mathrm{d}}p_{1}\simeq p^{*}_{1}\tau^{(1)}.\qed

3.2. Integrability

We now consider the problem of existence of prolongations and integrability.

Definition 3.7.

Given a relative algebroid (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}):

  • its first prolongation is the relative algebroid (B,p1,D(1))(B,p_{1},D^{(1)}) given by Proposition 3.3, provided it exists;

  • its kk-th prolongation is defined iteratively through

    (B(k1),pk,D(k)):=(B(k2),pk1,D(k1))(1),(B^{(k-1)},p_{k},D^{(k)}):=(B^{(k-2)},p_{k-1},D^{(k-1)})^{(1)},

    provided it exists.

The relative algebroid (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) is called kk-integrable if all the prolongations up to and including kk exist, and formally integrable if it is kk-integrable for all kk\in\mathbb{N}. If M(k)=M(k+1)M^{(k)}=M^{(k+1)} for some kk, then we say that the relative algebroid is of finite type, otherwise we say that it is of infinite type.

Our next result shows that the curvature is precisely the second order obstruction to integrability of a relative algebroid. In order to state it, note that – see also Example 1.36:

  1. (a)

    If we view τ(1)\tau^{(1)} as a classical tableau, then the cohomology group H1,2(τ(1))H^{-1,2}(\tau^{(1)}) is isomorphic to Hom(2B,)/imδ\operatorname{Hom}(\wedge^{2}B,\mathcal{F})/\operatorname{im}\delta;

  2. (b)

    If we view τ(1)\tau^{(1)} as a tableau of derivations, then the cohomology group H1,2(τ(1))H^{-1,2}(\tau^{(1)}) is isomorphic to 𝒟B2/imδ\mathcal{D}^{2}_{B}/\operatorname{im}\delta. This is where the torsion class of the first prolongation lives.

  3. (c)

    The curvature class Ω\Omega of of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) lives in H0,2(τ)=kerδ/imδH^{0,2}(\tau)=\ker\delta/\operatorname{im}\delta.

By Lemma 1.27, we have an inclusion

kerδHom(2B,)\textstyle{\ker\delta\subset\operatorname{Hom}(\wedge^{2}B,\mathcal{F})\,\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟B2\textstyle{\mathcal{D}^{2}_{B}}

which sends imδ\operatorname{im}\delta isomorphically onto imδ\operatorname{im}\delta, and hence induces an inclusion in Spencer cohomology

H0,2(τ)H1,2(τ(1)).\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 18.04823pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-18.04823pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{H^{0,2}(\tau)\,\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 18.04825pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@hook{1}}}}}}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 42.04823pt\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 42.04823pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{H^{-1,2}(\tau^{(1)})}$}}}}}}}\ignorespaces}}}}\ignorespaces. (3.2)

We can state the result about 2nd order obstructions as follows.

Theorem 3.8 (Fundamental Theorem of Prolongation).

Let (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) be a 1-integrable relative algebroid. Under the inclusion map (3.2) the torsion class of the first prolongation coincides with the curvature class of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}).

Remark 3.9.

The extension problem for relative algebroids asks: given a relative algebroid (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}), is there an extension D~\tilde{{\mathrm{D}}} of D{\mathrm{D}} to ΩB\Omega^{\bullet}_{B} such that D~2=0\tilde{{\mathrm{D}}}^{2}=0? In other words, can one complete the relative algebroid to a Lie algebroid? The curvature serves as an obstruction to this extension problem.

On the other hand, we will see later (cf. Proposition 3.19) that realizations of a relative algebroid are in a 1:1 correspondence with realizations of its first prolongation. Thus, by Proposition 2.12, the torsion of the first prolongation is an obstruction to the existence of realizations.

Therefore, the Fundamental Theorem of Prolongation implies that, at the formal level, the extension problem is equivalent to the realization problem.

Proof of Theorem 3.8.

The statement depends only pointwise on elements in M(1)M^{(1)}. So, in a neighborhood of a point in M(1)M^{(1)}, we can assume that we have a flat Ehresmann connection, which we view as bundle map h:p1TMTM(1)h\colon p_{1}^{*}TM\to TM^{(1)} satisfying

dp(h(D~m,v))=v,for all vTmM.{\mathrm{d}}p(h(\tilde{{\mathrm{D}}}_{m},v))=v,\quad\text{for all }v\in T_{m}M.

The connection hh induces a splitting of the map 𝒟B(1)1𝒟p11p1𝒟B1\mathcal{D}^{1}_{B^{(1)}}\to\mathcal{D}^{1}_{p_{1}^{*}}\cong p_{1}^{*}\mathcal{D}^{1}_{B} compatible with the symbol exact sequences

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p1Hom(2B,B)\textstyle{p_{1}^{*}\operatorname{Hom}(\wedge^{2}B,B)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}p1𝒟B1\textstyle{p_{1}^{*}\mathcal{D}^{1}_{B}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h\scriptstyle{h_{*}}p1Hom(B,TM)\textstyle{p^{*}_{1}\operatorname{Hom}(B,TM)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h\scriptstyle{h_{*}}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(2B(1),B(1))\textstyle{\operatorname{Hom}(\wedge^{2}B^{(1)},B^{(1)})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟B(1)1\textstyle{\mathcal{D}^{1}_{B^{(1)}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(B(1),TM(1))\textstyle{\operatorname{Hom}(B^{(1)},TM^{(1)})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0.\textstyle{0.}

For a point in M(1)M^{(1)}, let σ:MM(1)\sigma\colon M\to M^{(1)} be the unique local flat section through that point, so that

h(D~m,v)=dmσ(v).h(\tilde{{\mathrm{D}}}_{m},v)={\mathrm{d}}_{m}\sigma(v). (3.3)

We denote by D~σ\tilde{{\mathrm{D}}}_{\sigma} the derivation of ΩB\Omega^{\bullet}_{B} determined by σ\sigma, so

D~σ|m:=σ(m).\tilde{{\mathrm{D}}}_{\sigma}|_{m}:=\sigma(m).

(it is convenient to keep the distinction between σ\sigma as a section and σ\sigma as a derivation). The theorem will follow from the identity:

σ(hD(1)D(1))=D~σ2.\sigma^{*}\circ(h_{*}{\mathrm{D}}^{(1)}\circ{\mathrm{D}}^{(1)})=\tilde{{\mathrm{D}}}_{\sigma}^{2}. (3.4)

Note that hD(1)h_{*}{\mathrm{D}}^{(1)} is a derivation in 𝒟B(1)1\mathcal{D}^{1}_{B^{(1)}} that extends D(1){\mathrm{D}}^{(1)}, so hD(1)D(1)h_{*}{\mathrm{D}}^{(1)}\circ{\mathrm{D}}^{(1)} is a 2-derivation in 𝒟p12p1𝒟B2\mathcal{D}^{2}_{p^{*}_{1}}\cong p_{1}^{*}\mathcal{D}^{2}_{B} (see Section 1.5). Precomposing with σ\sigma^{*} is the same as pulling back this 2-derivation to σp1𝒟B2𝒟B2\sigma^{*}p_{1}^{*}\mathcal{D}^{2}_{B}\cong\mathcal{D}^{2}_{B}, so the equation makes sense.

Assume (3.4) holds. Passing to Spencer cohomology, the left hand side of this equation gives

[σ(hD(1)D(1))]=Θ(1),[\sigma^{*}\circ(h_{*}{\mathrm{D}}^{(1)}\circ{\mathrm{D}}^{(1)})]=\Theta^{(1)},

i.e., the torsion class of M(1)M^{(1)}. On the other hand, since D~σ\tilde{{\mathrm{D}}}_{\sigma} is torsionless, the right hand side of (3.4) gives the curvature class

[D~σ2]=Ω,[\tilde{{\mathrm{D}}}^{2}_{\sigma}]=\Omega,

so the theorem follows.

It remains to prove that equation (3.4) holds. For that, we only need to observe that both sides act in the same way on Ω(B,¯)1\Omega^{1}_{(B,\overline{\nabla})}, and that both derivations have the same symbol precisely because hh is flat, i.e, because of (3.3). ∎

For PDEs, Goldschmidt formulated a criterion for when a PDE is formally integrable ([20, Thm 8.1]). The following result is an analogue (or rather, extension) of that result for relative algebroids.

Theorem 3.10 (Goldschmidt’s formal integrability criterion).

Let (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) be a relative algebroid. Suppose that

  1. (i)

    (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) is 1-integrable,

  2. (ii)

    Hk,2(τ)=0H^{k,2}(\tau)=0 for all k0k\geq 0,

then (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) is formally integrable.

For the proof we need the following two lemmas.

Lemma 3.11.

Let (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) be a relative algebroid with tableau map τ\tau. Suppose that p1:M(1)Mp_{1}\colon M^{(1)}\to M is surjective and that τ(1)M\tau^{(1)}\to M has constant rank. Then (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) is 1-integrable.

Proof.

Follows from Proposition 3.5 in [20]. ∎

Lemma 3.12.

Let τ\tau be a tableau bundle such that τ(1)\tau^{(1)} has constant rank and Hk,2(τ)=0H^{k,2}(\tau)=0 for all k0k\geq 0. Then τ(k)\tau^{(k)} has constant rank for all k0k\geq 0.

Proof.

The proof of Lemma 1.5.6 in [32] holds in this setting. ∎

Proof of Theorem 3.10.

We use induction to show that the kk-th prolongation of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) exists. By assumption, the 1st prolongation exists. So assume that k1k\geq 1 and that we already know that the kk-th prolongation exists. We claim that

  1. (a)

    τ(k+1)=(τk)(1)\tau^{(k+1)}=(\tau^{k})^{(1)} has constant rank, and

  2. (b)

    pk:M(k+1)=(M(k))(1)M(k)p_{k}\colon M^{(k+1)}=(M^{(k)})^{(1)}\to M^{(k)} is surjective.

Then Lemma 3.11 shows that the kk-th prolongation is 1-integrable, so we are done.

Item (a) follows immediately from Lemma 3.12 and the assumptions in the statement. To prove item (b), note that by Theorem 3.8, under the map

H0,2(τ(k1))H1,2(τ(k)),H^{0,2}(\tau^{(k-1)})\hookrightarrow H^{-1,2}(\tau^{(k)}),

the torsion class of the kk-th prolongation takes values in H(0,2)(τ(k1))H^{(0,2)}(\tau^{(k-1)}). But,

H0,2(τ(k1))Hk,2(τ),H^{0,2}(\tau^{(k-1)})\simeq H^{k,2}(\tau),

which vanishes under our assumptions. The vanishing of this class implies that the projection pkp_{k} is surjective. ∎

Corollary 3.13.

If (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) has an involutive tableau τ\tau and vanishing torsion class, then it is formally integrable.

Definition 3.14.

A relative Lie algebroid is a formally integrable almost relative algebroid.

The appearance of the term “Lie” in the terminology is motivated by the full prolongation tower of a formally integrable relative algebroid:

(B(),D()):\textstyle{{\left(B^{(\infty)},D^{(\infty)}\right)\colon\ldots}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B(k)\textstyle{B^{(k)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B(k1)\textstyle{B^{(k-1)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D(k)\scriptstyle{{\mathrm{D}}^{(k)}}\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B(1)\textstyle{B^{(1)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}B\textstyle{B\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D(1)\scriptstyle{{\mathrm{D}}^{(1)}}M():\textstyle{M^{(\infty)}\colon\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M(k)\textstyle{M^{(k)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pk\scriptstyle{p_{k}}M(k1)\textstyle{M^{(k-1)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M(1)\textstyle{M^{(1)}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p1\scriptstyle{p_{1}}M\textstyle{M}

The derivation D(){\mathrm{D}}^{(\infty)}, defined on profinite sections of B()M()B^{(\infty)}\to M^{(\infty)} (i.e., on those that locally factor through some M(k)M^{(k)}) does square to zero:

(D())2=0,\left({\mathrm{D}}^{(\infty)}\right)^{2}=0,

so (B(),D())(B^{(\infty)},{\mathrm{D}}^{(\infty)}) is a “profinite Lie algebroid of finite rank”. We leave a deeper study of these objects for future work (see also Section 7 for further discussion).

3.3. Some examples

In this section, we discuss several simple examples that illustrate the various issues that can arise with prolongations of relative algebroids. We provide examples of relative algebroids that: (i) are of infinite type, (ii) are of finite type, (iii) do not admit a prolongation, and (iv) have a first prolongation but not a second prolongation.

While such examples already exist in the context of PDEs, the ones presented here are independent of PDE theory. These examples should also help the reader develop intuition for the general framework.

Example 3.15 (Relative vector fields and control systems).

If a relative algebroid is 1-integrable and has an involutive tableau with non-zero Cartan characters, it is always of infinite type. Such examples, with the smallest rank, arise from relative vector fields.

A vector field relative to a submersion p:MNp\colon M\to N is a section XΓ(pTN)X\in\Gamma(p^{*}TN). It gives rise to a relative algebroid

¯\textstyle{\underline{\mathbb{R}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}¯\textstyle{\underline{\mathbb{R}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}DX\scriptstyle{{\mathrm{D}}_{X}}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}N\textstyle{N}

where the derivation is given by

DX:C(N)Ω1(¯),DX(f)m=Xm(f)dt,{\mathrm{D}}_{X}\colon C^{\infty}(N)\to\Omega^{1}(\underline{\mathbb{R}}),\quad D_{X}(f)_{m}=X_{m}(f){\mathrm{d}}t,

where dt{\mathrm{d}}t is a basis of \mathbb{R}^{*}. Conversely, any relative algebroid with vector bundle ¯\underline{\mathbb{R}} gives rise to a relative vector field. In particular, the prolongation of a relative vector field is again a relative vector field.

For a relative vector field XX, the tableau map of the corresponding algebroid is

τ=¯X:\displaystyle\tau={\overline{\nabla}}X\colon kerdpHom(,pTN)pTN,\displaystyle\ker{\mathrm{d}}p\to\operatorname{Hom}(\mathbb{R},p^{*}TN)\cong p^{*}TN,
τ(v)(λ):\displaystyle\tau(v)(\lambda): =λddtXγ(t)|t=0\displaystyle=\lambda\left.\frac{{\mathrm{d}}}{{\mathrm{d}}t}X_{\gamma(t)}\right|_{t=0}

where γ(t)\gamma(t) is a curve in a fiber of pp with γ˙(0)=v\dot{\gamma}(0)=v. The Spencer differential δτ:Hom(,kerdp)Hom(2,pTN)\delta_{\tau}\colon\operatorname{Hom}(\mathbb{R},\ker{\mathrm{d}}p)\to\operatorname{Hom}(\wedge^{2}\mathbb{R},p^{*}TN) vanishes identically, and therefore the tableau is involutive, with Cartan character

s1=rank(kerdp)=dimMdimN.s_{1}=\operatorname{rank}(\ker{\mathrm{d}}p)=\dim M-\dim N.

A lift of DX{\mathrm{D}}_{X} to ¯M\underline{\mathbb{R}}\to M is a vector field X~\tilde{X} on MM such that dp(X~)=X{\mathrm{d}}p(\tilde{X})=X. The corresponding derivation D~X\tilde{{\mathrm{D}}}_{X} always squares to zero, so the relative algebroid has vanishing torsion and curvature, and is therefore formally integrable.

A local realization can be described explicitly as a curve γM:IM\gamma^{M}\colon I\to M such that γN:=pγM\gamma^{N}:=p\circ\gamma^{M} satisfies the ODE

γ˙N(t)=XγM(t).\dot{\gamma}^{N}(t)=X_{\gamma^{M}(t)}. (3.5)

In local coordinates, if (xμ,yϱ)(x^{\mu},y^{\varrho}) are submersion coordinates on MM, then the vector field looks like

X=Xμ(x,y)xμ,X=X^{\mu}(x,y)\frac{\partial}{\partial x^{\mu}},

so the ODE (3.5) takes the form

x˙μ(t)=Xμ(x(t),y(t)).\dot{x}^{\mu}(t)=X^{\mu}(x(t),y(t)). (3.6)

This shows that locally, realizations (integral curves) are completely determined by the choice of the dimMdimN\dim M-\dim N family of functions yϱ(t)y^{\varrho}(t) and the choice of an initial point (x0,y0)(x_{0},y_{0}). Equation (3.6) is a control system (with unspecified observables), where the choice of the functions yϱ(t)y^{\varrho}(t) is the control input. In particular, according to the results in [22], when the vector field is real analytic, then NN is partitioned into a singular foliation with the property that for each point n0Nn_{0}\in N, the integral curves of XX through n0n_{0} saturate a neighborhood of n0n_{0} inside the leaf through n0n_{0}.

As a very particular case, consider p:p\colon\mathbb{R}\to* with the zero relative vector field. A realization of the corresponding relative algebroid is just an \mathbb{R}-valued function x1(t)x^{1}(t) defined on an interval. The prolongation tower consists of a sequence of vector fields XkX_{k} relative to the projection pk:kk1p_{k}\colon\mathbb{R}^{k}\to\mathbb{R}^{k-1},

(¯,DX):\textstyle{{\left(\underline{\mathbb{R}},{\mathrm{D}}_{X_{\infty}}\right)\colon\ldots}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}¯\textstyle{\underline{\mathbb{R}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}¯\textstyle{\underline{\mathbb{R}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}DXk1\scriptstyle{{\mathrm{D}}_{X_{k-1}}}\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}¯\textstyle{\underline{\mathbb{R}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}¯\textstyle{\underline{\mathbb{R}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}DX0\scriptstyle{{\mathrm{D}}_{X_{0}}}:\textstyle{\mathbb{R}^{\infty}\colon\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}k\textstyle{\mathbb{R}^{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pk\scriptstyle{p_{k}}k1\textstyle{\mathbb{R}^{k-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\mathbb{R}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}\textstyle{*}

Denoting by (x1,,xk)(x^{1},\dots,x^{k}) the coordinates on k\mathbb{R}^{k}, the relative vector fields XkX_{k} are given by

Xk=i=1kxi+1xi,X=i=1xi+1xi.X_{k}=\sum_{i=1}^{k}x^{i+1}\frac{\partial}{\partial x^{i}},\quad X_{\infty}=\sum_{i=1}^{\infty}x^{i+1}\frac{\partial}{\partial x^{i}}.

