09.01.2015 Views

SECTIONS ALONG A MAP APPLIED TO HIGHER-ORDER ...

SECTIONS ALONG A MAP APPLIED TO HIGHER-ORDER ...

SECTIONS ALONG A MAP APPLIED TO HIGHER-ORDER ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>SECTIONS</strong> <strong>ALONG</strong> A <strong>MAP</strong> <strong>APPLIED</strong> <strong>TO</strong> <strong>HIGHER</strong>-<strong>ORDER</strong><br />

LAGRANGIAN MECHANICS. NOETHER’S THEOREM<br />

JOSÉ F. CARIÑENA⋆ , CARLOS LÓPEZ† , AND EDUARDO MARTÍNEZ⋆<br />

Abstract. We show that the concept of section along a map is a fundamental<br />

concept in the framework of the geometrical description of Classical Mechanics.<br />

We review the higher-order Lagrangian Mechanics formulation, and simpler<br />

redefinitions of basic objects appear in a natural way. As an application, a<br />

Noether’s theorem for higher-order Lagrangian Mechanics admitting a converse<br />

is developed.<br />

1. Introduction<br />

Geometric techniques have been applied to Physics in several ways during the<br />

last years, and so the Geometry of the Tangent and Cotangent Bundles has been<br />

shown to play a very relevant role in mechanics of time-independent systems from a<br />

geometric perspective. Geometric concepts as vector fields and forms are now well<br />

established in physics literature and the language of jets has more recently appeared<br />

as related to the study of (partial) differential equations and variational calculus.<br />

The theory of connections in principal bundles as geometric counterpart of gauge<br />

theories has also been very illuminating. We feel however that a generalization of<br />

the concept of vector field and form, or more generally, that of a section of a vector<br />

bundle, we will deal with, is also worthy of attention. It has been very often used<br />

in physics but without any previous geometric interpretation and we hope that a<br />

deeper geometric study of this concept and its applications to physics may be very<br />

fruitful for a better understanding of Classical Mechanics and Field Theory. In fact,<br />

among other things, it is the geometric concept necessary to deal with non-point<br />

transformations in Classical Mechanics.<br />

The concept of vector field along a map, well-known in other fields of Mathematics<br />

[17], has been introduced in order to describe in Lagrangian Mechanics<br />

objects like the time evolution operator [11] or symmetries of a Lagrangian system<br />

[5], [12], [13]. This concept has provided a new and simpler approach to some<br />

problems, like the connection between Lagrangian and Hamiltonian formalisms or<br />

the converse of Noether’s theorem. We aim to show in this paper that all the geometric<br />

framework of Mechanics may be based on this concept, from the very first<br />

definitions of the basic canonical objects, that so appear in a much more natural<br />

way.<br />

1991 Mathematics Subject Classification. 58F05.<br />

Key words and phrases. Higher-order tangent bundles,Higher-order Lagrangian Mechanics,<br />

Sections along a map, Derivations.<br />

Eduardo Martínez thanks Ministerio de Educación y Ciencia for a grant. Partial financial<br />

support by CICYT is also acknowledged.<br />

1


2 JOSÉ F. CARIÑENA⋆ , CARLOS LÓPEZ† , AND EDUARDO MARTÍNEZ⋆<br />

We propose an extensive use of the concept of section along a map in the study<br />

of the different problems of Mechanics, not only because of the simplification so<br />

obtained, but because of the necessity of this concept in order to enlarge the field<br />

of available tools too.<br />

Vector fields along a map f : N → M act on functions on M giving rise to<br />

functions on N in much the same way as vector fields act on functions on a manifold.<br />

On the other hand, a theory of derivations of forms along a map, similar to the<br />

usual Frölicher and Nijenhuis’ theory, has recently been developed by Pidello and<br />

Tulczyjew [16]. We will make an extensive use of this theory.<br />

Along the paper we shall review from this new perspective basic objects like the<br />

Poincaré form, total time derivative, complete lifts, etc. Once a Lagrangian function<br />

has been introduced, objects associated to L, like the Poincaré-Cartan form, the<br />

time evolution operator, symmetries and other related ones, are again seen from<br />

this new perspective. In particular it is explicitely shown how the Cartan form and<br />

the Legendre transformation are but two different features of the same object.<br />