The profinite vector field XX_{\infty} appears later in Section 6.

Example 3.16 (Finite-type relative algebroid).

Consider the submersion p:32,(x,y,z)(x,y)p\colon\mathbb{R}^{3}\to\mathbb{R}^{2},(x,y,z)\mapsto(x,y) and the relative algebroid (¯2,p,D)(\underline{\mathbb{R}}^{2},p,{\mathrm{D}}) determined by

{Dθ1=θ1θ2,Dθ2=0,Dx=zθ1,Dy=zθ2.\begin{cases}{\mathrm{D}}\theta^{1}=\theta^{1}\wedge\theta^{2},\\ {\mathrm{D}}\theta^{2}=0,\\ {\mathrm{D}}x=z\theta^{1},\\ {\mathrm{D}}y=z\theta^{2}.\end{cases}

To compute the prolongation, we find an extension D~\tilde{{\mathrm{D}}} of D{\mathrm{D}} satisfying

0\displaystyle 0 =D~(Dx)=D~zθ1+zθ1θ2\displaystyle=\tilde{{\mathrm{D}}}({\mathrm{D}}x)=\tilde{{\mathrm{D}}}z\wedge\theta^{1}+z\theta^{1}\wedge\theta^{2}
0\displaystyle 0 =D~(Dy)=D~zθ2.\displaystyle=\tilde{{\mathrm{D}}}({\mathrm{D}}y)=\tilde{{\mathrm{D}}}z\wedge\theta^{2}.

Writing D~z=αθ1+βθ2\tilde{{\mathrm{D}}}z=\alpha\theta^{1}+\beta\theta^{2}, we see that we must have

D~z=zθ2.\tilde{{\mathrm{D}}}z=z\theta^{2}.

This defines a prolongation where M(1)M=3M^{(1)}\cong M=\mathbb{R}^{3}, so D~\tilde{{\mathrm{D}}} is a derivation relative to the identity. At this point, we only know that D~D=0\tilde{{\mathrm{D}}}\circ{\mathrm{D}}=0, but one checks easily that D~2=0\tilde{{\mathrm{D}}}^{2}=0. Hence, the first prolongation is actually a Lie algebroid and so M(k)M(1)M^{(k)}\simeq M^{(1)} for all k1k\geq 1. The resulting algebroid (2¯,D~)3(\underline{\mathbb{R}^{2}},\tilde{{\mathrm{D}}})\to\mathbb{R}^{3} is isomorphic to an action algebroid 𝔤3\mathfrak{g}\ltimes\mathbb{R}^{3}, where 𝔤\mathfrak{g} is the non-abelian two-dimensional Lie algebra.

Example 3.17 (1-Integrable relative algebroid that is not 2-integrable).

Let us modify the derivation in the previous example by setting

{Dθ1=θ1θ2,Dθ2=0,Dx=zθ1,Dy=zθ2θ1.\begin{cases}{\mathrm{D}}\theta^{1}=\theta^{1}\wedge\theta^{2},\\ {\mathrm{D}}\theta^{2}=0,\\ {\mathrm{D}}x=z\theta^{1},\\ {\mathrm{D}}y=z\theta^{2}-\theta^{1}.\end{cases}

Proceeding as before, the first prolongation D~\tilde{{\mathrm{D}}} is now given by

D~z=θ1+zθ2.\tilde{{\mathrm{D}}}z=\theta^{1}+z\theta^{2}.

However, in this case D~2z=2θ1θ20\tilde{{\mathrm{D}}}^{2}z=2\theta^{1}\wedge\theta^{2}\neq 0, so the second prolongation does not exist.

Example 3.18 (Torsion from tableau of derivations).

In the previous example, the equations determining the prolongation and obstructions to integrability arose solely from the symbol tableau of the relative algebroid. We now illustrate how the full tableau of derivations can give rise to torsion. Of course, this can only happen when the rank is at least three, otherwise the symbol torsion completely determines the torsion. One way to achieve this is to consider relative algebroids with zero anchor. So, consider the algebroid relative to (¯3,p,D)(\underline{\mathbb{R}}^{3},p,{\mathrm{D}}), with p:p\colon\mathbb{R}\to* and D{\mathrm{D}} determined by

{Dθ1=xθ1θ2,Dθ2=xθ2θ3,Dθ3=xθ3θ1,\begin{cases}{\mathrm{D}}\theta^{1}=x\theta^{1}\wedge\theta^{2},\\ {\mathrm{D}}\theta^{2}=x\theta^{2}\wedge\theta^{3},\\ {\mathrm{D}}\theta^{3}=x\theta^{3}\wedge\theta^{1},\end{cases} (3.7)

where xx in the coordinate on \mathbb{R}. A lift D~\tilde{{\mathrm{D}}} of D{\mathrm{D}} is determined by D~x=x1θ1+x2θ2+x3θ3\tilde{{\mathrm{D}}}x=x_{1}\theta^{1}+x_{2}\theta^{2}+x_{3}\theta^{3} and the torsion is the 2-derivation (D~D)(\tilde{{\mathrm{D}}}\circ D) relative to 3×\mathbb{R}^{3}\times\mathbb{R}\to* defined by

{D~Dθ1=(x3+x)θ1θ2θ3,D~Dθ2=(x1+x)θ1θ2θ3,D~Dθ3=(x2+x)θ1θ2θ3.\begin{cases}\tilde{{\mathrm{D}}}\circ{\mathrm{D}}\theta^{1}=(x_{3}+x)\theta^{1}\wedge\theta^{2}\wedge\theta^{3},\\ \tilde{{\mathrm{D}}}\circ{\mathrm{D}}\theta^{2}=(x_{1}+x)\theta^{1}\wedge\theta^{2}\wedge\theta^{3},\\ \tilde{{\mathrm{D}}}\circ{\mathrm{D}}\theta^{3}=(x_{2}+x)\theta^{1}\wedge\theta^{2}\wedge\theta^{3}.\end{cases}

We find that the first prolongation space is

M(1)={x1=x2=x3=x},M^{(1)}=\{x_{1}=x_{2}=x_{3}=-x\},

so the first prolongation is given by

D(1)x=x(θ1+θ2+θ3).{\mathrm{D}}^{(1)}x=-x(\theta^{1}+\theta^{2}+\theta^{3}).

Note that the curvature is only zero at x=0x=0:

(D(1))2x=x(θ1θ2+θ2θ3+θ3θ1).\left({\mathrm{D}}^{(1)}\right)^{2}x=-x(\theta^{1}\wedge\theta^{2}+\theta^{2}\wedge\theta^{3}+\theta^{3}\wedge\theta^{1}).

Hence, to obtain an actual Lie algebroid one has to restrict to the space x=0x=0, in which case the relative algebroid is just the abelian Lie algebra 3\mathbb{R}^{3}.

3.4. Realizations and prolongations

Prolongations arose as a tool for constructing and computing obstructions to the existence of realizations. The realizations of the prolongations are related to realizations of the original relative algebroid in the following manner.

Proposition 3.19.

Let (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) be a relative algebroid. Then:

  1. (i)

    If (B1,p1,D1)(B_{1},p_{1},{\mathrm{D}}_{1}) is a prolongation, then any realization of (B1,p1,D1)(B_{1},p_{1},{\mathrm{D}}_{1}) induces a realization of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}).

  2. (ii)

    If (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) is 1-integrable, then realizations of the canonical prolongation (B(1),p1,D(1))(B^{(1)},p_{1},{\mathrm{D}}^{(1)}) are in 1-1 correspondence with realizations of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}).

Proof.

To prove (i), let (B1,p1,D1)(B_{1},p_{1},{\mathrm{D}}_{1}) be a prolongation and let (P,r1,θ1)(P,r_{1},\theta_{1}) a realization of (B1,p1,D1)(B_{1},p_{1},{\mathrm{D}}_{1}). The fact that (P,p1r1,(p1)θ1)(P,p_{1}\circ r_{1},(p_{1})_{*}\circ\theta_{1}) is a realization of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) follows by restricting to Ω(B,¯)1\Omega^{1}_{(B,\overline{\nabla})} the identity valid on ΩB1\Omega^{1}_{B}:

d(θ1)=(θ1)D1.{\mathrm{d}}\circ\left(\theta_{1}\right)^{*}=\left(\theta_{1}\right)^{*}\circ{\mathrm{D}}_{1}.

To prove (ii), let (P,r,θ)(P,r,\theta) be a realization of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}). By Proposition 2.12, we can define the map

r(1):PM(1),r(1)(p)=D~θp:=(θp1)(dθ|p).r^{(1)}\colon P\to M^{(1)},\quad r^{(1)}(p)=\tilde{{\mathrm{D}}}_{\theta_{p}}:=\left(\theta^{-1}_{p}\right)^{*}\circ\left({\mathrm{d}}\theta^{*}\big{|}_{p}\right).

Clearly, p1r(1)=rp_{1}\circ r^{(1)}=r, and since B(1)=p1BB^{(1)}=p_{1}^{*}B there is a natural bundle map

(θ(1),r(1)):TPB(1), with (p1)θ(1)=θ.(\theta^{(1)},r^{(1)})\colon TP\to B^{(1)},\text{ with }(p_{1})_{*}\circ\theta^{(1)}=\theta.

Using that θ\theta is a realization, we find that

d(θ(1))p1=dθ=θD=(θ(1))D(1)p1, on Ω(B,¯).{\mathrm{d}}\circ\left(\theta^{(1)}\right)^{*}\circ p_{1}^{*}={\mathrm{d}}\circ\theta^{*}=\theta^{*}\circ{\mathrm{D}}=\left(\theta^{(1)}\right)^{*}\circ{\mathrm{D}}^{(1)}\circ p_{1}^{*},\mbox{ on $\Omega^{\bullet}_{(B,\overline{\nabla})}$.}

From our formula for r(1)r^{(1)} and the tautological nature of D(1){\mathrm{D}}^{(1)} it also follows that

d(r(1)))(f)=(θ(1))D(f), for fC(M),{\mathrm{d}}\circ\left(r^{(1)})^{*}\right)(f)=\left(\theta^{(1)}\right)^{*}\circ{\mathrm{D}}(f),\mbox{ for $f\in C^{\infty}(M)$,}

so (r(1),θ(1))(r^{(1)},\theta^{(1)}) is a realization of B(1)B^{(1)}. The construction in (i) when applied to the 1st prolongation is an inverse to this construction, so we have a 1:1 correspondence. ∎

Theorem 3.20 (Cartan-Bryant [8, Thm. 3 and Thm. 4]).

Let (B,¯,)(B,\overline{\nabla},\mathcal{F}) be an analytic relative Lie algebroid. Then, for each kk and each xM(k)x\in M^{(k)}, there exists a realization through xx.

Proof.

Since, by assumption, (B,¯,)(B,\overline{\nabla},\mathcal{F}) is formally integrable, induction and Proposition 3.19 shows that it is enough to prove the result for k=1k=1. Moreover, since every tableau is involutive after finitely many prolongations, we can assume that the tableau τ\tau is involutive. But then, the realization problem for B(1)B^{(1)} in local analytic coordinates (see Remark 2.8 and Example 3.5) satisfies the assumptions of [8, Thm. 3]). The latter result shows that there exists a realization through every point in M(1)M^{(1)}. ∎

3.5. Naturality

Let (φ,p):(B1,¯1)(B2,¯2)(\varphi,p)\colon(B_{1},{\overline{\nabla}}^{1})\to(B^{2},{\overline{\nabla}}^{2}) be a morphism of flat foliated vector bundles covering a map p:(M1,1)(M2,2)p\colon(M_{1},\mathcal{F}_{1})\to(M_{2},\mathcal{F}_{2}), such that φ\varphi is a fiberwise isomorphism. As we saw in Section 1.5, φ\varphi induces a bundle map

φ:𝒟(B1,¯1)k𝒟(B2,¯2)k\varphi_{*}\colon\mathcal{D}^{k}_{(B_{1},{\overline{\nabla}}^{1})}\to\mathcal{D}^{k}_{(B_{2},{\overline{\nabla}}^{2})}

More generally, in this situation, there is a bundle map

φ:Hom(lB1,𝒟(B1,¯1)k)Hom(lB2,𝒟(B2,¯2)k)\varphi_{*}\colon\operatorname{Hom}\left(\wedge^{l}B_{1},\mathcal{D}^{k}_{(B_{1},{\overline{\nabla}}^{1})}\right)\to\operatorname{Hom}\left(\wedge^{l}B_{2},\mathcal{D}^{k}_{(B_{2},{\overline{\nabla}}^{2})}\right)

that intertwines the Spencer differentials: δφ=φδ\delta\circ\varphi_{*}=\varphi_{*}\circ\delta. All constructions with derivations that we discussed before behave naturally relative to these induced maps.

Proposition 3.21.

Let (φ,p):(B1,¯1,D1)(B2,¯2,D2)(\varphi,p)\colon(B_{1},{\overline{\nabla}}^{1},{\mathrm{D}}_{1})\to(B_{2},{\overline{\nabla}}^{2},{\mathrm{D}}_{2}) be a morphism of relative algebroids, which is a fiberwise isomorphism. Then

  1. (i)

    The tableau maps are (φ,p)(\varphi_{*},p_{*})-related: φτ1=τ2p\varphi_{*}\circ\tau_{1}=\tau_{2}\circ p_{*}. In particular, φ\varphi induces a morphism in Spencer cohomology

    φ:Hk,l(τ1)Hk,l(τ2),\varphi_{*}\colon H^{k,l}(\tau_{1})\to H^{k,l}(\tau_{2}),

    that maps the torsion class of B1B_{1} to the one of B2B_{2};

  2. (ii)

    If (P,r,θ)(P,r,\theta) is a realization of (B1,¯1,D1)(B_{1},{\overline{\nabla}}^{1},{\mathrm{D}}_{1}), then (P,pr,φθ)(P,p\circ r,\varphi\circ\theta) is a realization of (B2,¯2,D2)(B_{2},{\overline{\nabla}}^{2},{\mathrm{D}}_{2});

  3. (iii)

    If (B1,¯1,D1)(B_{1},{\overline{\nabla}}^{1},{\mathrm{D}}_{1}) and (B2,¯2,D2)(B_{2},{\overline{\nabla}}^{2},{\mathrm{D}}_{2}) are kk-integrable, then φ\varphi induces a morphism of relative algebroids (φ(k),p(k)):(B1(k),(p1)k,D1(k))(B2(k),(p2)k,D2(k))(\varphi^{(k)},p^{(k)})\colon(B_{1}^{(k)},(p_{1})_{k},{\mathrm{D}}^{(k)}_{1})\to(B_{2}^{(k)},(p_{2})_{k},{\mathrm{D}}^{(k)}_{2}) that is fiberwise an isomorphism;

  4. (iv)

    If (B1,¯1,D1)(B_{1},{\overline{\nabla}}^{1},{\mathrm{D}}_{1}) and (B2,¯2,D2)(B_{2},{\overline{\nabla}}^{2},{\mathrm{D}}_{2}) are formally integrable, then φ\varphi induces a morphism (φ(),p()):(B1(),D1())(B2(),D2())(\varphi^{(\infty)},p^{(\infty)})\colon(B_{1}^{(\infty)},{\mathrm{D}}^{(\infty)}_{1})\to(B_{2}^{(\infty)},{\mathrm{D}}^{(\infty)}_{2}) of profinite Lie algebroids.

Notice that if (B,p1,D1)(B,p_{1},{\mathrm{D}}_{1}) is a prolongation of a relative algebroid (B,¯,D)(B,{\overline{\nabla}},{\mathrm{D}}), then the torsion class of (B,p1,D1)(B,p_{1},{\mathrm{D}}_{1}) is in the kernel of (p1)(p_{1})_{*}. Hence, the proposition implies:

Corollary 3.22.

If a relative algebroid (B,¯,D)(B,{\overline{\nabla}},{\mathrm{D}}) admits a prolongation (B,p1,D1)(B,p_{1},{\mathrm{D}}_{1}), then its torsion class must vanish.

Proof.

First, we show that φ\varphi_{*} and pp_{*} intertwine the tableau maps. For that, notice that if αΩ(B1,¯1)\alpha\in\Omega_{(B_{1},{\overline{\nabla}}^{1})} then, for any XX\in\mathcal{F}, we have

(¯X1D1)α=¯X1(D1α).({\overline{\nabla}}^{1}_{X}{\mathrm{D}}_{1})\alpha={\overline{\nabla}}^{1}_{X}({\mathrm{D}}_{1}\alpha).

Using this and the definition of the tableaux, we find for αΩ(B2,¯2)\alpha\in\Omega^{\bullet}_{(B_{2},{\overline{\nabla}}^{2})}:

(φτ1(X))α\displaystyle(\varphi_{*}\tau_{1}(X))\alpha =(φ(¯X1D1))α=φ(¯X1(D1φα))\displaystyle=\left(\varphi_{*}\left({\overline{\nabla}}^{1}_{X}D_{1}\right)\right)\alpha=\varphi_{*}\left({\overline{\nabla}}^{1}_{X}\left({\mathrm{D}}_{1}\varphi^{*}\alpha\right)\right)
=φ(¯X1φ(D2α))=φ(φ¯p(X)2(D2α))\displaystyle=\varphi_{*}\left({\overline{\nabla}}^{1}_{X}\varphi^{*}({\mathrm{D}}_{2}\alpha)\right)=\varphi_{*}\left(\varphi^{*}{\overline{\nabla}}^{2}_{p_{*}(X)}({\mathrm{D}}_{2}\alpha)\right)
=¯p(X)2(D2α)=(¯p(X)2D2)(α)=τ2(p(X))(α).\displaystyle={\overline{\nabla}}^{2}_{p_{*}(X)}({\mathrm{D}}_{2}\alpha)=({\overline{\nabla}}^{2}_{p_{*}(X)}{\mathrm{D}}_{2})(\alpha)=\tau_{2}(p_{*}(X))(\alpha).

Since φ\varphi_{*} intertwines the Spencer differentials, it induces maps on the Spencer complexes of τ1\tau_{1} and τ2\tau_{2}, commuting with the differentials, and therefore φ\varphi_{*} descends to the level of cohomology. To check that it relates the torsion classes, let D~1\tilde{{\mathrm{D}}}_{1} be a pointwise lift of D1{\mathrm{D}}_{1} above xM1x\in M_{1}. Its torsion is the 2-derivation T1|D~1=D~1DT^{1}\big{|}_{\tilde{{\mathrm{D}}}_{1}}=\tilde{{\mathrm{D}}}_{1}\circ{\mathrm{D}}. Because D1{\mathrm{D}}_{1} and D2{\mathrm{D}}_{2} are φ\varphi-related, the derivation φ(D~1)\varphi_{*}(\tilde{{\mathrm{D}}}_{1}) is a pointwise lift of D2{\mathrm{D}}_{2} above p(x)M2p(x)\in M_{2}. The 2-derivations T1|D~1T^{1}\big{|}_{\tilde{{\mathrm{D}}}_{1}} and T2|D~2T^{2}\big{|}_{\tilde{{\mathrm{D}}}_{2}} are also φ\varphi_{*}-related since

φ(T1|D~1)(α)\displaystyle\varphi_{*}\left(T^{1}\big{|}_{\tilde{{\mathrm{D}}}_{1}}\right)(\alpha) =φ(D~1D1)(α)=φD~1D1φ(α)\displaystyle=\varphi_{*}(\tilde{{\mathrm{D}}}_{1}\circ{\mathrm{D}}_{1})(\alpha)=\varphi_{*}\circ\tilde{{\mathrm{D}}}_{1}\circ{\mathrm{D}}_{1}\circ\varphi^{*}(\alpha)
=φD~1φD2(α)=φ(D~1)D2(α)\displaystyle=\varphi_{*}\circ\tilde{{\mathrm{D}}}_{1}\circ\varphi^{*}\circ{\mathrm{D}}_{2}(\alpha)=\varphi_{*}(\tilde{{\mathrm{D}}}_{1})\circ{\mathrm{D}}_{2}(\alpha)
=D~2D2(α)=T2|D~2(α),\displaystyle=\tilde{{\mathrm{D}}}_{2}\circ{\mathrm{D}}_{2}(\alpha)=T^{2}\big{|}_{\tilde{{\mathrm{D}}}_{2}}(\alpha),

for any αΩ(B2,¯2)\alpha\in\Omega^{\bullet}_{(B_{2},{\overline{\nabla}}^{2})}. Passing down to the Spencer cohomology item (i) follows.

Item (ii) follows by observing that, for αΩ(B2,¯2)\alpha\in\Omega^{\bullet}_{(B_{2},{\overline{\nabla}}^{2})}, one has

(φθ)D2α\displaystyle(\varphi\circ\theta)^{*}{\mathrm{D}}_{2}\alpha =θφD2α=θD1φα=d(φθ)α.\displaystyle=\theta^{*}\varphi^{*}{\mathrm{D}}_{2}\alpha=\theta^{*}{\mathrm{D}}_{1}\varphi^{*}\alpha={\mathrm{d}}(\varphi\circ\theta)^{*}\alpha.

In order to prove item (iii) it is enough to prove the case k=1k=1. By the previous calculation, if D~1\tilde{{\mathrm{D}}}_{1} is a torsionless lift of D1{\mathrm{D}}_{1}, then φ(D~1)\varphi_{*}(\tilde{{\mathrm{D}}}_{1}) is also a torsionless lift of D2{\mathrm{D}}_{2}. So φ\varphi_{*} restricts to a map between the base spaces of the first prolongation p(1):=φ:M1(1)M2(2)p^{(1)}:=\varphi_{*}\colon M_{1}^{(1)}\to M_{2}^{(2)}, satisfying

(p2)1p(1)=(p2)1φ=p(p1)1.(p_{2})_{1}\circ p^{(1)}=(p_{2})_{1}\circ\varphi_{*}=p\circ(p_{1})_{1}.

This map and φ\varphi combine into a map between the pullbacks Bi(1)=(pi)1BiB^{(1)}_{i}=(p_{i})_{1}^{*}B_{i}:

φ(1):=(φ,p(1)):B1(1)B2(1).\varphi^{(1)}:=(\varphi,p^{(1)})\colon B^{(1)}_{1}\to B^{(1)}_{2}.

Note that φ(1)\varphi^{(1)} is again a fiberwise isomorphism and also a map of relative algebroids because of the tautological nature of the derivations D1(1){\mathrm{D}}^{(1)}_{1} and D2(1){\mathrm{D}}_{2}^{(1)}. Indeed, we find

D1(1)(φ(α))|D~1\displaystyle{\mathrm{D}}^{(1)}_{1}(\varphi^{*}(\alpha))\big{|}_{\tilde{{\mathrm{D}}}_{1}} =D~1(φα)=φ(φ(D~1)(α))\displaystyle=\tilde{{\mathrm{D}}}_{1}(\varphi^{*}\alpha)=\varphi^{*}\big{(}\varphi_{*}(\tilde{{\mathrm{D}}}_{1})(\alpha)\big{)}
=φ(D2(1)(α)|φ(D~1))=(φ(1))(D2(1)(α))\displaystyle=\varphi^{*}\left({\mathrm{D}}^{(1)}_{2}(\alpha)\big{|}_{\varphi_{*}(\tilde{{\mathrm{D}}}_{1})}\right)=(\varphi^{(1)})^{*}\big{(}{\mathrm{D}}^{(1)}_{2}(\alpha)\big{)}

for any αΩ(B2,¯2)\alpha\in\Omega^{\bullet}_{(B_{2},{\overline{\nabla}}^{2})}, where we used that p(1)=φp^{(1)}=\varphi_{*}.

Finally, item (iv) follows from item (iii). ∎

4. Constructions

To better understand relative algebroids, we will now develop several important constructions involving them.

4.1. The universal relative algebroid

Let ANA\to N be any vector bundle, and p1:𝒟AkNp_{1}\colon\mathcal{D}^{k}_{A}\to N its bundle of kk-derivations. There is a tautological kk-derivation D˘\breve{\mathrm{D}} relative to the projection p1:𝒟AkNp_{1}\colon\mathcal{D}^{k}_{A}\to N, which at each point Dx(𝒟Ak)x{\mathrm{D}}_{x}\in\left(\mathcal{D}^{k}_{A}\right)_{x} is the derivation itself:

D˘(α)(Dx)=Dxαk+lAx,for αΩl(A).\breve{\mathrm{D}}(\alpha)({\mathrm{D}}_{x})={\mathrm{D}}_{x}\alpha\in\wedge^{k+l}A_{x}^{*},\quad\text{for }\alpha\in\Omega^{l}(A).
Definition 4.1.

The universal relative Lie algebroid of a vector bundle ANA\to N is the triple (A,p1,D˘)(A,p_{1},\breve{\mathrm{D}}), where p1:𝒟A1Np_{1}\colon\mathcal{D}^{1}_{A}\to N and D˘\breve{\mathrm{D}} is the tautological 1-derivation.

The bundle over 𝒟A1\mathcal{D}^{1}_{A} of the universal relative algebroid will be denoted B˘:=p1A\breve{B}:=p_{1}^{*}A. In order to justify the use of the term “universal”, let us introduce the following notation.

Definition 4.2.

Let (A,p,D)(A,p,{\mathrm{D}}) be an algebroid relative to a submersion p:MNp\colon M\to N. The classifying map cDc_{{\mathrm{D}}} of (A,p,D)(A,p,{\mathrm{D}}) is the composition

MDcDp𝒟A1pr𝒟A1.\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 8.39583pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&&\crcr}}}\ignorespaces{\hbox{\kern-8.39583pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 14.72221pt\raise-5.39166pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-2.39166pt\hbox{$\scriptstyle{{\mathrm{D}}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 32.39583pt\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\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}}{}}}}\ignorespaces{}\ignorespaces{}{}{}{{}{}}\ignorespaces\ignorespaces{\hbox{\kern 38.4003pt\raise 23.19028pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-0.82361pt\hbox{$\scriptstyle{c_{{\mathrm{D}}}}$}}}\kern 3.0pt}}}}}}\ignorespaces{}{}{}{{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}{}}{\hbox{\kern 80.66096pt\raise 6.04568pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{}{{}}{{}{}{}\lx@xy@spline@}{}}}}\ignorespaces{}\ignorespaces\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{{}{}{{}}{{}{}{}}{}}}}\ignorespaces{}{\hbox{\kern 32.39583pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{p^{*}\mathcal{D}^{1}_{A}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 60.39288pt\raise-5.18748pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-0.8264pt\hbox{$\scriptstyle{\operatorname{pr}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 80.66597pt\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 80.66597pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathcal{D}^{1}_{A}}$}}}}}}}\ignorespaces}}}}\ignorespaces.
Proposition 4.3.

Every relative algebroid (A,p,D)(A,p,{\mathrm{D}}) is canonically isomorphic to the pullback of the universal relative Lie algebroid (A,p1,D˘)(A,p_{1},\breve{\mathrm{D}}) along its classifying map.

Proof.

Note that p1cD=pp_{1}\circ c_{D}=p, and that the pullback cDD˘c_{D}^{*}\breve{\mathrm{D}} coincides with D{\mathrm{D}} under the canonical identification cDp1𝒟A1p𝒟A1c^{*}_{D}p_{1}^{*}\mathcal{D}^{1}_{A}\cong p^{*}\mathcal{D}^{1}_{A}. ∎

Example 4.4.

When N={}N=\{*\}, so A=VA=V is a vector space, we have

𝒟V1Hom(2V,V).\mathcal{D}^{1}_{V}\cong\operatorname{Hom}(\wedge^{2}V,V)\to*.

A relative algebroid (V,p,D)(V,p,{\mathrm{D}}), for p:M{}p\colon M\to\{*\}, is the same thing as a skew-symmetric bilinear map [,]:V×VC(M,V)[\cdot,\cdot]\colon V\times V\to C^{\infty}(M,V), viewed a bracket on VV relative to pp. This is determined by a map cD:MHom(2V,V)c_{{\mathrm{D}}}\colon M\to\operatorname{Hom}(\wedge^{2}V,V), cD(x)(v,w)=[v,w](x)c_{{\mathrm{D}}}(x)(v,w)=[v,w](x), which is precisely the classifying map. This construction was already considered by Bryant in relation to his Theorem 4 in [8].

Example 4.5.

For almost Lie algebroids, the previous proposition says that any such structure D{\mathrm{D}} on a fixed vector bundle p:ANp\colon A\to N is obtained by pulling back the tautological derivation D˘\breve{\mathrm{D}} along a section cD:M𝒟A1c_{{\mathrm{D}}}\colon M\to\mathcal{D}^{1}_{A}: cDD˘=Dc_{{\mathrm{D}}}^{*}\breve{\mathrm{D}}={\mathrm{D}}.

Example 4.6.

The classifying map of the first prolongation of a 1-integrable relative algebroid (B,¯,D)(B,{\overline{\nabla}},{\mathrm{D}}) is the inclusion M(1)𝒟B1M^{(1)}\hookrightarrow\mathcal{D}^{1}_{B}. In fact, given an algebroid (B,p1,D1)(B,p_{1},{\mathrm{D}}_{1}) relative to a submersion p1:M1Mp_{1}\colon M_{1}\to M with classifying map cD1:M1𝒟B1c_{{\mathrm{D}}_{1}}\colon M_{1}\to\mathcal{D}^{1}_{B}, one has that:

  1. (i)

    (B,p1,D1)(B,p_{1},{\mathrm{D}}_{1}) is an extension of (B,¯,D)(B,{\overline{\nabla}},{\mathrm{D}}) if and only if cD1c_{{\mathrm{D}}_{1}} takes values in the space of pointwise lifts L𝒟B1L\subset\mathcal{D}^{1}_{B};

  2. (ii)

    (B,p1,D1)(B,p_{1},{\mathrm{D}}_{1}) is a prolongation of (B,¯,D)(B,{\overline{\nabla}},{\mathrm{D}}) if and only if the image of cD1c_{{\mathrm{D}}_{1}} is contained in M(1)M^{(1)}.

In particular, among all prolongations of (B,¯,D)(B,{\overline{\nabla}},{\mathrm{D}}), the canonical prolongation is the universal one.

Next, we will study properties of the universal relative Lie algebroid. We start by justifying the use of the term “Lie”.

Proposition 4.7.

The universal relative Lie algebroid of a vector bundle ANA\to N is formally integrable.

Proof.

The tableau of the universal algebroid is the identity map 𝒟A1𝒟A1\mathcal{D}^{1}_{A}\to\mathcal{D}^{1}_{A}, so it is involutive by Proposition 1.48. The torsion class vanishes automatically because H1,2(𝒟A1)=0H^{-1,2}(\mathcal{D}^{1}_{A})=0 by Lemma 1.49. Then, by Corollary 3.13 of Goldschmidt’s Formal Integrability Criterion, the universal algebroid is formally integrable. ∎

The realizations of the universal relative Lie algebroid have a nice geometric interpretation. For that, given a manifold PP and a vector bundle ANA\to N, by an AA-coframe on PP we mean a bundle map θA:TPA\theta_{A}\colon TP\to A covering a map rN:PNr_{N}\colon P\to N which is fiberwise an isomorphism. These objects have appeared in the literature under the name “generalized coframes” in the context of Dirac spinors coupled to Einstein’s equations – see [29].

Proposition 4.8.

The realizations of the universal relative Lie algebroid of ANA\to N are in one-to-one correspondence with manifolds equipped with AA-coframes.

Proof.

In one direction, it is clear that a realization (θ,r):TPp1A(\theta,r)\colon TP\to p_{1}^{*}A of the universal relative Lie algebroid (A,p,D˘)(A,p,\breve{\mathrm{D}}), gives rise to the AA-frame θA:TPA\theta_{A}\colon TP\to A covering rN:=p1rr_{N}:=p_{1}\circ r given by

Φ=prAθ.\Phi=\operatorname{pr}_{A}\circ\,\theta.

In the opposite direction, given an AA-frame θA:TPA\theta_{A}\colon TP\to A covering a map rN:PNr_{N}\colon P\to N, we can define a bundle map (θ,r):TPp1A(\theta,r)\colon TP\to p_{1}^{*}A which is a fiberwise isomorphism by

{r(p):=DrN(p),θ(vp):=(DrN(p),θA(vp)),(vpTpP),\begin{cases}r(p):={\mathrm{D}}_{r_{N}(p)},\\ \theta(v_{p}):=({\mathrm{D}}_{r_{N}(p)},\theta_{A}(v_{p})),\end{cases}\quad(v_{p}\in T_{p}P),

where DrN(p)(𝒟A1)rN(p){\mathrm{D}}_{r_{N}(p)}\in(\mathcal{D}^{1}_{A})_{r_{N}(p)} is given by

DrN(p)α:=(θA)dpθA(α),(αΩ1(A)).{\mathrm{D}}_{r_{N}(p)}\alpha:=(\theta_{A})_{*}\circ{\mathrm{d}}_{p}\circ\theta_{A}^{*}(\alpha),\quad(\alpha\in\Omega^{1}(A)).

This ensures that the condition

θD˘=dθp1\theta^{*}\circ\breve{\mathrm{D}}={\mathrm{d}}\circ\theta^{*}\circ p_{1}^{*}

holds, so (θ,r)(\theta,r) is a realization of (A,p1,D˘).(A,p_{1},\breve{\mathrm{D}}).

These constructions are inverse to each other, so the proposition follows. ∎

Recall that a relative algebroid (A,p,D)(A,p,{\mathrm{D}}) is called standard when its tableau map τ:kerdpp𝒟A1\tau\colon\ker{\mathrm{d}}p\to p^{*}\mathcal{D}^{1}_{A} is fiberwise injective.

Proposition 4.9.

An algebroid (A,p,D)(A,p,{\mathrm{D}}) relative to a submersion p:MNp\colon M\to N is standard if and only if its classifying map cD:M𝒟A1c_{{\mathrm{D}}}\colon M\to\mathcal{D}^{1}_{A} is an immersion.

Proof.

Note that p=p1cDp=p_{1}\circ c_{{\mathrm{D}}}. This means that the classifying map cD:M𝒟A1c_{{\mathrm{D}}}\colon M\to\mathcal{D}^{1}_{A} is an immersion if and only if its vertical derivative dcD|kerdp:kerdpkerdp1{\mathrm{d}}c_{{\mathrm{D}}}\big{|}_{\ker{\mathrm{d}}p}\colon\ker{\mathrm{d}}p\to\ker{\mathrm{d}}p_{1} is fiberwise injective. We claim that the tableau map τ:kerdpp𝒟A1\tau\colon\ker{\mathrm{d}}p\to p^{*}\mathcal{D}^{1}_{A} is given by

τ(Xm)=(m,dcD(Xm)),(Xmkerdmp),\tau(X_{m})=(m,{\mathrm{d}}c_{{\mathrm{D}}}(X_{m})),\quad(X_{m}\in\ker{\mathrm{d}}_{m}p), (4.1)

so the result follows. To prove the above formula, just observe that we have, by Proposition 4.3,

τ(X)m=¯XmD=¯Xm(cDD˘)=¯dcD(Xm)D˘=dcD(Xm).\tau(X)_{m}={\overline{\nabla}}_{X_{m}}{\mathrm{D}}={\overline{\nabla}}_{X_{m}}(c_{{\mathrm{D}}}^{*}\breve{\mathrm{D}})={\overline{\nabla}}_{{\mathrm{d}}c_{{\mathrm{D}}}(X_{m})}\breve{\mathrm{D}}={\mathrm{d}}c_{\mathrm{D}}(X_{m}).\qed

4.2. Restriction

Almost relative algebroids are so flexible that they can easily be restricted to subspaces. From the point of view of Cartan’s realization problem, imposing restrictions on a relative algebroid structure is equivalent to adding extra conditions to the realization problem of the original relative algebroid. This will be clear from examples at the end of this section. For now, let us be precise about what we mean by “restriction”.

In this section we fix an ambient relative algebroid (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}).

Definition 4.10.