The paper is organized as follows: in section 2 we introduce the basic notation,<br />

while in section 3 we give some properties of sections along a map and explain how<br />

some of these objects appear in a canonical way. In section 4 it is shown how a<br />

vector field along a higher-order tangent bundle projection may be prolonged, and<br />

then we study its properties. In section 5 we study the canonical vector field along<br />

the (first) tangent bundle projection, T (0) , its prolongations and relations with<br />

those of other vector fields. In section 6 some properties of the variational operator<br />

and related operators are studied. Higher-order differential equations are studied<br />

in section 7. Finally, in section 8 we show the application of all the preceding<br />

concepts to the study of Lagrangian Mechanics, and in particular, we give a new<br />

version of Noether’s theorem showing that there exists a one-to-one correspondence<br />

between the symmetries of the Lagrangian and those of the Hamiltonian dynamical<br />

system defined by this Lagrangian, and, in consequence, between symmetries of the<br />

Lagrangian and first integrals of the dynamics defined by it. We also point out the<br />

difficulties that can be found when trying to generalize the theory developed by<br />

Marmo and Mukunda for first order Lagrangians [14].<br />

2. Notation<br />

The basic notation of this paper is mainly that of [8]. See also [20] and [9] for a<br />

more detailed introduction to general concepts and definitions.<br />

Capital letters will refer to smooth manifolds and the set of smooth functions<br />

on a manifold M will be denoted by C ∞ (M). Given a manifold M, C ∞ (R, M)<br />

will be the set of differentiable curves over M, C ∞ (R, M) = {ρ: R → M}. The<br />

equivalence relation defined in C ∞ (R, M) for curves having contact of first order<br />

as follows<br />

ρ 1 ≡ ρ 2<br />

iff<br />

ρ 1 (0) = ρ 2 (0) and<br />

(2.1)<br />

∀f ∈ C ∞ d<br />

(M)<br />

dt {f ◦ ρ 1} ∣ t=0<br />

= d dt {f ◦ ρ 2} ∣ t=0<br />

endowes the factor space with a differentiable manifold structure which is the tangent<br />

manifold of M,<br />

T M = C ∞ (R, M)/ ≡ . (2.2)


<strong>SECTIONS</strong> ALNG A <strong>MAP</strong> IN MECHANICS 3<br />

The natural projection τ M : T M → M (afterwards denoted by τ), mapping each<br />

equivalence class of curves [ρ] to the initial point m = ρ(0) ∈ M, defines a vector<br />

bundle structure (T M, τ, M), the well known tangent bundle. The dual bundle of<br />

T M is the cotangent bundle π : T ∗ M → M. The set of sections of τ and π, i.e.,<br />

vector fields and 1-forms on M, respectively, will be denoted by X(M) and ∧1 (M).<br />

Higher order tangent bundles of a manifold M are defined in a similar way by<br />

making use of an equivalence relation for curves having a contact of higher order.<br />

The set of equivalence classes of curves on C ∞ (R, M) with a contact of order k is the<br />

k-th order tangent manifold of M, T k M, with natural projection τ k,0 : T k M → M.<br />

For k > 1 the bundle structure of (T k M, τ k,0 , M) is not linear, but of polynomial<br />

type. Elements of T k M are denoted [ρ] k , where ρ is a curve of the corresponding<br />

equivalence class. Sometimes we write Q for an element of T k M if we do not need<br />

to make any reference to an element of the class Q.<br />

Given a local coordinate system of M, (U, q i ) i = 1, . . . , d = dim M, with<br />

U ⊂ M an open subset, there are induced coordinate systems for T k M given by<br />

(U k , q(n) i ) i = 1, . . . , d, n = 0, . . . , k<br />

(2.3)<br />

U k = τ −1<br />

k (U) qi (n) ([ρ]k ) = dn {<br />

q i<br />

dt n ◦ ρ } ∣ t=0<br />

.<br />

The natural projections τ k,l : T k M → T l M, for 0 < l < k, are given by<br />

τ k,l ([ρ] k ) = [ρ] l . A tangent vector v ∈ T (T k M) is said to be τ k,l -vertical (or<br />

vertical of order l) if τ k,l∗ v = 0, i. e. if v is tangent to the fibres of τ k,l . The set of<br />

τ k,l -vertical vectors will be denoted V (τ k,l ) and it is a vector subbundle of T (T k M).<br />