A map q:QMq\colon Q\to M is invariant for (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) if the image of dq{\mathrm{d}}q contains the image of the anchor of D{\mathrm{D}}, i.e., if for every xQx\in Q,

σ(Dq(x))Hom(Bq(x),im(dxq)/q(x))Hom(Bq(x),ν(q(x))).\sigma({\mathrm{D}}_{q(x)})\in\operatorname{Hom}(B_{q(x)},\operatorname{im}({\mathrm{d}}_{x}q)/\mathcal{F}_{q(x)})\subseteq\operatorname{Hom}(B_{q(x)},\nu(\mathcal{F}_{q(x)})).
Remark 4.11.

In the special case of an algebroid (A,p,D)(A,p,{\mathrm{D}}) relative to a submersion p:MNp\colon M\to N, the invariance condition on a map q:QMq\colon Q\to M says that

imρp(x)imTx(pq), for all xQ.\operatorname{im}\rho_{p(x)}\subseteq\operatorname{im}T_{x}(p\circ q),\quad\mbox{ for all $x\in Q$.}

If (A,D)(A,D) is an almost Lie algebroid qq is an immersion, we recover the usual notion of invariant submanifold for such an algebroid.

Proposition 4.12.

Suppose that a map q:QMq\colon Q\to M satisfies the following conditions:

  1. (a)

    qq is invariant for DD;

  2. (b)

    (dxq)1(q(x))(d_{x}q)^{-1}(\mathcal{F}_{q(x)}) has constant rank for all xQx\in Q.

Then there exists a unique relative algebroid (qB,q¯,DQ)(q^{*}B,q^{*}\overline{\nabla},{\mathrm{D}}_{Q}) for which the map q:qBBq_{*}\colon q^{*}B\to B is a morphism of relative algebroids.

Remark 4.13.

The morphism q:(qB,q¯,DQ)(B,¯,D)q_{*}\colon(q^{*}B,q^{*}{\overline{\nabla}},{\mathrm{D}}_{Q})\to(B,{\overline{\nabla}},{\mathrm{D}}) is a fiberwise an isomorphism, so it follows from from Proposition 3.21 that:

  1. (i)

    realizations of (qB,q¯,DQ)(q^{*}B,q^{*}\overline{\nabla},{\mathrm{D}}_{Q}) yield realizations of (B,¯,D)(B,{\overline{\nabla}},{\mathrm{D}});

  2. (ii)

    there is an induced morphism at level of the Spencer cohomologies of the tableaux q:Hk,l(τQ)Hk,l(τ)q_{*}\colon H^{k,l}(\tau_{Q})\to H^{k,l}(\tau), which relates the torsion classes;

  3. (iii)

    if both algebroids are kk-integrable, there is an induced algebroid morphism q:((qB)(k),(pQ)k,DQ(k))(B(k),pk,D(k))q_{*}\colon((q^{*}B)^{(k)},(p_{Q})_{k},{\mathrm{D}}_{Q}^{(k)})\to(B^{(k)},p_{k},{\mathrm{D}}^{(k)})

  4. (iv)

    if both algebroids are formally integrable, there is an induced algebroid morphism q:((qB),DQ)(B,D)q_{*}\colon((q^{*}B)^{\infty},{\mathrm{D}}_{Q}^{\infty})\to(B^{\infty},{\mathrm{D}}^{\infty}).

Note also that if q:QMq\colon Q\to M is transverse to \mathcal{F}, then both conditions (a) and (b) in the proposition are automatically satisfied.

Proof.

The regularity condition (b) means that that the pullback foliation q!q^{!}\mathcal{F} exists. It follows that one has a flat foliated vector bundle (qB,q¯)(q^{*}B,q^{*}\overline{\nabla}) over (Q,q!)(Q,q^{!}\mathcal{F}), as well as a map of foliated vector bundles

q:=(q,pr):(qB,q¯)(B,¯).q_{*}:=(q,\operatorname{pr})\colon(q^{*}B,q^{*}\overline{\nabla})\to(B,{\overline{\nabla}}).

Since this is a fiberwise isomorphism, as we saw in Section 1.5, it induces a bundle map

q:𝒟(qB,q¯)1𝒟(B,¯)1.q_{*}\colon\mathcal{D}^{1}_{(q^{*}B,q^{*}\overline{\nabla})}\to\mathcal{D}^{1}_{(B,{\overline{\nabla}})}.

which covers qq.

It is easy to check that the short exact sequence in Lemma 1.25 is natural in qq_{*}. This means that there is a commutative diagram of short exact sequences

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(2qB,qB)\textstyle{\operatorname{Hom}(\wedge^{2}q^{*}B,q^{*}B)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q_{*}}𝒟(qB,q¯)1\textstyle{\mathcal{D}^{1}_{(q^{*}B,q^{*}\overline{\nabla})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q_{*}}Hom(qB,ν(q!))\textstyle{\operatorname{Hom}(q^{*}B,\nu(q^{!}\mathcal{F}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q_{*}}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}qHom(2B,B)\textstyle{q^{*}\operatorname{Hom}(\wedge^{2}B,B)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q𝒟(B,¯)1\textstyle{q^{*}\mathcal{D}^{1}_{(B,\overline{\nabla})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}qHom(B,ν())\textstyle{q^{*}\operatorname{Hom}(B,\nu(\mathcal{F}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

Both vertical arrows on the sides are injective, and therefore so is the the middle vertical arrow. Since q:QMq\colon Q\to M is an invariant map, the section qDq^{*}{\mathrm{D}} is in the image of qq_{*}, so there is a unique relative derivation DQ{\mathrm{D}}_{Q} on (Q,q!,qB,q¯)(Q,q^{!}\mathcal{F},q^{*}B,q^{*}\overline{\nabla}) that satisfies q(DQ)=qDq_{*}({\mathrm{D}}_{Q})=q^{*}{\mathrm{D}}. This completes the proof. ∎

Definition 4.14.

If qq is an injective immersion that satisfies the conditions of Proposition 4.12, we call (qB,q¯,DQ)(q^{*}B,q^{*}\overline{\nabla},D_{Q}) the restriction of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) to QQ.

Example 4.15.

If (A,p,D)(A,p,{\mathrm{D}}) is a 1-integrable relative algebroid, then its first prolongation (B,p1,D(1))(B,p_{1},D^{(1)}) is the restriction of the tautological relative algebroid (A,p1,D˘)(A,p_{1},\breve{\mathrm{D}}) to M(1)𝒟A1M^{(1)}\hookrightarrow\mathcal{D}^{1}_{A}.

It should be noted that, in general, restriction does not preserve any kind of integrability of the relative algebroid. From the point of view of the realization problem, restriction amounts to add extra equations to the problem. So a problem that originally had realizations may stop having them.

The vanishing locus of the anchor is always invariant.

Proposition 4.16.

Let (B,¯,D)(B,{\overline{\nabla}},{\mathrm{D}}) be a relative algebroid and let QMQ\subset M be a submanifold along which ρ\rho vanishes. Then QQ is an invariant submanifold.

Proof.

In this case, imρx=0\operatorname{im}\rho_{x}=0 for all xQx\in Q, so QQ is clearly invariant. ∎

If (B(),D())(B^{(\infty)},{\mathrm{D}}^{(\infty)}) is the prolongation tower of a relative algebroid (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}), then the kernel of the anchor at each point x0kerρ()x_{0}\in\ker\rho^{(\infty)} is a Lie algebra. This is the Lie algebra of a group of symmetries of a realization whose image contains p(x0)Mp_{\infty}(x_{0})\in M. So, in general, the larger the group of symmetries of a realization of a geometric problem is, the larger the kernel of the anchor of the corresponding relative algebroid must be.

Example 4.17 (Submanifolds where anchor has constant rank).

In general, submanifolds along which the rank of the anchor is constant are not invariant, unless involutivity conditions are imposed.

For example, take the algebroid ¯22\underline{\mathbb{R}}^{2}\to\mathbb{R}^{2} (relative to the identity), with derivation defined by

{Dθ1=Dθ2=0,Dx=yθ1,Dy=θ2.\begin{cases}{\mathrm{D}}\theta^{1}={\mathrm{D}}\theta^{2}=0,\\ {\mathrm{D}}x=-y\theta^{1},\\ {\mathrm{D}}y=\theta^{2}.\end{cases}

The corresponding anchor satisfies ρ(e1)=yx\rho(e_{1})=-y\partial_{x}, ρ(e2)=y\rho(e_{2})=\partial_{y}. Therefore, the submanifold Q={rankρ=1}={y=0}Q=\{\operatorname{rank}\rho=1\}=\{y=0\} is not invariant. In this example, there are no realizations anywhere, as D2x=θ1θ20{\mathrm{D}}^{2}x=\theta^{1}\wedge\theta^{2}\neq 0.

Understanding manifolds along which the anchor has constant rank requires studying the full prolongation tower. This is more delicate and will be discussed in future work.

Example 4.18 (Universal bundle of Lie algebras).

The universal bundle of Lie algebra structures on a vector space VV can be constructed, in an ad hoc manner, as the bundle of Lie algebras – a Lie algebroid with zero anchor – V¯𝔤^\underline{V}\to\hat{\mathfrak{g}} where

𝔤^={cHom(2V,V)|Jacc=0}\hat{\mathfrak{g}}=\{c\in\operatorname{Hom}(\wedge^{2}V,V)\ |\ \operatorname{Jac}_{c}=0\}

with Jacc(v1,v2,v3)=c(v1,c(v2,v3))+c.p.\operatorname{Jac}_{c}(v_{1},v_{2},v_{3})=c(v_{1},c(v_{2},v_{3}))+\mathrm{c.p.} for v1,v2,v3Vv_{1},v_{2},v_{3}\in V. The bracket on V¯𝔤^\underline{V}\to\hat{\mathfrak{g}} is the tautological bracket: [v1,v2](c)=c(v1,v2)[v_{1},v_{2}](c)=c(v_{1},v_{2}). A realization of this algebroid above a point c𝔤^c\in\hat{\mathfrak{g}} is a local Lie group integrating the Lie algebra (V,c)(V,c).

This Lie algebroid arises by imposing maximal symmetry on the classifying algebroid for all VV-coframes. Namely, let p1:𝒟V1Hom(2V,V)p_{1}\colon\mathcal{D}^{1}_{V}\cong\operatorname{Hom}(\wedge^{2}V,V)\to* be the tautological relative algebroid corresponding to a vector space VV. Consider its prolongation (p1V,p2,D(1))(p_{1}^{*}V,p_{2},{\mathrm{D}}^{(1)}), relative to the projection p2:(𝒟V1)(1)𝒟V1p_{2}\colon(\mathcal{D}^{1}_{V})^{(1)}\to\mathcal{D}^{1}_{V}, where

(𝒟V1)(1)={ξcp1Hom(V,Hom(2V,V))|Jacc=δξc},p2(ξc)=c.(\mathcal{D}^{1}_{V})^{(1)}=\big{\{}\xi_{c}\in p_{1}^{*}\operatorname{Hom}(V,\operatorname{Hom}(\wedge^{2}V,V))\ |\ \operatorname{Jac}_{c}=\delta\xi_{c}\big{\}},\quad p_{2}(\xi_{c})=c.

The anchor of this prolongation at ξc(𝒟V1)(1)\xi_{c}\in(\mathcal{D}^{1}_{V})^{(1)} vanishes if and only if ξc=0\xi_{c}=0. The restriction of p2p_{2} to this invariant subvariety gives a canonical identification {ρ(1)=0}𝔤^\{\rho^{(1)}=0\}\cong\hat{\mathfrak{g}}. The restricted relative algebroid is precisely V¯𝔤^\underline{V}\to\hat{\mathfrak{g}}, the universal bundle of Lie algebra structures on VV.

Realizations of the relative algebroid V¯𝔤^\underline{V}\to\hat{\mathfrak{g}} correspond to VV-coframes with maximal symmetry: these are just local Lie groups with their Maurer-Cartan forms!

Example 4.19 (Jacobi manifolds).

Fixing a vector space VV, Bryant in [8] defines a Jacobi manifold as a submanifold M𝒟V1Hom(2V,V)M\subset\mathcal{D}^{1}_{V}\cong\operatorname{Hom}(\wedge^{2}V,V) such that

Jaccδ(Hom(V¯,TcM)),for all cM,\operatorname{Jac}_{c}\in\delta(\operatorname{Hom}(\underline{V},T_{c}M)),\quad\mbox{for all $c\in M$},

where δ:Hom(V,𝒟V1)𝒟V2\delta\colon\operatorname{Hom}(V,\mathcal{D}^{1}_{V})\to\mathcal{D}^{2}_{V} is the Spencer differential. In our language, this is the same as saying that the restriction of the tautological algebroid to MM has vanishing torsion class. Theorem 4 in [8] states that, when MM is real analytic, if TcM𝒟V1T_{c}M\subset\mathcal{D}^{1}_{V} is an involutive tableau of derivations, then there exists a realization of the restricted algebroid through every point in MM.

Example 4.20 (Riemannian metrics).

As we recalled in the introductory section, the orthonormal frame bundle of a Riemannian manifold has a canonical coframe (θ,ω)(\theta,\omega) with values in n𝔬(n)\mathbb{R}^{n}\oplus\mathfrak{o}(n) satisfying the Cartan’s structure equations (0.1). This suggests that the realization problem for (locally orthonormal frame bundles of) Riemannian metrics can be obtained by restricting 𝒟n𝔬(𝔫)1\mathcal{D}^{1}_{\mathbb{R}^{n}\oplus\mathfrak{o(n)}} to derivations of the form

{Dθ=ωθ,Dω=R(θθ)ωω\begin{cases}{\mathrm{D}}\theta=-\omega\wedge\theta,\\ {\mathrm{D}}\omega=R(\theta\wedge\theta)-\omega\wedge\omega\end{cases} (4.2)

The space of such derivations is an affine subspace of 𝒟n𝔬(𝔫)1\mathcal{D}^{1}_{\mathbb{R}^{n}\oplus\mathfrak{o(n)}} parametrized by RHom(2n,𝔬(n))R\in\operatorname{Hom}(\wedge^{2}\mathbb{R}^{n},\mathfrak{o}(n)).

The resulting restricted algebroid has a non-zero torsion class, and therefore the algebroid needs to be restricted further to the subspace where the torsion class vanishes. The torsion class can be computed by taking an extension D~\tilde{{\mathrm{D}}} of D{\mathrm{D}}. Such extension is completely determined by D~R\tilde{{\mathrm{D}}}R, and to find this value one applies D~\tilde{{\mathrm{D}}} to (4.2):

0\displaystyle 0 =D2θ=Dωθ+ωDθ=R(θθ)θ\displaystyle={\mathrm{D}}^{2}\theta=-{\mathrm{D}}\omega\wedge\theta+\omega\wedge{\mathrm{D}}\theta=R(\theta\wedge\theta)\wedge\theta
0\displaystyle 0 =D~Dω=D~R(θθ)+R(Dθθ)R(θDθ)Dωω+ωDω\displaystyle=\tilde{{\mathrm{D}}}{\mathrm{D}}\omega=\tilde{{\mathrm{D}}}R\wedge(\theta\wedge\theta)+R({\mathrm{D}}\theta\wedge\theta)-R(\theta\wedge{\mathrm{D}}\theta)-{\mathrm{D}}\omega\wedge\omega+\omega\wedge{\mathrm{D}}\omega
=D~R(θθ)\displaystyle=\tilde{{\mathrm{D}}}R\wedge(\theta\wedge\theta)

The torsion class can be identified with R(θθ)θR(\theta\wedge\theta)\wedge\theta, and its vanishing is precisely the first (algebraic) Bianchi identity.

The vanishing of the curvature class, when imposing 𝔬(n)\mathfrak{o}(n)-invariance of the prolongation, gives rise to the second Bianchi identity. A precise formulation requires the notion of a relative GG-structure algebroid, which will be the subject of future work (see Section 7).

4.3. Systatic foliation and reduction

Recall that a relative algebroid is standard when its tableau map τ\tau is fiberwise injective. As we saw in the previous section, this happens precisely when the classifying map is an immersion. We will now show that the directions in which the tableau map is zero – i.e., the directions in which the classifying map is constant – are essentially “redundant” from the perspective of the realization problem. The notion of systastic space and inessential invariants goes back to Cartan – a modern formulation and discussion can be found in [16].

Again, in this section we fix a relative algebroid (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) over (M,)(M,\mathcal{F}). We also assume that the kernel of its tableau map τ:𝒟(B,¯)1\tau\colon\mathcal{F}\to\mathcal{D}^{1}_{(B,{\overline{\nabla}})} has constant rank.

Definition 4.21.

The systatic foliation of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) is the foliation

sys:=kerτ,\mathcal{F}_{\text{sys}}:=\ker\tau\subset\mathcal{F},

where τ\tau is the tableau map of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}).

Note that sys\mathcal{F}_{\text{sys}} is involutive because the \mathcal{F}-connection ¯\overline{\nabla} induced on 𝒟(B,¯)1\mathcal{D}^{1}_{(B,\overline{\nabla})} is flat. Next, we present two useful characterizations of the systatic foliation.

Proposition 4.22.

For a a relative algebroid (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) the systatic foliation is given by

sys={X:¯X[b1,b2]=0, for all b1,b2Γ(B,¯) },\mathcal{F}_{\mathrm{sys}}=\{X\in\mathcal{F}:\overline{\nabla}_{X}[b_{1},b_{2}]=0,\mbox{ for all $b_{1},b_{2}\in\Gamma_{(B,\overline{\nabla})}$ }\},

where [,][\cdot,\cdot] denotes the bracket associated to D{\mathrm{D}}.

Proof.

By (2.4), for each XX\in\mathcal{F}, we have

¯XD=0¯X[b1,b2]=0, for all b1,b2Γ(B,¯).{\overline{\nabla}}_{X}{\mathrm{D}}=0\quad\Longleftrightarrow\quad\overline{\nabla}_{X}[b_{1},b_{2}]=0,\text{ for all }b_{1},b_{2}\in\Gamma_{(B,\overline{\nabla})}.\qed
Proposition 4.23.

The systatic foliation of an algebroid (A,p,D)(A,p,{\mathrm{D}}) relative to a submersion coincides with the connected components of the fibers of the classifying map cDc_{{\mathrm{D}}}.

Proof.

By (4.1), we have kerτ=kerdcD\ker\tau=\ker{\mathrm{d}}c_{{\mathrm{D}}}, so the result follows. ∎

Let us now discuss how to get rid of directions along the systatic foliation.

Definition 4.24.

Let (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) be a relative algebroid. A foliation 0sys\mathcal{F}_{0}\subset\mathcal{F}_{\mathrm{sys}} is called inessential if it is a simple foliation and ¯\overline{\nabla} has no holonomy along its leaves.

Notice that a inessential foliation of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) is given by the fibers of a submersion q:MMredq\colon M\to M_{\mathrm{red}}. The foliation \mathcal{F} descends to a foliation red\mathcal{F}_{\mathrm{red}} of the leaf space MredM_{\mathrm{red}} characterized by

=(dq)1(red).\mathcal{F}=({\mathrm{d}}q)^{-1}(\mathcal{F}_{\mathrm{red}}).

Moreover, by Corollary 1.17, there is a unique flat foliated bundle (Bred,¯)(B_{\mathrm{red}},{\overline{\nabla}}) over MredM_{\mathrm{red}} whose pullback under q:MMredq\colon M\to M_{\mathrm{red}} is (B,¯)(B,{\overline{\nabla}}).

We use these notations in the statement of the following theorem showing that the derivation D{\mathrm{D}} can also be reduced, preserving all its essential properties.

Theorem 4.25 (Reduction).

Let (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) be a relative algebroid and let 0sys\mathcal{F}_{0}\subset\mathcal{F}_{\mathrm{sys}} be an inessential foliation with leaf space q:MMredq\colon M\to M_{\mathrm{red}}. Then:

  1. (i)

    There exists a unique structure of a relative algebroid algebroid (Bred,¯,Dred)(B_{\mathrm{red}},\overline{\nabla},{\mathrm{D}}_{\mathrm{red}}) such that the q:BBredq_{*}\colon B\to B_{\mathrm{red}} is a morphism of relative algebroids;

  2. (ii)

    The tableaux τ\tau and τred\tau_{\mathrm{red}} of (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) and (Bred,¯,Dred)(B_{\mathrm{red}},\overline{\nabla},{\mathrm{D}}_{\mathrm{red}}), have naturally isomorphic Spencer cohomologies;

  3. (iii)

    The relative algebroid (B,¯,D)(B,\overline{\nabla},{\mathrm{D}}) is kk-integrable if and only if (Bred,¯,Dred)(B_{\mathrm{red}},\overline{\nabla},{\mathrm{D}}_{\mathrm{red}}) is kk-integrable;

  4. (iv)

    There is a one-to-one correspondence

    {realizations (P,r,θ)of (B,¯,D)}1:1{realizations (P,r~,θ~) of (Bred,¯,Dred)together with a lift r of r~: MqPrr~Mred }.\left\{\begin{subarray}{c}\mbox{realizations $(P,r,\theta)$}\\ \\ \mbox{of $(B,\overline{\nabla},{\mathrm{D}})$}\end{subarray}\right\}\overset{1:1}{\longleftrightarrow}\left\{\begin{subarray}{c}\mbox{realizations $(P,\tilde{r},\tilde{\theta})$ of $(B_{\mathrm{red}},\overline{\nabla},{\mathrm{D}}_{\mathrm{red}})$}\\ \mbox{together with a lift $r$ of $\tilde{r}$:}\\ \mbox{ $\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 5.73315pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\\&\crcr}}}\ignorespaces{\hbox{\kern-3.0pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{}$}}}}}}}{\hbox{\kern 32.46083pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 39.2379pt\raise-13.23193pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-0.57848pt\hbox{$\scriptstyle{q}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 39.2379pt\raise-18.68056pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern-5.73315pt\raise-26.46388pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{P\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces{}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 12.62526pt\raise-9.1771pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-1.05486pt\hbox{$\scriptstyle{r}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 34.79063pt\raise-3.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}\ignorespaces\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces{\hbox{\lx@xy@drawline@}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{}{\hbox{\lx@xy@droprule}}\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 13.84116pt\raise-33.075pt\hbox{{}\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\hbox{\hbox{\kern 0.0pt\raise-3.61111pt\hbox{$\scriptstyle{\tilde{r}}$}}}\kern 3.0pt}}}}}}\ignorespaces{\hbox{\kern 29.73315pt\raise-26.46388pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@tip{1}\lx@xy@tip{-1}}}}}}{\hbox{\lx@xy@droprule}}{\hbox{\lx@xy@droprule}}{\hbox{\kern 29.73315pt\raise-26.46388pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{M_{\mathrm{red}}}$}}}}}}}\ignorespaces}}}}\ignorespaces$ }\end{subarray}\right\}.

Moreover, if 0=sys\mathcal{F}_{0}=\mathcal{F}_{\mathrm{sys}} then the reduced algebroid (Bred,¯,Dred)(B_{\mathrm{red}},\overline{\nabla},{\mathrm{D}}_{\mathrm{red}}) is standard.

Remark 4.26.

The reduced algebroid (Bred,¯,Dred)(B_{\mathrm{red}},{\overline{\nabla}},{\mathrm{D}}_{\mathrm{red}}) can be seen as a quotient of the relative algebroid (B,¯,D)(B,{\overline{\nabla}},{\mathrm{D}}) by the pseudogroup generated by the flows of vector fields tangent to the inessential foliation 0\mathcal{F}_{0}. A detailed discussion of such pseudogroups of symmetries is left for future work (see Section 7.1).

Remark 4.27.

The prolongation of the reduction can not be a (systatic) reduction of the prolongation. The reason is that the prolongation is always standard (see Lemma 3.6). Intuitively, the prolongation before reduction adds extra equations that (locally) encode a map from a realization to the fibers of qq, while this is lost in the prolongation of the reduction. However, since the morphism q:(B,¯,D)(Bred,¯,Dred)q_{*}\colon(B,\overline{\nabla},{\mathrm{D}})\to(B_{\mathrm{red}},\overline{\nabla},{\mathrm{D}}_{\mathrm{red}}) is a fiberwise isomorphism, it follows from from Proposition 3.21 that, assuming kk-integrability, there is an induced morphism between the kk-prolongations.

Proof.

To prove item (i), note that the map of flat foliated vector bundles

q:(B,¯)(Bred,¯)q_{*}\colon(B,\overline{\nabla})\to(B_{\mathrm{red}},\overline{\nabla})

is a fiberwise isomorphism. It follows from Section 1.5 that we have a bundle map

q:𝒟(B,¯)1𝒟(Bred,¯)1.q_{*}\colon\mathcal{D}^{1}_{(B,\overline{\nabla})}\to\mathcal{D}^{1}_{(B_{\mathrm{red}},\overline{\nabla})}.

We claim that there is a unique Dred{\mathrm{D}}_{\mathrm{red}} which is (q,q)(q,q_{*})-related to D{\mathrm{D}}, so (i) holds.

To prove this claim, note that we have a commutative diagram of short exact sequences

0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(k+1B,B)\textstyle{\operatorname{Hom}(\wedge^{k+1}B,B)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q_{*}}𝒟(B,¯)k\textstyle{\mathcal{D}^{k}_{(B,\overline{\nabla})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q_{*}}Hom(kB,ν())\textstyle{\operatorname{Hom}(\wedge^{k}B,\nu(\mathcal{F}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q_{*}}0\textstyle{0}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(k+1Bred,Bred)\textstyle{\operatorname{Hom}(\wedge^{k+1}B_{\mathrm{red}},B_{\mathrm{red}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟(Bred,¯)k\textstyle{\mathcal{D}^{k}_{(B_{\mathrm{red}},\overline{\nabla})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(kBred,ν(red))\textstyle{\operatorname{Hom}(\wedge^{k}B_{\mathrm{red}},\nu(\mathcal{F}_{\mathrm{red}}))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}0\textstyle{0}

Since the sides are fiberwise isomorphisms, so is the map in the middle. Now, because ker(dq)kerτ\ker({\mathrm{d}}q)\subseteq\ker\tau, we have that ¯XD=0{\overline{\nabla}}_{X}{\mathrm{D}}=0 for any Xker(dq)X\in\ker({\mathrm{d}}q), and since the holonomy of ¯{\overline{\nabla}} vanishes along the directions of ker(dq)\ker({\mathrm{d}}q), we conclude that the section DΓ(𝒟(B,¯)1{\mathrm{D}}\in\Gamma(\mathcal{D}^{1}_{(B,{\overline{\nabla}})} is the pullback of a section DredΓ(𝒟(Bred,)1){\mathrm{D}}_{\mathrm{red}}\in\Gamma(\mathcal{D}^{1}_{(B_{\mathrm{red}},\nabla)}). This also means that D{\mathrm{D}} and Dred{\mathrm{D}}_{\mathrm{red}} are (q,q)(q,q_{*})-related, so the claim follows.

In order to prove item (ii), we look at the relationship between the tableau maps of BB and BredB_{\mathrm{red}}. According to Proposition 3.21, the map qq_{*} intertwines the tableau maps and the Spencer differentials. For k1k\geq 1 this amounts to the commutativity of the following diagram where the rows are short exact sequences:

Hom(lB,Hom(SkB,ker(dq))\textstyle{{\operatorname{Hom}(\wedge^{l}B,\operatorname{Hom}(S^{k}B,\ker({\mathrm{d}}q))}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δ\scriptstyle{\delta}Hom(lB,τ(k))\textstyle{{\operatorname{Hom}(\wedge^{l}B,\tau^{(k)})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q_{*}}δ\scriptstyle{\delta}qHom(lBred,τred(k))\textstyle{{q^{*}\operatorname{Hom}(\wedge^{l}B_{\mathrm{red}},\tau^{(k)}_{\mathrm{red}})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δ\scriptstyle{\delta}Hom(l+1B,Hom(Sk1B,ker(dq))\textstyle{{\operatorname{Hom}(\wedge^{l+1}B,\operatorname{Hom}(S^{k-1}B,\ker({\mathrm{d}}q))}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(l+1B,τ(k1))\textstyle{{\operatorname{Hom}(\wedge^{l+1}B,\tau^{(k-1)})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q_{*}}qHom(l+1Bred,τred(k1))\textstyle{{q^{*}\operatorname{Hom}(\wedge^{l+1}B_{\mathrm{red}},\tau^{(k-1)}_{\mathrm{red}})}}

while for k=0k=0 one has the commutative diagram

Hom(lB,ker(dq))\textstyle{{\operatorname{Hom}(\wedge^{l}B,\ker({\mathrm{d}}q))}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hom(lB,)\textstyle{{\operatorname{Hom}(\wedge^{l}B,\mathcal{F})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q_{*}}δτ\scriptstyle{\delta_{\tau}}qHom(lBred,red)\textstyle{{q^{*}\operatorname{Hom}(\wedge^{l}B_{\mathrm{red}},\mathcal{F}_{\mathrm{red}})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}δτred\scriptstyle{\delta_{\tau_{\mathrm{red}}}}0\textstyle{0\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝒟(B,¯)l+1\textstyle{{\mathcal{D}^{l+1}_{(B,{\overline{\nabla}})}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}q\scriptstyle{q_{*}}q𝒟(Bred,¯)l+1\textstyle{{q^{*}\mathcal{D}^{l+1}_{(B_{\mathrm{red}},{\overline{\nabla}})}}}

Setting l=1l=1, we conclude that τ(k)\tau^{(k)} has constant rank if and only if τred(k)\tau_{\mathrm{red}}^{(k)} has constant rank. Moreover, since the restriction τ|ker(dq)\tau\big{|}_{\ker({\mathrm{d}}q)} is the zero tableau map, which is involutive, the map qq_{*} descends to an isomorphism in cohomology:

q:Hk,l(τ)qHk,l(τred).q_{*}\colon H^{k,l}(\tau)\xrightarrow{\sim}q^{*}H^{k,l}(\tau_{\mathrm{red}}).

Now, using this isomorphism, item (iii) also follows: by Proposition 3.21 (i), torsion classes of the kk-th prolongations are qq_{*}-related; hence, if BB and BredB_{\mathrm{red}} are (k1)(k-1)-integrable, then BB is kk-integrable if and only if BredB_{\mathrm{red}} is kk-integrable.

Finally, to prove item (iv), observe that by Proposition 3.21 (ii) every realization (P,r,θ)(P,r,\theta) of (B,¯,D)(B,{\overline{\nabla}},{\mathrm{D}}) induces a realization (P,r~,θ~)(P,\tilde{r},\tilde{\theta}) of (Bred,¯,Dred)(B_{\mathrm{red}},{\overline{\nabla}},{\mathrm{D}}_{\mathrm{red}}) with r~=qr\tilde{r}=q\circ r. Conversely, given a a realization (P,r~,θ~)(P,\tilde{r},\tilde{\theta}) of (Bred,¯,Dred)(B_{\mathrm{red}},{\overline{\nabla}},{\mathrm{D}}_{\mathrm{red}}) and a lift r:PMr\colon P\to M of r~\tilde{r}, using the fact that q:𝒟(B,¯)1q𝒟(Bred,¯)1q_{*}\colon\mathcal{D}^{1}_{(B,\overline{\nabla})}\to q^{*}\mathcal{D}^{1}_{(B_{\mathrm{red}},\overline{\nabla})} is an isomorphism of vector bundles, there is a unique vector bundle map (θ,r):TPB(\theta,r)\colon TP\to B such that θ~=qθ\tilde{\theta}=q_{*}\circ\theta. Moreover, since qq_{*} is a morphism of relative algebroids, we have

q(dθθD)\displaystyle q_{*}({\mathrm{d}}\circ\theta^{*}-\theta^{*}\circ{\mathrm{D}}) =dθqθDq\displaystyle={\mathrm{d}}\circ\theta^{*}\circ q^{*}-\theta^{*}\circ{\mathrm{D}}\circ q^{*}
=d(qθ)(qθ)Dred=dθ~θ~D=0.\displaystyle={\mathrm{d}}\circ(q_{*}\circ\theta)^{*}-(q_{*}\circ\theta)^{*}\circ{\mathrm{D}}_{\mathrm{red}}={\mathrm{d}}\circ\tilde{\theta}^{*}-\tilde{\theta}\circ{\mathrm{D}}=0.

Since qq_{*} is a fiberwise isomorphism we must have dθ=θD{\mathrm{d}}\circ\theta^{*}=\theta^{*}\circ{\mathrm{D}}, so (θ,r)(\theta,r) is a morphism of almost relative algebroids.

These constructions are inverse to each other, so this completes the proof. ∎

5. Relative connections and PDEs

We will now show that any partial differential equation can be recast as a relative algebroid in such a way that the formal theory of prolongations [20] coincides with the prolongation theory for relative algebroids. The relative algebroid of a PDE arises from a relative connection, so we start by discussing this notion.

5.1. Relative connections and the Cartan distribuition

Let N𝑞XN\xrightarrow{q}X be a submersion. A connection for qq is a splitting TNker(dq)qTXTN\cong\ker({\mathrm{d}}q)\oplus q^{*}TX. Sometimes, however, the splitting does not arise on the level of NN, but rather depends on additional coordinates. Let us illustrate this with an important example.

Example 5.1.

Let q:NXq\colon N\to X be any submersion. The bundle p1:J1qNp_{1}\colon J^{1}q\to N of first jets of local sections can be identified with the bundle of horizontal compliments of ker(dq)\ker({\mathrm{d}}q) in TNTN, that is

(J1q)n{CnTnN:Cnker(dnq)=TnN}.(J^{1}q)_{n}\cong\{C_{n}\subset T_{n}N:C_{n}\oplus\ker({\mathrm{d}}_{n}q)=T_{n}N\}.

There is no canonical splitting of TNTN into the components ker(dq)\ker({\mathrm{d}}q) and qTXq^{*}TX, but there is a tautological splitting of p1TNp_{1}^{*}TN, given by the identifications

p1TNp1ker(dq)𝒞,(p1TN)Cn=kerTnqCnp_{1}^{*}TN\cong p_{1}^{*}\ker({\mathrm{d}}q)\oplus\mathcal{C},\quad\left(p_{1}^{*}TN\right)_{C_{n}}=\ker T_{n}q\oplus C_{n}

The subbundle 𝒞p1TN\mathcal{C}\subset p_{1}^{*}TN is called the Cartan distribution. We recall that its relevance arises from the fact that it detects which sections τ\tau of q1:J1qXq_{1}\colon J^{1}q\to X are holonomic, i.e., of the form τ=j1σ\tau=j^{1}\sigma, with σ:XN\sigma:X\to N section of q:NXq\colon N\to X. In fact, one has:

  • A local section τ:XJ1q\tau\colon X\to J^{1}q is holonomic if and only if it is tangent to the Cartan distribution, i.e., if

    imdx(p1τ)𝒞τ(x) for all xdomτ.\operatorname{im}{\mathrm{d}}_{x}(p_{1}\circ\tau)\subset\mathcal{C}_{\tau(x)}\mbox{ for all $x\in\operatorname{dom}\tau$}.

The notion of a relative connection formalizes this type of behavior found in the previous example.

Definition 5.2.

Let M𝑝N𝑞XM\xrightarrow{p}N\xrightarrow{q}X be two submersions. A connection on qq relative to pp is a vector bundle CpTNC\subset p^{*}TN complementary to pker(dq)p^{*}\ker({\mathrm{d}}q):

pTNCpker(dq).p^{*}TN\cong C\oplus p^{*}\ker({\mathrm{d}}q).

Equivalently, it is a map c:MJ1qc\colon M\to J^{1}q such that p1c=pp_{1}\circ c=p, where p1:J1qNp_{1}\colon J^{1}q\to N is the projection.

A relative connection CC on M𝑝N𝑞XM\xrightarrow{p}N\xrightarrow{q}X gives rise to an algebroid (A,p,D)(A,p,{\mathrm{D}}) relative to pp. For the vector bundle, we take A=qTXA=q^{*}TX and, under the identification pACp^{*}A\xrightarrow{\sim}C, the anchor map corresponds to the inclusion

ρ:pACpTN.\rho\colon p^{*}A\xrightarrow{\sim}C\subset p^{*}TN.

If X1,X2Γ(TX)X_{1},X_{2}\in\Gamma(TX) are vector fields on XX, then the relative bracket is determined by

[qX1,qX2]D=pq[X1,X2],[q^{*}X_{1},q^{*}X_{2}]_{{\mathrm{D}}}=p^{*}q^{*}[X_{1},X_{2}],

and extended to any sections of AA through the Leibniz rule using the anchor. The resulting relative algebroid is an example of a relative algebroid with injective anchor.