A covector α ∈ T ∗ (T k M) is said to be τ k,l -semibasic (or semibasic of order l) if it<br />

vanishes on τ k,l -vertical vectors: 〈α, v〉 = 0 ∀v ∈ V (τ k,l ). The set of τ k,l -semibasic<br />

covectors is a vector subbundle of T ∗ (T k M). A vector field V ∈ X(T k M) is said<br />

to be τ k,l -vertical if its image is into V (τ k,l ). Similarly, a τ k,l -semibasic 1-form α is<br />

a 1-form α ∈ ∧1 (T k M) such that its image is made up by τ k,l -semibasic covectors.<br />

Every function f ∈ C ∞ (M) has associated a set of functions {f(n) k ∈ C∞ (T k M)}<br />

for n = 0, . . . , k given by<br />

f k (n) ([ρ]k ) = dn<br />

dt n {f ◦ ρ} ∣ ∣<br />

t=0<br />

(2.4)<br />

Similarly, for every curve ρ: R → M there is a natural lifted curve ˜ρ k : R → T k M,<br />

defined by<br />

˜ρ k (t) = [ρ t ] k with ρ t (s) = ρ(t + s). (2.5)<br />

In a local coordinate system (U, q), if q i (ρ(t)) = χ i (t), then q i (n) (˜ρk (t)) = χ ik (n) (t).<br />

Natural lifts of curves allow us to define the imbeddings i k,l : T k+l M → T l (T k M)<br />

defined by i k,l ([ρ] k+l ) = [˜ρ k ] l . These imbeddings i k,l and projections τ k,l always<br />

define commutative diagrams:<br />

T k+l M<br />

i k,l<br />

T l T k M<br />

(2.6)<br />

τ k+l,l<br />

<br />

T l M<br />

T l M<br />

Finally, as a concept generalizing the differential of a map φ: N → M, we defined<br />

the k-order prolongation φ k : T k N → T k M given by φ k [ρ] k = [φ ◦ ρ] k , where ρ is a<br />

curve in N.<br />

τ l k,0


4 JOSÉ F. CARIÑENA⋆ , CARLOS LÓPEZ† , AND EDUARDO MARTÍNEZ⋆<br />

3. Sections along a map<br />

Let π : E → M be a fibre bundle, f : N → M a differentiable map and denote<br />

f ∗ E the pull-back of the bundle E by the map f [17]. By a section of E along the<br />

map f (or over f) we mean a differentiable map σ : N → E such that π◦σ = f. The<br />

point to be remarked here is that there exists a one-to-one canonical correspondence<br />

between the set of sections along f and that of sections of the bundle f ∗ E over N.<br />

Moreover, if E is a vector bundle, then both sets can be endowed with C ∞ (N)-<br />

module structures and this correspondence is an isomorphism of C ∞ (N)-modules.<br />

For details see e.g. [17]. The most important cases for our proposals is when the<br />

vector bundle is either T M or (T ∗ M) ∧p or (T ∗ M) ⊗p ⊗ (T M) ⊗q , or something<br />

similar else, and the sections along f are then said to be vector fields, p-forms<br />

or (p, q)-type tensor fields along f, and will be denoted X(f), ∧p (f) or T p,q (f),<br />

respectively.<br />

The more traditional concepts of vector fields and forms in M arise here as vector<br />

fields and forms along the identity map in the base M, i.e. X(M) = X(id M ) and<br />

∧ p<br />

(M) = ∧p (id M ), and similarly T p,q (M) = T p,q (id M ).<br />

Two special instances of vector fields over f are the restriction of a vector field<br />

X ∈ X(M) to f(N) obtained by composition of f : N → M and X : M → T M<br />

(usually denoted X ◦f), and the image of a vector field Y ∈ X(N) under f, denoted<br />

f ∗ ◦ Y and given by (f ∗ ◦ Y )(n) = f ∗n (Y (n)). Similarly, if α ∈ ∧p (N), then the<br />

restriction α ◦ f of α to f(N) is a p-form along f.<br />

More generally, if g : P → N is a differenciable map, for any X ∈ X(f) and<br />

Y ∈ X(g), then X ◦ g and f ∗ ◦ Y are vector fields along f ◦ g. In the same way, if<br />

α ∈ ∧p (f), then α ◦ g is a p-form along f ◦ g.<br />

The set of vector fields over f are endowed with a C ∞ (N)-module structure:<br />

they can be added to each other and multiplied by functions on N. In particular,<br />

if (V, z) and (U, x) are charts in N and M, respectively, such that f(V) ⊂ U, X is<br />

a vector field along f and n ∈ V, then the expression of X n in these coordinates is<br />