To identify the derivation D{\mathrm{D}} of this relative algebroid one proceeds as follows. Recall from Section 1.5 that there is a canonical map q:𝒟qTX1q𝒟TX1q_{*}\colon\mathcal{D}^{1}_{q^{*}TX}\to q^{*}\mathcal{D}^{1}_{TX}, which pulls back to a map

pq:p𝒟qTX1pq𝒟TX1p^{*}q_{*}\colon p^{*}\mathcal{D}^{1}_{q^{*}TX}\to p^{*}q^{*}\mathcal{D}^{1}_{TX}

The derivation DD is a lift of the de Rham differential d{\mathrm{d}} in the sense that one has pq(D)=pqdp^{*}q_{*}(D)=p^{*}q^{*}{\mathrm{d}}, where d{\mathrm{d}} is interpreted as a section dΓ(𝒟TX1){\mathrm{d}}\in\Gamma(\mathcal{D}^{1}_{TX}).

p𝒟qTX1\textstyle{p^{*}\mathcal{D}^{1}_{q^{*}TX}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pq\scriptstyle{p^{*}q_{*}}pq𝒟TX1\textstyle{p^{*}q^{*}\mathcal{D}^{1}_{TX}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M\textstyle{M\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D\scriptstyle{{\mathrm{D}}}pqd\scriptstyle{p^{*}q^{*}{\mathrm{d}}}

This is the defining feature of the relative derivation associated to a relative connection.

Proposition 5.3.

Let M𝑝N𝑞XM\xrightarrow{p}N\xrightarrow{q}X be two submersions. There is a one-to-one correspondence between pp-relative connections CC on qq and sections DΓ(p𝒟qTX1)D\in\Gamma(p^{*}\mathcal{D}^{1}_{q^{*}TX}) with pqD=pqdp^{*}q_{*}D=p^{*}q^{*}{\mathrm{d}}. In particular, given CC the corresponding derivation DC{\mathrm{D}}_{C} is determined by

{DCf=(pdf)|C,if fC(N),DC(qα)=pqdα,if αΩ(X).\begin{cases}{\mathrm{D}}_{C}f=(p^{*}{\mathrm{d}}f)|_{C},&\text{if }f\in C^{\infty}(N),\\ {\mathrm{D}}_{C}(q^{*}\alpha)=p^{*}q^{*}{\mathrm{d}}\alpha,&\text{if }\alpha\in\Omega^{\bullet}(X).\end{cases}
Proof.

Let DΓ(p𝒟qTX1){\mathrm{D}}\in\Gamma(p^{*}\mathcal{D}^{1}_{q^{*}TX}). The equation

(pqD)(qf)=pqd(qf),(fC(X)),(p^{*}q_{*}{\mathrm{D}})(q^{*}f)=p^{*}q^{*}{\mathrm{d}}(q^{*}f),\quad(f\in C^{\infty}(X)),

holds if and only if for any vector field XΓ(TX)X\in\Gamma(TX) one has

(pqD)(qf)(pqX)=(pqd(qf))(pqX)(p^{*}q_{*}{\mathrm{D}})(q^{*}f)(p^{*}q^{*}X)=(p^{*}q^{*}{\mathrm{d}}(q^{*}f))(p^{*}q^{*}X)

and this is equivalent to:

ρD(pqX)(pqf)=pqX(f).\rho_{{\mathrm{D}}}(p^{*}q^{*}X)(p^{*}q^{*}f)=p^{*}q^{*}X(f).

This last equation holds if and only if ρD:pApTN\rho_{{\mathrm{D}}}\colon p^{*}A\to p^{*}TN is injective with image a subbundle CC complementary to pker(dq)p^{*}\ker({\mathrm{d}}q).

On the other hand, the equation

(pqD)(qα)=pqd(qα),(αΩ1(X)),(p^{*}q_{*}{\mathrm{D}})(q^{*}\alpha)=p^{*}q^{*}{\mathrm{d}}(q^{*}\alpha),\quad(\alpha\in\Omega^{1}(X)),

holds if and only if any vector fields X1,X2Γ(TX)X_{1},X_{2}\in\Gamma(TX) one has

(pqD)(qα)(pqX1,pqX2)=(pqd(qα))(pqX1,pqX2),(p^{*}q_{*}{\mathrm{D}})(q^{*}\alpha)(p^{*}q^{*}X_{1},p^{*}q^{*}X_{2})=(p^{*}q^{*}{\mathrm{d}}(q^{*}\alpha))(p^{*}q^{*}X_{1},p^{*}q^{*}X_{2}),

and in terms of the bracket of D{\mathrm{D}} this is equivalent to

[qX1,qX2]D=pq[X1,X2],[q^{*}X_{1},q^{*}X_{2}]_{{\mathrm{D}}}=p^{*}q^{*}[X_{1},X_{2}],

so the result follows. ∎

The Cartan distribution 𝒞p1TN\mathcal{C}\subset p_{1}^{*}TN in Example 5.1, being a relative distribution for the submersions J1qp1N𝑞XJ^{1}q\xrightarrow{p_{1}}N\xrightarrow{q}X, has an associated relative algebroid with derivation

D𝒞:Ω(qTX)Ω+1(q1TX).{\mathrm{D}}_{\mathcal{C}}\colon\Omega^{\bullet}(q^{*}TX)\to\Omega^{\bullet+1}(q^{*}_{1}TX).

where q1:=qp1:J1qXq_{1}:=q\circ p_{1}\colon J^{1}q\to X. Notice that, by the previous proposition, one has

{D𝒞f=(p1df)|𝒞,if fC(N),D𝒞(qα)=q1dα,if αΩ(X).\begin{cases}{\mathrm{D}}_{\mathcal{C}}f=(p_{1}^{*}{\mathrm{d}}f)|_{\mathcal{C}},&\text{if }f\in C^{\infty}(N),\\ {\mathrm{D}}_{\mathcal{C}}(q^{*}\alpha)=q_{1}^{*}{\mathrm{d}}\alpha,&\text{if }\alpha\in\Omega^{\bullet}(X).\end{cases} (5.1)
Definition 5.4.

The derivation D𝒞D_{\mathcal{C}} is called the Cartan derivation.

The previous proposition leads to the following description of D𝒞{\mathrm{D}}_{\mathcal{C}}.

Corollary 5.5.

Let q:NXq\colon N\to X be a submersion. The bundle of first jets is naturally isomorphic to

J1qq1(d):={Dn𝒟qTX1|q(Dn)=dq(n)},J^{1}q\cong q_{*}^{-1}({\mathrm{d}}):=\left\{{\mathrm{D}}_{n}\in\mathcal{D}^{1}_{q^{*}TX}\ |\ q_{*}({\mathrm{D}}_{n})={\mathrm{d}}_{q(n)}\right\},

where the natural isomorphism is given by restriction of the symbol map σ:𝒟qTX1Hom(qTX,TN)\sigma\colon\mathcal{D}^{1}_{q^{*}TX}\to\operatorname{Hom}(q^{*}TX,TN). The relative algebroid (qTX,p1,D𝒞)(q^{*}TX,p_{1},{\mathrm{D}}_{\mathcal{C}}) has classifying map the resulting inclusion

cD𝒞:J1q𝒟qTX1.\lx@xy@svg{\hbox{\raise 0.0pt\hbox{\kern 19.13422pt\hbox{\ignorespaces\ignorespaces\ignorespaces\hbox{\vtop{\kern 0.0pt\offinterlineskip\halign{\entry@#!@&&\entry@@#!@\cr&\crcr}}}\ignorespaces{\hbox{\kern-19.13422pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{c_{{\mathrm{D}}_{\mathcal{C}}}\colon J^{1}q\,\ignorespaces\ignorespaces\ignorespaces\ignorespaces}$}}}}}}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces{\hbox{\kern 19.13422pt\raise-2.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\lx@xy@hook{1}}}}}}{\hbox{\lx@xy@droprule}}\ignorespaces{\hbox{\kern 43.13422pt\raise-2.0pt\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 43.13422pt\raise 0.0pt\hbox{\hbox{\kern 0.0pt\raise 0.0pt\hbox{\hbox{\kern 3.0pt\raise 0.0pt\hbox{$\textstyle{\mathcal{D}^{1}_{q^{*}TX}}$}}}}}}}\ignorespaces}}}}\ignorespaces.
Proof.

The previous result applied to p=idp=\operatorname{id} shows that for a fixed nNn\in N, we have

{Dn(𝒟qTX1)n|q(Dn)=dq(n)}={CnTnN:Cnker(dnq)=TnN}.\left\{{\mathrm{D}}_{n}\in(\mathcal{D}^{1}_{q^{*}TX})_{n}\ |\ q_{*}(D_{n})={\mathrm{d}}_{q(n)}\right\}=\{C_{n}\subset T_{n}N:C_{n}\oplus\ker({\mathrm{d}}_{n}q)=T_{n}N\}.

The left side is canonically isomorphic to (J1q)n(J^{1}q)_{n}. ∎

Example 5.6 (Cartan derivation in local coordinates).

Let us assume that we have fixed local charts (V,xi)(V,x^{i}) for XX and (U=q1(V),xi,ua)(U=q^{-1}(V),x^{i},u^{a}) for NN, so that

q:NX,q(xi,ua)=xi.q\colon N\to X,\quad q(x^{i},u^{a})=x^{i}.

Then we have an induced chart (p11(U),xi,ua,uia)(p_{1}^{-1}(U),x^{i},u^{a},u^{a}_{i}) on the total space of the first jet bundle so that

p1:J1qN,p1(xi,ua,uia)=(xi,ua).p_{1}\colon J^{1}q\to N,\quad p_{1}(x^{i},u^{a},u^{a}_{i})=(x^{i},u^{a}).

Also, let ei=xie_{i}=\partial_{x^{i}} be the corresponding local frame for TXTX with dual coframe θi=dxi\theta^{i}={\mathrm{d}}x^{i}. Then a form αΩk(qTV)\alpha\in\Omega^{k}(q^{*}TV) can be expressed as

α=i1<<ikαi1,,ik(x,u)dxi1dxik,\alpha=\sum_{i_{1}<\cdots<i_{k}}\alpha_{i_{1},\dots,i_{k}}(x,u)\,{\mathrm{d}}x^{i_{1}}\wedge\cdots\wedge{\mathrm{d}}x^{i_{k}},

and it follows from (5.1) that the Cartan derivation acts on such a form as the total exterior derivative

D𝒞α=ji1<<ik(Djαi1,,ik)dxjdxi1dxik,{\mathrm{D}}_{\mathcal{C}}\alpha=\sum_{j}\sum_{i_{1}<\cdots<i_{k}}(D_{j}\alpha_{i_{1},\dots,i_{k}})\,{\mathrm{d}}x^{j}\wedge{\mathrm{d}}x^{i_{1}}\wedge\cdots\wedge{\mathrm{d}}x^{i_{k}},

where

Djf:=fxj+afuauja.D_{j}f:=\frac{\partial f}{\partial x^{j}}+\sum_{a}\frac{\partial f}{\partial u^{a}}u^{a}_{j}.

One obtains higher order Cartan derivations by considering the higher order jet spaces of the submersion q:NXq\colon N\to X. For any integer k1k\geq 1, one constructs a connection on Jk1qXJ^{k-1}q\to X relative to pk:JkqJk1qp_{k}\colon J^{k}q\to J^{k-1}q by considering first the inclusion JkqJ1(Jk1q)J^{k}q\hookrightarrow J^{1}(J^{k-1}q) and then restricting the Cartan distrbution 𝒞pkTJk1qJ1(Jk1q)\mathcal{C}\subset p_{k}^{*}TJ^{k-1}q\to J^{1}(J^{k-1}q) to JkqJ^{k}q. The resulting relative connection will also be called the (higher order) Cartan distribution.

The higher order Cartan distibution being a relative distribution for the submersions JkqpkJk1qqk1XJ^{k}q\xrightarrow{p_{k}}J^{k-1}q\xrightarrow{q_{k-1}}X, has an associated relative algebroid with derivation

D𝒞:Ω(qk1TX)Ω+1(qkTX).{\mathrm{D}}_{\mathcal{C}}\colon\Omega^{\bullet}(q^{*}_{k-1}TX)\to\Omega^{\bullet+1}(q^{*}_{k}TX).

We also call D𝒞{\mathrm{D}}_{\mathcal{C}} the (higher order) Cartan derivation. Again, applying Proposition 5.3, one has

{D𝒞f=(pkdf)|𝒞,if fC(Jk1q),D𝒞(qk1α)=qkdα,if αΩ(X).\begin{cases}{\mathrm{D}}_{\mathcal{C}}f=(p_{k}^{*}{\mathrm{d}}f)|_{\mathcal{C}},&\text{if }f\in C^{\infty}(J^{k-1}q),\\ {\mathrm{D}}_{\mathcal{C}}(q_{k-1}^{*}\alpha)=q_{k}^{*}{\mathrm{d}}\alpha,&\text{if }\alpha\in\Omega^{\bullet}(X).\end{cases} (5.2)
Example 5.7 (Higher order Cartan derivation in local coordinates).

Similar to the case of first order, we can describe the higher order Cartan derivation in local coordinates as follows. We let d=dimXd=\dim X (number of dependent variables) and n=dimNdimXn=\dim N-\dim X (number of independent variables) and we fix local charts (V,xi)(V,x^{i}) for XX and (U=qk11(V),xi,ua,uJa)(U=q_{k-1}^{-1}(V),x^{i},u^{a},u^{a}_{J}) for Jk1qJ^{k-1}q, with J=(j1,,jr)J=(j_{1},\dots,j_{r}) all unordered rr-tuples of integers with 1jln1\leq j_{l}\leq n, and #J=rk1\#J=r\leq k-1. Then a form αΩk(qk1TV)\alpha\in\Omega^{k}(q_{k-1}^{*}TV) can be expressed as

α=i1<<ikαi1,,ik(x,u,uJ)dxi1dxik,\alpha=\sum_{i_{1}<\cdots<i_{k}}\alpha_{i_{1},\dots,i_{k}}(x,u,u_{J})\,{\mathrm{d}}x^{i_{1}}\wedge\cdots\wedge{\mathrm{d}}x^{i_{k}},

and the Cartan derivation (5.2) acts on such a form as the total exterior derivative

D𝒞α=ji1<<ik(Djαi1,,ik)dxjdxi1dxik,{\mathrm{D}}_{\mathcal{C}}\alpha=\sum_{j}\sum_{i_{1}<\cdots<i_{k}}(D_{j}\alpha_{i_{1},\dots,i_{k}})\,{\mathrm{d}}x^{j}\wedge{\mathrm{d}}x^{i_{1}}\wedge\cdots\wedge{\mathrm{d}}x^{i_{k}},

where now

Djf:=fxj+a=1n#J<k#JfuJuj,Ja.D_{j}f:=\frac{\partial f}{\partial x^{j}}+\sum_{a=1}^{n}\sum_{\#J<k}\frac{\partial^{\#J}f}{\partial u^{J}}u^{a}_{j,J}.

5.2. The relative algebroid of a PDE

We now wish to associate a relative algebroid to a PDE. By the latter we mean:

Definition 5.8.

A partial differential equation (PDE) of order kk on a submersion q:NXq\colon N\to X is a submanifold EJkqE\subseteq J^{k}q. A solution to EE is a (local) section σ\sigma of q:NXq\colon N\to X such that imjkσE\operatorname{im}j^{k}\sigma\subset E.

If we will assume that the image pk(E)Jk1qp_{k}(E)\subseteq J^{k-1}q is a manifold and that the map qk=qk1pk:EXq_{k}=q_{k-1}\circ p_{k}\colon E\to X is a submersion, then the Cartan distribution restricts to a connection of qk1:pk(E)Xq_{k-1}\colon p_{k}(E)\to X relative to the submersions Epkpk(E)qk1XE\xrightarrow{p_{k}}p_{k}(E)\xrightarrow{q_{k-1}}X. The corresponding relative algebroid of the PDE is an algebroid (qk1TX,pk,DE)(q_{k-1}^{*}TX,p_{k},{\mathrm{D}}_{E}) relative to the submersion pk:Epk(E)p_{k}\colon E\to p_{k}(E), and the derivation DE{\mathrm{D}}_{E} is determined by

{DEf=(pkdf)|𝒞,if fC(pk(E)),DE(qk1α)=iEqkdα,if αΩ(X).\begin{cases}{\mathrm{D}}_{E}f=(p_{k}^{*}{\mathrm{d}}f)|_{\mathcal{C}},&\text{if }f\in C^{\infty}(p_{k}(E)),\\ {\mathrm{D}}_{E}(q_{k-1}^{*}\alpha)=i^{*}_{E}q_{k}^{*}{\mathrm{d}}\alpha,&\text{if }\alpha\in\Omega^{\bullet}(X).\end{cases} (5.3)

where iE:EJkqi_{E}\colon E\hookrightarrow J^{k}q is the inclusion.

Definition 5.9.

We call (qk1TX,pk,DE)(q_{k-1}^{*}TX,p_{k},D_{E}) the relative algebroid of the PDE EJkqE\subset J^{k}q.

Remark 5.10.

If we make the weaker assumption that the PDE EJkqE\subseteq J^{k}q intersects the fibers of pk:JkqJk1qp_{k}\colon J^{k}q\to J^{k-1}q in submanifolds of a fixed dimension. Then, by Proposition 4.12, the relative algebroid (qk1TX,pk,D𝒞)(q_{k-1}^{*}TX,p_{k},{\mathrm{D}}_{\mathcal{C}}) associated to the Cartan distribution can be restricted to EE. Hence, we still have a relative algebroid (qk1TX,¯,DE)(q_{k-1}^{*}TX,\overline{\nabla},D_{E}) associated to EE. The results that follow are valid in this more general setting, replacing (qk1TX,pk,DE)(q_{k-1}^{*}TX,p_{k},D_{E}) by this algebroid relative to a foliation. To simplify the exposition, we choose to stay within the framework of algebroids relative to a submersion.

Our next result shows that the relative algebroid of the PDE encodes its solutions.

Theorem 5.11.

Let EJkqE\subset J^{k}q be a PDE with relative algebroid (qk1TX,pk,DE)(q_{k-1}^{*}TX,p_{k},D_{E}). Then germs of solutions to EE are in 1-1 correspondence with germs of realizations of (qk1TX,pk,DE)(q_{k-1}^{*}TX,p_{k},D_{E}) modulo diffeomorphisms.

Proof.

Let σ:UN\sigma\colon U\to N be a local solution of EE. The we construct a realization (θ,r):TPqkTX(\theta,r)\colon TP\to q_{k}^{*}TX of (qk1TX,pk,DE)(q_{k-1}^{*}TX,p_{k},D_{E}) by setting:

P:=U,r(x):=jxqσ,θ(vx)=(r(x),vx).P:=U,\quad r(x):=j^{q}_{x}\sigma,\quad\theta(v_{x})=(r(x),v_{x}).

That θ\theta preserves anchors is clear. Using (5.3) one finds that for any αΩ1(X)\alpha\in\Omega^{1}(X)

θDE(qk1α)=θiEqkdα=dα=dθα.\theta^{*}{\mathrm{D}}_{E}(q^{*}_{k-1}\alpha)=\theta^{*}i^{*}_{E}q_{k}^{*}{\mathrm{d}}\alpha={\mathrm{d}}\alpha={\mathrm{d}}\theta^{*}\alpha.

Conversely, assume that (P,r,θ)(P,r,\theta) is a realization around pPp\in P of (qk1TX,pk,DE)(q_{k-1}^{*}TX,p_{k},D_{E}) such that r(p)=eEr(p)=e\in E. Then the map qkrq_{k}\circ r is a local diffeomorphism, so in a neighborhood of a pPp\in P it factors through a (local) section τ:XE\tau\colon X\to E.

P\textstyle{P\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}r\scriptstyle{r}E\textstyle{E\ignorespaces\ignorespaces\ignorespaces\ignorespaces}qk\scriptstyle{q_{k}}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}σ\scriptstyle{\sigma}

The compatibility of (r,θ)(r,\theta) with the anchor, gives

imdx(pkτ)𝒞τ(x) for all xdomτ.\operatorname{im}{\mathrm{d}}_{x}(p_{k}\circ\tau)\subset\mathcal{C}_{\tau(x)}\mbox{ for all $x\in\operatorname{dom}\tau$}.

So τ\tau is tangent to the Cartan distribuition, and we can conclude that it is holonomic. Hence, τ=jkσ\tau=j^{k}\sigma for a local section σ:XN\sigma\colon X\to N, which is the desired local solution of EE.

5.3. Prolongation and integrability of PDEs

We will now show that the formal theory of prolongations [20] for PDEs coincides with the prolongation theory for the associated relative algebroids.

Theorem 5.12.

Let EJkqE\subset J^{k}q be a PDE with relative algebroid (qk1TX,pk,DE)(q_{k-1}^{*}TX,p_{k},D_{E}). Then:

  1. (i)

    EE is a 1-integrable PDE if and only if the relative algebroid (qk1TX,pk,DE)(q_{k-1}^{*}TX,p_{k},D_{E}) is 1-integrable;

  2. (ii)

    If EE is a 1-integrable PDE, then the relative algebroid corresponding to the prolongation E(1)Jk+1qE^{(1)}\subset J^{k+1}q is the prolongation of (qk1TX,pk,DE)(q_{k-1}^{*}TX,p_{k},D_{E}).

In particular, a PDE is formally integrable if and only if its associated relative algebroid is.

Remark 5.13 (Variational bicomplex).

When E=J1πE=J^{1}\pi Theorem 5.12 shows that the prolongations of EE correspond to the higher order Cartan derivations. From their expression in local coordinates (see Example 5.7), one sees that they assemble together into the horizontal differential of the first row the variational bicomplex (see, e.g., [5]). In other words, the profinite Lie algebroid corresponding to the prolongation tower has as derivation the horizontal differential of the variational bicomplex:

(πTX,dH):\textstyle{{\left(\pi_{\infty}^{*}TX,{\mathrm{d}}_{H}\right)\colon\ldots}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πkTX\textstyle{\pi_{k}^{*}TX\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πk1TX\textstyle{\pi_{k-1}TX\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D𝒞\scriptstyle{{\mathrm{D}}_{\mathcal{C}}}\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π1TX\textstyle{\pi_{1}^{*}TX\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πTX\textstyle{\pi^{*}TX\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D𝒞\scriptstyle{{\mathrm{D}}_{\mathcal{C}}}Jπ:\textstyle{J^{\infty}\pi\colon\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Jkπ\textstyle{J^{k}\pi\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pk\scriptstyle{p_{k}}Jk1π\textstyle{J^{k-1}\pi\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}J1π\textstyle{J^{1}\pi\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p\scriptstyle{p}N.\textstyle{N.}
Proof of Theorem 5.12.

Let us consider first the case E=JkqE=J^{k}q, so DE=D𝒞{\mathrm{D}}_{E}={\mathrm{D}}_{\mathcal{C}}. On the one hand, EE as a PDE, has first prolongation

E(1)=Jk+1q{jy1τ:τΓ(Jkq) holonomic}J1(Jkq).E^{(1)}=J^{k+1}q\cong\left\{j^{1}_{y}\tau\colon\tau\in\Gamma(J^{k}q)\text{ holonomic}\right\}\subset J^{1}(J^{k}q).

On the other hand, by Corollary 5.5, we have the following description of the first jet bundle of JkqJ^{k}q:

J1(Jkq)={Dy𝒟qkTX1:(qk)Dy=dqk(y)}.J^{1}(J^{k}q)=\left\{{\mathrm{D}}_{y}\in\mathcal{D}^{1}_{q_{k}^{*}TX}\colon(q_{k})_{*}{\mathrm{D}}_{y}={\mathrm{d}}_{q_{k}(y)}\right\}.

This gives a description of E(1)E^{(1)} in terms of derivations as

E(1)\displaystyle E^{(1)}\cong {Dy𝒟qkTX1:(qk)Dy=dqk(y),DyD𝒞=0}\displaystyle\left\{{\mathrm{D}}_{y}\in\mathcal{D}^{1}_{q_{k}^{*}TX}\colon(q_{k})_{*}{\mathrm{D}}_{y}={\mathrm{d}}_{q_{k}(y)},\ {\mathrm{D}}_{y}\circ{\mathrm{D}}_{\mathcal{C}}=0\right\}
={Dy𝒟qkTX1:(qk)Dy=(D𝒞)pk(y),DyD𝒞=0}\displaystyle=\left\{{\mathrm{D}}_{y}\in\mathcal{D}^{1}_{q_{k}^{*}TX}\colon(q_{k})_{*}{\mathrm{D}}_{y}=({\mathrm{D}}_{\mathcal{C}})_{p_{k}(y)},\ {\mathrm{D}}_{y}\circ{\mathrm{D}}_{\mathcal{C}}=0\right\}

which is precisely the first prolongation space of the relative algebroid (qk1TX,pk,D𝒞)(q_{k-1}^{*}TX,p_{k},D_{\mathcal{C}}). It follows that the theorem holds in this case.

For general EJkqE\subset J^{k}q, the result follows because the first prolongation of EE can be described as E(1)=J1EJk+1qE^{(1)}=J^{1}E\cap J^{k+1}q in J1(Jkq)J^{1}(J^{k}q), which corresponds to the first prolongation of the relative algebroid using the description for Jk+1qJ^{k+1}q in terms of derivations. This proves both (i) and (ii). ∎

Example 5.14.

We illustrate the theorem with the simple PDE ux=yu_{x}=y, where u=u(x,y)u=u(x,y). As a manifold, this PDE has coordinates (uy,u,x,y)(u_{y},u,x,y) and sits inside

{(x,uy,u,x,y)}J1π={(ux,uy,u,x,y)},\{(x,u_{y},u,x,y)\}\subset J^{1}\pi=\{(u_{x},u_{y},u,x,y)\},

where π:32\pi\colon\mathbb{R}^{3}\to\mathbb{R}^{2} is the projection π(u,x,y)=(x,y)\pi(u,x,y)=(x,y). The corresponding relative algebroid is obtained by restricting the Cartan derivation to EE, so it can be described by the trivial vector bundle 2¯3\underline{\mathbb{R}^{2}}\to\mathbb{R}^{3}, with derivation D{\mathrm{D}} determined by

{Dθ1=Dθ2=0,Dx=θ1,Dy=θ2,Du=yθ1+uyθ2.\begin{cases}{\mathrm{D}}\theta^{1}={\mathrm{D}}\theta^{2}=0,\\ {\mathrm{D}}x=\theta^{1},\\ {\mathrm{D}}y=\theta^{2},\\ {\mathrm{D}}u=y\theta^{1}+u_{y}\theta^{2}.\end{cases}

The free variable is uyu_{y}, so to compute the prolongation we start with an extension of D{\mathrm{D}}, which is determined by D~uy=uxyθ1+uyyθ2\tilde{{\mathrm{D}}}u_{y}=u_{xy}\theta^{1}+u_{yy}\theta^{2}. We find that D~\tilde{{\mathrm{D}}} must satisfy

0=D~Du=θ2θ1+(uxyθ1+uyyθ2)θ1=(uxy1)θ1θ2.0=\tilde{{\mathrm{D}}}{\mathrm{D}}u=\theta^{2}\wedge\theta^{1}+(u_{xy}\theta^{1}+u_{y}y\theta^{2})\wedge\theta^{1}=(u_{xy}-1)\theta^{1}\wedge\theta^{2}.

We find that the 1st prolongation of the relative algebroid is characterized by uxy=1u_{xy}=1 and that uyyu_{yy} is the new variable. This is corresponds exactly to the relative algebroid underlying the prolongation E(1)J2πE^{(1)}\subset J^{2}\pi of the PDE: the latter is given by the equations {ux=y,uxx=0,uxy=1}\{u_{x}=y,u_{xx}=0,u_{xy}=1\}, so E(1)E^{(1)} is parametrized by {(uyy,uy,u,x,y)}\{(u_{yy},u_{y},u,x,y)\}.

Remark 5.15 (PDEs with symmetries).

Symmetries of PDEs are given by pseudogroups of diffeomorphism (see, e.g., [26]). In future work, we will show that the symmetries of a PDE also preserve the underling relative algebroid, so that the structure of the relative algebroid descends to the quotient.

Remark 5.16 (Pfaffian fibrations).

A different framework for PDEs with symmetries comes from Pfaffian fibrations and Pfaffian actions of Pfaffian groupoids [1, 2, 10, 11, 30]. We will explain in future work how relative algebroids underlie Pfaffian fibration, and how Pfaffian actions give rise to symmetries of the underlying relative algebroids.

6. Postlude: an example

In this final section, we will revisit Example 0.2 from the Introduction and we discuss it using the framework developed in the paper. This example, considered by Bryant in [8, §5.1], is simple enough that can be solved directly, but it is extremely insightful to study it from the perspective of relative derivations.

As discussed in the Introduction, the existence and classification problem of surfaces with a metric whose Gauss curvature satisfies |K|=1|\nabla K|=1 is govern by the equations

{dθ1=θ3θ1,dθ2=θ3θ1,dθ3=Kθ1θ2,dK=cos(φ)θ1+sin(φ)θ2.\begin{cases}{\mathrm{d}}\theta^{1}=-\theta^{3}\wedge\theta^{1},\\ {\mathrm{d}}\theta^{2}=\theta^{3}\wedge\theta^{1},\\ {\mathrm{d}}\theta^{3}=K\theta^{1}\wedge\theta^{2},\\ {\mathrm{d}}K=\cos(\varphi)\theta^{1}+\sin(\varphi)\theta^{2}.\end{cases} (6.1)

These equations define a derivation DA{\mathrm{D}}_{A} on the trivial vector bundle A:=3¯A:=\underline{\mathbb{R}^{3}}\to\mathbb{R}, relative to the projection p:SS1×p\colon\SS^{1}\times\mathbb{R}\to\mathbb{R}, where \mathbb{R} has coordinate KK and SS1\SS^{1} has coordinate φ\varphi. Here {θ1,θ2,θ3}\{\theta^{1},\theta^{2},\theta^{3}\} is a basis of sections of AA^{*} and if we let {e1,e2,e3}\{e_{1},e_{2},e_{3}\} be the dual basis of sections of AA, the anchor of this relative algebroid is given by

ρ(e1)\displaystyle\rho(e_{1}) =cos(φ)K,\displaystyle=\cos(\varphi)\partial_{K}, ρ(e2)\displaystyle\rho(e_{2}) =sin(φ)K,\displaystyle=\sin(\varphi)\partial_{K}, ρ(e3)\displaystyle\rho(e_{3}) =0,\displaystyle=0,

while the bracket takes the form

[e1,e2]\displaystyle[e_{1},e_{2}] =Ke3,\displaystyle=-Ke_{3}, [e2,e3]\displaystyle[e_{2},e_{3}] =e1,\displaystyle=-e_{1}, [e3,e1]\displaystyle[e_{3},e_{1}] =e2.\displaystyle=-e_{2}.

Using this expression for the anchor, a straightforward computation shows that this relative algebroid has tableau map τ:ker(dp)pHom(3¯,T)\tau\colon\ker({\mathrm{d}}p)\to p^{*}\operatorname{Hom}(\underline{\mathbb{R}^{3}},T\mathbb{R}) given by

τ(φ)=(sin(φ)θ1+cos(φ)θ2)K.\tau(\partial_{\varphi})=\left(-\sin(\varphi)\theta^{1}+\cos(\varphi)\theta^{2}\right)\otimes\partial_{K}.

In the sequel, it will be convenient to use the following notation for the expression appearing in the tableau:

φ(dK):=sin(φ)θ1+cos(φ)θ2.\partial_{\varphi}({\mathrm{d}}K):=-\sin(\varphi)\theta^{1}+\cos(\varphi)\theta^{2}.

All higher prolongations can be explicitly computed, giving the full prolongation tower

3¯\textstyle{\underline{\mathbb{R}^{3}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}3¯\textstyle{\underline{\mathbb{R}^{3}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}3¯\textstyle{\underline{\mathbb{R}^{3}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}3¯\textstyle{\underline{\mathbb{R}^{3}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}×SS1×\textstyle{\mathbb{R}^{\infty}\times\SS^{1}\times\mathbb{R}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}k×SS1×\textstyle{\mathbb{R}^{k}\times\SS^{1}\times\mathbb{R}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}pk\scriptstyle{p_{k}}k1×SS1×\textstyle{\mathbb{R}^{k-1}\times\SS^{1}\times\mathbb{R}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}p1\scriptstyle{p_{1}},\textstyle{\mathbb{R},}

where k×SS1×\mathbb{R}^{k}\times\SS^{1}\times\mathbb{R} has coordinates (ck,,c1,φ,K)(c_{k},\ldots,c_{1},\varphi,K), and the relative derivation is determined by (6.1) together with

Dck=fk[c1,,ck]dK+ck+1φ(dK){\mathrm{D}}c_{k}=-f_{k}[c_{1},\dots,c_{k}]{\mathrm{d}}K+c_{k+1}\partial_{\varphi}({\mathrm{d}}K) (6.2)

where fk[c1,,ck]f_{k}[c_{1},\dots,c_{k}] are polynomials given by

f1[c1]\displaystyle f_{1}[c_{1}] =c12K\displaystyle=-c_{1}^{2}-K
f2m[c1,,c2m]\displaystyle f_{2m}[c_{1},\dots,c_{2m}] =i=1m(2m+1i)cic2m+1i\displaystyle=-\sum_{i=1}^{m}\binom{2m+1}{i}c_{i}c_{2m+1-i}
f2m+1[c1,,c2m+1]\displaystyle f_{2m+1}[c_{1},\dots,c_{2m+1}] =(2m+1m)cm+12i=1m(2m+2i)cic2m+2i\displaystyle=-\binom{2m+1}{m}c_{m+1}^{2}-\sum_{i=1}^{m}\binom{2m+2}{i}c_{i}c_{2m+2-i}

for m1m\geq 1. Note that we can also interpret (6.2) as defining a profinite derivation on 3¯×SS1×\underline{\mathbb{R}^{3}}\to\mathbb{R}^{\infty}\times\SS^{1}\times\mathbb{R}

Equation (6.2) suggests using the new global coframe {dK,φ(dK),θ3}\{{\mathrm{d}}K,\partial_{\varphi}({\mathrm{d}}K),\theta^{3}\}, for which the anchor is decoupled such that the profinite part is concentrated in one basis vector only. The frame {b1,b2,b3}\{b_{1},b_{2},b_{3}\} dual to this coframe is given by

b1=(cosφ)e1+(sinφ)e2b2=(sinφ)e1+(cosφ)e2c1e3b3=e3.\displaystyle\begin{split}b_{1}&=(\cos\varphi)e_{1}+(\sin\varphi)e_{2}\\ b_{2}&=-(\sin\varphi)e_{1}+(\cos\varphi)e_{2}-c_{1}e_{3}\\ b_{3}&=e_{3}.\end{split} (6.3)

The profinite Lie algebroid 3¯×SS1×\underline{\mathbb{R}^{3}}\to\mathbb{R}^{\infty}\times\SS^{1}\times\mathbb{R}, with respect to the new frame, has bracket

[b1,b2]=c1b2,[b1,b3]=[b2,b3]=0.[b_{1},b_{2}]=-c_{1}b_{2},\qquad[b_{1},b_{3}]=[b_{2},b_{3}]=0. (6.4)

and anchor

ρ(b1)\displaystyle\rho_{\infty}(b_{1}) =K+k=1fk[c1,,ck]ck,\displaystyle=\partial_{K}+\sum_{k=1}^{\infty}f_{k}[c_{1},\dots,c_{k}]\partial_{c_{k}}, ρ(b2)\displaystyle\rho_{\infty}(b_{2}) =k=1ck+1ck,\displaystyle=\sum_{k=1}^{\infty}c_{k+1}\partial_{c_{k}}, ρ(b3)\displaystyle\rho_{\infty}(b_{3}) =φ.\displaystyle=\partial_{\varphi}. (6.5)

The algebroid decouples as the product of TSS1SS1T\SS^{1}\to\SS^{1} and 2¯×\underline{\mathbb{R}^{2}}\to\mathbb{R}^{\infty}\times\mathbb{R}, where the latter has global frame {b1,b2}\{b_{1},b_{2}\}. It will be convenient to set Σk:=k×\Sigma^{k}:=\mathbb{R}^{k}\times\mathbb{R}, and denote by

A:=2¯ΣA:=\underline{\mathbb{R}^{2}}\to\Sigma^{\infty}

the algebroid with global frame {b1,b2}\{b_{1},b_{2}\}.