X(n) = ξ i (z) ∂<br />

∂x i ∣<br />

∣f(n) , (3.1)<br />

so that every vector field X along f can be locally written as a linear combination<br />

X = ξ i( ∂<br />

∂x i ◦ f) (3.2)<br />

of the restrictions of the coordinate vector fields ∂/∂x i with coefficients ξ i in<br />

C ∞ (N).<br />

In the same way a p-form along f, α, has a local expression<br />

α = α i1...i p<br />

(dx i1 ◦ f) ∧ · · · ∧ (dx ip ◦ f), (3.3)<br />

where α i1...i p<br />

∈ C ∞ (N).<br />

As another interesting example, let us consider the tangent bundle τ : T M → M<br />

where τ plays de role of f. There exists a canonical vector field along τ, which we<br />

will denote T (0) , given by the identity map in T M. Its coordinate expression is<br />

T (0) = v i( ∂<br />

∂q i ◦ τ) . (3.4)<br />

This object will be considered in detail later on.<br />

If f is now the projection π : T ∗ M → M of the tangent bundle, the canonical<br />

object to be considered is the 1-form along π, ˇθ 0 , given by the identity map in


<strong>SECTIONS</strong> ALNG A <strong>MAP</strong> IN MECHANICS 5<br />

T ∗ M. Its coordinate expression is<br />

ˇθ 0 = p i (dq i ◦ π). (3.5)<br />

Let us suppose that f is a submersion and denote ∧ p<br />

0<br />

(f) the set of p-forms α on N<br />

satisfying the following condition:<br />

If n ∈ N and u ∈ T n N are such that f ∗n u = 0, then i u α = 0. (3.6)<br />

There exists a one-to-one correspondence between ∧p (f) and ∧ p<br />

0 (f). If ω ∈ ∧p (f)<br />

then f ∗ ◦ ω defined by<br />

(f ∗ ◦ ω) n (u 1 , . . . , u p ) = ω n (f ∗n u 1 , . . . , f ∗n u p ) ∀u i ∈ T n N i = 1, . . . , p (3.7)<br />

is a p-form in ∧ p<br />

0 (f). Conversely, if α ∈ ∧ p<br />

0 (f), n ∈ N, v 1, . . . , v p ∈ T f(n) M and<br />

v ↑ 1 , . . . , v↑ p ∈ T n N are such that f ∗n v ↑ i = v i then ˇα defined by<br />

is a p-form along f such that f ∗ ◦ ˇα = α.<br />

In coordinates, this map is given by<br />

ˇα n (v 1 , . . . , v p ) = α n (v ↑ 1 , . . . , v↑ p) (3.8)<br />

α i1...i p<br />

(dx i1 ◦ f) ∧ · · · ∧ (dx ip ◦ f) ↦→ α i1...i p<br />

dx i1 ∧ · · · ∧ dx ip .<br />

In the last example the 1-form θ 0 = π ∗ ◦ ˇθ 0 is simply the well-known canonical<br />

1-form in T ∗ M.<br />

Another interesting example of a (1, 1)-tensor field along τ is the vertical endomorphism.<br />

Let T : T M → T ∗ (T M) ⊗ T M be the (1, 1)-tensor field along τ<br />

defined as follows: T (v)X = τ ∗v X for X ∈ T v (T M). We also recall that there<br />

exists a natural identification of the vector space T m M with its tangent space at<br />

v ∈ T m M, T v (T m M) ≡ V v (T M), through the denoted vertical lift [7]<br />

ξ v : T m M → T v (T m M) = V v (T M).<br />

The vertical endomorphism S [7] can then be defined by<br />

S(X)(v) = ξ v ◦ T (v)(X)<br />

∀X ∈ T (T M).<br />

Therefore, up to the vertical lift, the vertical endomorphism is but the natural (1, 1)<br />

tensor T along τ. In local coordinates the expression of the associated tensor field<br />

is<br />

T (x, v) = δ j ∂<br />

i<br />

∂x j ⊗ dxi .<br />

∂<br />

∂x i → ∂<br />

∂v i , the vertical endomorphism S be-<br />

After the vertical lift identification<br />

comes<br />

S(x, v) =<br />

∂<br />

∂v i ⊗ dxi .<br />

Leaving aside the canonical (1,1) tensor S on T M, S : T (T M) → T (T M), there<br />

is another canonical map, the involution map J : T (T M) → T (T M) (see e.g. [6])<br />

which is not a tensor field on T M, but it can be seen as a vector field along the<br />

tangent projection τ ∗ , J ∈ X(τ ∗ ), because it has the property τ T M ◦ J = τ ∗ :<br />