The vector field ρ(b1)\rho_{\infty}(b_{1}) is levelwise profinite in nature, and its flow can be explicitly computed in terms of the solution of a Riccati equation. However, by [13, Lemma 3.3], the flow of ρ(b2)\rho_{\infty}(b_{2}) has no flow defined on profinite open subsets. For this reason, there can not be a smooth groupoid, whose source fibers are 2-dimensional manifolds; if there was one, the right-invariant vector field corresponding to b2b_{2} would have a flow restricted to each source fiber, which would descend to a flow of ρ(b2)\rho_{\infty}(b_{2}) on the base Σ\Sigma^{\infty}.

The vector field ρ(b2)\rho_{\infty}(b_{2}) does have a flow on a different space. For this, note that this vector field lives entirely on \mathbb{R}^{\infty}, with coordinates c=(ck)kc_{\infty}=(c_{k})_{k\in\mathbb{N}}, and is completely decoupled from the coordinate KK on Σ\Sigma^{\infty}. So let C0ω()C^{\omega}_{0}(\mathbb{R}) and C0()C^{\infty}_{0}(\mathbb{R}) be the space of germs of analytic, respectively smooth, functions around 0. The jet map j0j^{\infty}_{0} relates the vector field ρ(b2)\rho_{\infty}(b_{2}) to the vector field t\partial_{t} on the spaces of germs.

(C0(),t)\textstyle{{(C^{\infty}_{0}(\mathbb{R}),\partial_{t})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j0\scriptstyle{j_{0}^{\infty}}(C0ω(),t)\textstyle{{(C^{\omega}_{0}(\mathbb{R}),\partial_{t})}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}j0\scriptstyle{j_{0}^{\infty}}(,kck+1ck)\textstyle{{\left(\mathbb{R}^{\infty},\sum_{k}c_{k+1}\partial_{c_{k}}\right)}}

Now observe that:

  • On the space of analytic germs C0ω()C^{\omega}_{0}(\mathbb{R}), the vector field t\partial_{t} has a flow Φt\Phi_{t}, given by

    Φt(germ0(f()))=germ0(f(+t))\Phi_{t}(\operatorname{germ}_{0}(f(\cdot)))=\operatorname{germ}_{0}(f(\cdot+t))

    whenever tt is in the maximal domain to which the germ of ff can be extended;

  • On C0()C^{\infty}_{0}(\mathbb{R}), the vector field t\partial_{t} has no flow, since there are many distinct integral curves through a point, due to the existence of flat functions;

  • On \mathbb{R}^{\infty}, as we already mentioned, the vector field kck+1ck\sum_{k}c_{k+1}\partial_{c_{k}} also has no flow.

It is therefore natural to restrict ρ(f2)\rho_{\infty}(f_{2}) to the space of convergent power series:

ω:={c:lim supk(|ck|k!)1/k<}.\mathbb{R}^{\omega}:=\left\{c_{\infty}\in\mathbb{R}^{\infty}\colon\limsup_{k}\left(\frac{|c_{k}|}{k!}\right)^{1/k}<\infty\right\}.

This space is in bijection with C0()C^{\infty}_{0}(\mathbb{R}) and it is more natural to equip ω\mathbb{R}^{\omega} with the smooth structure of C0()C^{\infty}_{0}(\mathbb{R}) rather than the profinite smooth structure of \mathbb{R}^{\infty}. In a similar vein, we let

Σω:=ω×Σ,\Sigma^{\omega}:=\mathbb{R}^{\omega}\times\mathbb{R}\subset\Sigma^{\infty},

and consider the restricted algebroid

Aω:=A|ΣωΣω,A^{\omega}:=A|_{\Sigma^{\omega}}\to\Sigma^{\omega},

which makes sense since ρ(f2)\rho_{\infty}(f_{2}) is tangent to Σω\Sigma^{\omega}. The integral manifolds of AωA^{\omega} partition Σω\Sigma^{\omega} into well-defined leaves, and the algebroid can (at least) leafwise be integrated to a smooth groupoid 𝒢ωΣω\mathcal{G}^{\omega}\rightrightarrows\Sigma^{\omega} whose source-fibers are simply connected manifolds.

The original classification problem can be solved on this space: it is governed by the SS1\SS^{1}-structure algebroid TSS1×AωSS1×ΣωT\SS^{1}\times A^{\omega}\to\SS^{1}\times\Sigma^{\omega} whose canonical SS1\SS^{1}-integration is the SS1\SS^{1}-structure groupoid (SS1×SS1)×𝒢ωSS1×Σω(\SS^{1}\times\SS^{1})\times\mathcal{G}^{\omega}\rightrightarrows\SS^{1}\times\Sigma^{\omega}, whose source fibers are of the form SS1×s𝒢ω1(c)\SS^{1}\times s_{\mathcal{G}^{\omega}}^{-1}(c_{\infty}). These are the coframe bundles of non-extendable simply connected solutions to the realization problem! The leaves and the groupoid 𝒢ω\mathcal{G}^{\omega} can be explicitly described, and their isometry groups can be listed. The appropriate smooth structure on the total space of this groupoid has yet to be studied, but we believe its a type of diffeological groupoid that differentiates to the given algebroid, in the sense of Aintablian and Blohmann [4].

Solutions with additional symmetry

Even without delving into the complicated theory of infinite-dimensional geometry, to make sense of the algebroid governing the full realization problem, a glance at the profinite algebroid TSS1×ASS1×ΣT\SS^{1}\times A\to\SS^{1}\times\Sigma^{\infty} can already lead to interesting insights and solutions. For example, it is possible to find solutions by looking at the locus where the anchor drops rank, which amounts to imposing extra symmetry on realizations.

The expression for the anchor shows that the algebroid AΣA\to\Sigma^{\infty} drops rank on the subspace

Σ0={(c,K)Σ:ck=0 for k2}.\Sigma^{\infty}_{0}=\{(c_{\infty},K)\in\Sigma^{\infty}:c_{k}=0\mbox{ for $k\geq 2$}\}.

This locus is finite-dimensional! One obtains a restricted algebroid with anchor and bracket given by

ρ(b1)=K(K+c12)c1,ρ(b2)=0,[b1,b2]=c1b2.\rho(b_{1})=\partial_{K}-(K+c_{1}^{2})\partial_{c_{1}},\qquad\rho(b_{2})=0,\qquad[b_{1},b_{2}]=-c_{1}b_{2}.

This algebroid can be explicitly integrated (in terms of solutions to a Riccati equation) to a finite dimensional Lie groupoid. By the work of Fernandes and Struchiner [18], it represents the stack of complete, simply connected, solutions of the realization problem with translational symmetry. Geometrically, the extra symmetry is translation in the direction orthogonal to K\nabla K.

7. Outlook

In this paper, we have established the foundational framework of the theory. There are numerous directions to further explore. Below, we outline a few key directions that we are currently investigating and its relationships with existing literature.

7.1. PDEs with symmetries

In the context of the formal theory of PDEs, one of the key advantages of the framework of relative algebroids is its stability under quotients by symmetries. In Section 5, we saw that every PDE has an associated canonical relative algebroid whose realizations, up to diffeomorphism, correspond to solutions of the PDE. Classification problems in geometry are often governed by PDEs with large symmetry groups. In future work, we will precisely define symmetries of relative algebroids and their quotients by symmetries. Moreover, one can show that the symmetries of a PDE correspond to the symmetries of its associated relative algebroid. Our ultimate goal is to provide a rigorous explanation of how Bryant’s equations arise from a geometric problem formulated as a PDE.

We expect that the quotient algebroid associated with a PDE with symmetries will be related to various existing approaches to symmetries in the literature. For example, given a Lie pseudogroup Γ\Gamma acting on a differential equation EJkqE\subset J^{k}q, it is known that, under relatively mild assumptions, the space of differential invariants of the PDE is finitely generated and can be computed through established algorithms (see, e.g., [24, 28]). We conjecture that the differential algebra of these invariants coincides with the exterior algebra of the prolongation tower of the quotient relative algebroid. This would provide a Lie-theoretic interpretation of the results in [24, 28]. Establishing this precise connection should not only broaden the theory developed in loc. cit. but also lead to powerful new tools.

On a related note, we saw in Remark 5.13 that the profinite differential of the prolongation tower of a formally integrable PDE EJkqE\subset J^{k}q corresponds to the horizontal differential (at the bottom row) of the variational bicomplex of EE. The prolongation tower of the symmetry quotient of the relative algebroid then should correspond to the horizontal differential (at the bottom row) of the invariant variational bicomplex of the PDE [23].

The tableau of the quotient algebroid associated with a PDE with symmetries is much smaller than the tableau of the PDE itself. Moreover, PDEs with very large symmetry groups (such as the group of diffeomorphisms) can never possess desirable properties like ellipticity or finite type. In certain cases, however, PDEs with symmetries may exhibit these properties within a fixed gauge. This suggests that the tableau of the corresponding quotient may also retain such properties. One of our goals is to investigate whether properties of the tableau of the relative algebroid can lead to existence results.

Another connection to the existing literature arises through Pfaffian fibrations and Pfaffian actions [1, 2, 10, 11, 30]. Just as PDEs give rise to relative algebroids, Pfaffian fibrations induce relative algebroids, and Pfaffian actions give rise to symmetries of these algebroids. We hope to explore this connection further in future work.

7.2. Relative GG-structure algebroids

The Bryant-Cartan existence result for realizations in the analytic setting (Theorem 3.20) provides local manifolds with coframes that solve the realization problem. However, in many realization problems, such as those arising from Riemannian manifolds or more general GG-structures, the realization problem is naturally formulated in terms of coframes on the orthonormal frame bundle or a principal GG-bundle. In general, the local solutions obtained from the Bryant-Cartan theorem do not yield such principal bundles, as it does not account for the presence of a structure group. Therefore, it is desirable to incorporate structure groups into the theory of relative algebroids.

In the case of finite-dimensional Lie algebroids, this has been achieved in [18], where a theory of GG-structure algebroids and GG-structure groupoids is developed. That work also establishes necessary and sufficient conditions for the existence of GG-structure realizations. We aim to extend this theory to relative algebroids to better understand the role of the structure group GG in solving the realization problem for GG-structure relative algebroids. In some preliminary work we have established the existence of a universal (profinite) GG-structure algebroid through which every GG-structure algebroid factors.

7.3. Profinite Lie algebroids

Profinite-dimensional manifolds and bundles appear extensively in the theory of formal PDEs (see, e.g., [1, 3, 21, 27]). Hence, their emergence in the theory of relative algebroids is not surprising: the base space of the prolongation tower of a relative algebroid is the space of formal realizations modulo symmetries.

There are several challenges regarding the existence of smooth groupoids integrating the prolongation tower of a relative algebroid. In fact, we suspect that such groupoids may not exist for any prolongation tower. Intuitively, this stems from the fact that there is no “continuous” way to assign a smooth function to each jet.

However, the prolongation tower of a relative algebroid exhibits much richer geometry when restricted to smaller spaces, such as the space of convergent power series. The algebroid, when restricted to this space, has leaves and can be leafwise integrated into a groupoid. Moreover, in this case, tools from finite-dimensional Lie theory become available to study the space of global solutions to the realization problem. One main goal is to understand the moduli stack of complete solutions, which requires a precise understanding of the smooth structure on the integrating object. One possible approach is through diffeologies; in this context, the recent work of Aintablian and Blohmann [4] on diffeological Lie groupoids and algebroids should be particularly relevant. Another possible approach, suggested to us by Ivan Contreras, is to consider formal Lie groupoids (see, e.g., [12, 14]) integrating a profinite Lie groupoid. Since formal Lie groupoids arise from power series, it seems plausible that integrating objects of such nature for profinite Lie algebroids may exist.

7.4. Applications to and interactions with control theory

Example 3.15 illustrates how a control system arises from a relative algebroid defined by a relative vector field. The interaction between control theory and relative algebroids should also work in the other direction.

As a particular example, the notion of controllability (the equivalence relation of points connected by realizations – see [22]) should also be present for relative algebroids. This notion should induce a partition of the base of a relative algebroid into invariant submanifolds. This bidirectional interaction suggests deeper connections between the two fields. Here are two other relations worth exploring:

In control theory, one typically works with a relative distribution HpTNH\subset p^{*}TN, where p:MNp\colon M\to N is a submersion. There seems to be no canonical way to define a derivation relative to pp on HH itself, without assuming some additional structure. However, there is a well-defined notion of an integral manifold, namely a submanifold LML\subset M such that dp(TL)=H|L{\mathrm{d}}p(TL)=H|_{L}. The precise relationship between relative distributions and relative algebroids remains to be fully understood.

Furthermore, since relative algebroids are particularly well-suited for studying realization problems with symmetry, they could also provide a useful framework for the study of control systems with symmetries.

References

  • [1] L. Accornero. Topics on Lie pseudogroups. Phd thesis, Utrecht University, 2021.
  • [2] L. Accornero, F. Cattafi, M. Crainic, and M. A. Salazar. Pseudogroups and Geometric Structures. To appear, 2025.
  • [3] L. Accornero and M. Crainic. Haefliger’s differentiable cohomology. Journal of Noncommutative Geometry, 2024.
  • [4] L. Aintablian and C. Blohmann. Differentiable groupoid objects and their abstract Lie algebroids. arXiv:2412.19697, 2024.
  • [5] I. M. Anderson. Introduction to the variational bicomplex. Contemp. Math, 132:51–73, 1992.
  • [6] R. L. Bryant. Recent advances in the theory of holonomy. Séminaire Bourbaki, 41:351–374, 1998-1999.
  • [7] R. L. Bryant. Bochner-Kähler metrics. Journal of the AMS, 14(3):623–715, 2001.
  • [8] R. L. Bryant. Notes on exterior differential systems. arXiv:1405.3116, 2014.
  • [9] R. L. Bryant, S.-S. Chern, R. B. Gardner, H. L. Goldschmidt, and P. A. Griffiths. Exterior differential systems. Springer, 1991.
  • [10] F. Cattafi. A general approach to almost structures in geometry. Phd thesis, Utrecht University, 2020.
  • [11] F. Cattafi, M. Crainic, and M. A. Salazar. From PDEs to Pfaffian fibrations. L’Enseignement Mathématique, 66(1):187–250, 2020.
  • [12] A. S. Cattaneo, B. Dherin, and G. Felder. Formal symplectic groupoid. Comm. Math. Phys., 253(3):645–674, 2005.
  • [13] V. N. Chetverikov. On the structure of integrable C-fields. Differential Geometry and its Applications, 1(4):309–325, 1991.
  • [14] I. Contreras. Models for formal groupoids. In Geometric and topological methods for quantum field theory, pages 322–339. Cambridge Univ. Press, Cambridge, 2013.
  • [15] M. Crainic and I. Moerdijk. Deformations of Lie brackets: cohomological aspects. J. Eur. Math. Soc. (JEMS), 10(4):1037–1059, 2008.
  • [16] M. Crainic and O. Yudilevich. Lie pseudogroups à la Cartan. Memoirs of the American Mathematical Society, 303(1527), 2024.
  • [17] R. L. Fernandes and I. Struchiner. The classifying Lie algebroid of a geometric structure I: Classes of coframes. Transactions of the American Mathematical Society, 366(5):2419–2462, 2014.
  • [18] R. L. Fernandes and I. Struchiner. The global solutions to Cartan’s realization problem. To appear in Memoirs of the American Mathematical Society, 2019.
  • [19] R. L. Fernandes and I. Struchiner. The classifying Lie algebroid of a geometric structure II: G-structures with connection. São Paulo Journal of Mathematical Sciences, 15:524–570, 2021.
  • [20] H. Goldschmidt. Integrability criteria for systems of nonlinear partial differential equations. Journal of Differential Geometry, 1(3-4):269–307, 1967.
  • [21] B. Güneysu, M. J. Pflaum, et al. The profinite dimensional manifold structure of formal solution spaces of formally integrable PDEs. Symmetry, Integrability and Geometry: Methods and Applications, 13:003, 2017.
  • [22] R. Hermann and A. Krener. Nonlinear controllability and observability. IEEE Transactions on automatic control, 22(5):728–740, 1977.
  • [23] I. A. Kogan and P. J. Olver. The invariant variational bicomplex. Contemporary Mathematics, 285:131–144, 2001.
  • [24] B. Kruglikov and V. Lychagin. Global Lie–Tresse theorem. Selecta Mathematica, 22:1357–1411, 2016.
  • [25] I. Moerdijk and J. Mrčun. Introduction to foliations and Lie groupoids, volume 91 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2003.
  • [26] P. J. Olver. Applications of Lie groups to differential equations, volume 107 of Graduate Texts in Mathematics. Springer-Verlag, New York, second edition, 1993.
  • [27] P. J. Olver and J. Pohjanpelto. Maurer-Cartan forms and the structure of Lie pseudo-groups. Selecta Math. (N.S.), 11(1):99–126, 2005.
  • [28] P. J. Olver and J. Pohjanpelto. Differential invariant algebras of Lie pseudo-groups. Advances in Mathematics, 222(5):1746–1792, 2009.
  • [29] J. Pierard de Maujouy. Physics and geometry from a Lagrangian: Dirac spinors on a generalised frame bundle. Journal of Geometry and Physics, 206:105297, 2024.
  • [30] M. A. Salazar. Pfaffian groupoids. Phd thesis, Utrecht University, 2013.
  • [31] W. M. Seiler. Spencer cohomology, differential equations, and Pommaret bases. Radon Ser. Comput. Appl. Math, 2:169–216, 2007.
  • [32] O. Yudilevich. Lie pseudogroups à la Cartan from a modern perspective. Phd thesis, Utrecht University, 2016.