T (T M) → T M; in local coordinates (x, v, ẋ, ˙v) on T (T M), the vector field J is<br />

given by<br />

( ) ( )<br />

∂<br />

∂<br />

J(x, v, ẋ, ˙v) = v<br />

+ ˙v<br />

∈ T (x,ẋ) (T M).<br />

∂x<br />

|(x,ẋ)<br />

∂ẋ<br />

|(x,ẋ)


6 JOSÉ F. CARIÑENA⋆ , CARLOS LÓPEZ† , AND EDUARDO MARTÍNEZ⋆<br />

Its higher order generalization J k : T k (T M) → T (T k M) is also a vector field along<br />

a map, J k ∈ X(τ k ), where τ k : T k (T M) → T k M is the k-th prolongation of the<br />

tangent bundle map τ,<br />

τ T k M ◦ J k = τ k .<br />

This vector field is given in local coordinates by<br />

J k (x, v; x 1 , v 1 ; . . . ; x k , v k ) =<br />

k∑<br />

l=0<br />

vl( i ∂<br />

∂x i ◦ τ k ).<br />

l<br />

The last example of this enumeration of natural sections along maps is the canonical<br />

isomorphism Ψ : T (T ∗ M) → T ∗ (T M) (see e.g. [6]), which shows, among other<br />

properties, the existence of a symplectic structure on the tangent manifold T (T ∗ M).<br />

The canonical isomorphism Ψ is a one form along the tangent projection π ∗ , Ψ ∈<br />

∧ (π∗ ), because of the property<br />

π T M ◦ Ψ = π ∗ : T (T ∗ M) → T M.<br />

Same as for the vector field J, Ψ has a higher order generalization, Ψ k : T k (T ∗ M) →<br />

T ∗ (T k M) ([6]) which is a one form along the k-th prolongation π k of π. In similar<br />

way to the case of the Poincaré 1-form θ 0 on T ∗ M, the one-form Ψ k is in<br />

correspondence with a semibasic one form πk ∗ ◦Ψ k on T k (T ∗ M), whose exterior differential<br />

defines a canonical symplectic structure on T k (T ∗ M). In local coordinates<br />

(x i , p i ; x i 1, p i1 ; . . . ; x i k , p ik) on T k (T ∗ M), the one form Ψ k is given by<br />

Ψ k (x i , p i ; . . . ; x i k, p ik ) =<br />

k∑<br />

( k l )p i(k−l) (dx i l ◦ π k ).<br />

Other examples in Lagrangian Mechanics will be given in section 8.<br />

Vector fields along f act on a function on M giving as result a function on N.<br />

If X ∈ X(f) and n ∈ N then X(n) is a tangent vector to M at the point f(n)<br />

which act on a function h ∈ C ∞ (M). Then, defining the function Xf ∈ C ∞ (N)<br />

by (Xf)(n) = X(n)f, the Leibnitz rule for tangent vectors implies the property<br />

X(hl) = f ∗ h Xl + f ∗ l Xh. A map satisfying this property is called a f ∗ -derivation<br />

(of degree 0).<br />

In a recent paper [16] the Frölicher and Nijenhuis’ theory has been generalized<br />

in order to include these new derivations. We recall that a f ∗ -derivation of degree<br />

r of scalar forms on M is a R-linear map D : ∧ (M) → ∧ (N) satisfying<br />

l=0<br />

D (∧p (M) ) ⊂ ∧ p+r (N)D(α ∧ β) = Dα ∧ f ∗ β + (−1) pr f ∗ α ∧ Dβ (3.9)<br />

for β ∈ ∧ (M) and α ∈ ∧p (M).<br />

We say that D is of type i ∗ when D vanishes on ∧0 (M) and D is of type d ∗ if<br />

D ◦ d = (−1) r d ◦ D, where d is the exterior differentiation (in M on the left hand<br />

side and in N on the right hand side).<br />

If we denote V p (f) the set of p-linear maps L from T N × · · · × T N to T M<br />

satisfying that for n ∈ N and v 1 , . . . , v p ∈ T n N, τ[L(v 1 , . . . , v p )] = f(n), then we<br />

can associate with each element L of this set a type i ∗ derivation of degree p − 1,<br />

denoted i L , as well as a type d ∗ derivation of degree p, denoted d L . Moreover every<br />

derivation D of degree r can be decomposed in a unique way as a sum D = i L + d R<br />

for some L ∈ V r+1 (f) and R ∈ V r (f).


<strong>SECTIONS</strong> ALNG A <strong>MAP</strong> IN MECHANICS 7<br />

In particular if X ∈ V 0 (f) ≡ X(f), then d X is a derivation of degree 0 such that<br />

its action over functions on M coincides with the the action of X over C ∞ (M) as<br />

defined above.<br />

Suppose that f is a submersion, P is another differentiable manifold and that<br />

g : P → N is also a submersion. If α ∈ ∧p ∧<br />

(f) and X ∈ X(g) then ᾱ = f ∗ ◦ α ∈<br />

p<br />

0 (f), and we can act with d X on ᾱ obtaining d X ᾱ ∈ ∧p (P ). It can be shown<br />

that d X ᾱ ∈ ∧ p<br />

0 (g) (for all α ∈ ∧p (f)) if and only if there exists Y ∈ X(f) such<br />

that f ∗ ◦ X = Y ◦ g. In this case (d X ᾱ) ∨ is a p-form along g that we will simply<br />

denote d X α.<br />

Similarly, for X ∈ X(g) we can act with i X on ᾱ, and so we have a (p − 1)-form<br />

i X ᾱ on P , that always belongs to ∧ p<br />

0<br />

(g). The point is that if there exists Y ∈ X(f)<br />

such that f ∗ ◦ X = Y ◦ g, then there will be a (p − 1)-form β in ∧ p−1<br />

0<br />

(f) such that<br />

f ∗ β = i X ᾱ, and we can consider β as being a (p − 1)-form along f. This form will<br />

be denoted i X α.<br />

4. Lifting vector fields along projections<br />

In the following we will be concerned with cases in which f are the natural<br />

projections τ l,k : T l M → T k M, for k and l non-negative integer numbers such that<br />

l > k (and τ 1,0 ≡ τ).<br />

It is well-known [8] that every function f ∈ C ∞ (M) defines functions f(n) r ∈<br />

C ∞ (T r M) for every n ≤ r by means of<br />

f(n)<br />

r (<br />

[ρ]<br />

r ) = dn<br />

dt n {f ◦ ρ} ∣ t=0<br />

. (4.1)<br />

We can extend this construction to functions F ∈ C ∞ (T k M) by making use of the<br />

canonical immersion<br />

i k,r : T k+r M −→ T r (T k M)<br />

[ρ] k+r ↦−→ [˜ρ k ] r , (4.2)<br />

where ˜ρ k is the lifting to T k M of the curve ρ, as follows: F defines F(n) r ∈<br />

C ∞ (T r (T k M)), and then F[n] r = i ∗ k,r F (n) r is a function on T k+r M. If [ρ] k+r ∈<br />

T k+r M then<br />

F[n]<br />

r (<br />

[ρ]<br />

k+r ) = dn {<br />

F ◦ ˜ρ<br />

k } ∣ dt n t=0<br />

. (4.3)<br />

Let now X be a vector field along τ k,0 . There exists one vector field X (r) over<br />

τ k+r,r such that<br />

X (r) f(n) r = (Xf)r [n] , (4.4)<br />

for any function f ∈ C ∞ (M).<br />

We will give an explicit construction of such vector field over τ k+r,r . Let us<br />

choose ¯X ∈ X(T k M) such that τ k,0∗ ◦ ¯X = X. Let [ρ] k+r ∈ T k+r M be an arbitrary<br />

but fixed point of T k+r M. Then the map χ: U ⊂ R 2 → M defined by<br />

χ(s, t) = (τ k,0 ◦ ¯X s ◦ ˜ρ k )(t), (4.5)<br />

where ¯X s is the flow of ¯X and U is an open neighbourhood of (0, 0) ∈ R 2 , can be<br />

used to determine a vector ξ tangent to T r M at the point [ρ] r . More precisely,<br />

for any real number s, the curve in M, χ s (t) = χ(s, t), defines a point in T r M.<br />

As a function of the parameter s we obtain a curve in T r M whose tangent vector<br />

at s = 0 is the abovementioned vector ξ at [ρ] r , and in this way we will find a<br />

vector field X (r) along τ k+r,r . This vector field is well defined: if ¯X′ ∈ X(T k M)


8 JOSÉ F. CARIÑENA⋆ , CARLOS LÓPEZ† , AND EDUARDO MARTÍNEZ⋆<br />

is a different vector field τ k,0∗ -related to X and ρ ′ is another curve in M defining<br />

the same (k + r)-velocity, [ρ] k+r = [ρ ′ ] k+r , then the tangent vectors defined by<br />

χ(s, t) = (τ k,0 ◦ ¯X s ◦ ˜ρ k )(t) and by χ ′ (s, t) = (τ k,0 ◦ ¯X s ′ ◦ ˜ρ ′k )(t) coincide. Indeed,<br />

note that for any function f ∈ C ∞ (M) we have (f ◦ χ)(0, t) = (f ◦ ρ)(t) and<br />

(f ◦ χ ′ )(0, t) = (f ◦ ρ ′ )(t) so that<br />

∂ n<br />

∂t n {f ◦ ρ} ∣ ∣<br />

(0,0)<br />

= ∂n<br />

∂t n {f ◦ ρ′ } ∣ ∣<br />

(0,0)<br />

, (4.6)<br />

because ρ and ρ ′ define the same r-velocity.<br />

Moreover, note that<br />

∂<br />

∂s {f ◦ χ} ∣ ∣<br />

s=0<br />

= ∂ ∂s<br />

{<br />

(f ◦ τk,0 ) ( X s<br />

(˜ρ k (t) ))} ∣ ∣<br />

s=0<br />

= (τ k,0∗ ¯X)˜ρk (t)f<br />

= X˜ρk (t)f<br />

and a similar result holds for χ ′ . Then:<br />

and<br />

so that<br />

= [ (Xf) ◦ ˜ρ k] (t),<br />

∂ n+1<br />

∂s∂t n {f ◦ χ} ∣ ∣<br />

(0,0)<br />

= dn<br />

dt n {<br />

(Xf) ◦ ˜ρ<br />

k } ∣ ∣<br />

t=0<br />

= (Xf) r [n](<br />

[ρ]<br />

k+r ) ,<br />

(4.7)<br />

(4.8)<br />

∂ n+1<br />

∂s∂t n {f ◦ χ′ } ∣ (0,0)<br />

= (Xf) r (<br />

[n] [ρ ′ ] k+r) , (4.9)<br />

∂ n+1<br />

∂s∂t n {f ◦ χ} ∣ (0,0)<br />

= ∂n+1<br />

∂s∂t n {f ◦ χ′ } ∣ (0,0)<br />

. (4.10)<br />

This proves that X (r)( [ρ] k+r) is well defined.<br />

From (4.9) we see that<br />

(Xf) r (<br />

[n] [ρ]<br />

k+r ) = ∂n+1<br />

∂s∂t n {f ◦ χ} ∣ (0,0)<br />

= d { ∂<br />

n<br />

ds ∂t n {f ◦ χ s} ∣ } ∣∣s=0<br />

t=0<br />

= d {<br />

f r<br />

ds<br />

(n)(<br />

[χs ] r)} ∣<br />

∣s=0<br />

(4.11)<br />

= ( X (r) f r (n))(<br />

[ρ]<br />

k+r ) ,<br />

so that X (r) satisfies the condition (4.4).<br />

Since two tangent vectors agreeing over functions of the type f(n) r must be equal,<br />

we can conclude that X (r) is uniquely defined.<br />

As a consequence of (4.4) we have for every r > l ≥ 0<br />

τ r,l∗ ◦ X (r) = X (l) ◦ τ k+r,k+l , (4.12)<br />

(provided that X (0) = X). Indeed, for n ≤ l we have that f(n) r = τ r,l ∗ f(n) l and<br />

F[n] r = τ k+r,k+l ∗ F[n] l , for f ∈ C∞ (M) and F ∈ C ∞ (T k M), so that from (4.4):<br />

(τ r,l∗ ◦ X (r) )f l (n) = X(r) (τ r,l ∗ f l (n) ) = X(r) (f r (n) ) = (Xf)r [n] , (4.13)


<strong>SECTIONS</strong> ALNG A <strong>MAP</strong> IN MECHANICS 9<br />

and<br />

(X (l) ◦ τ k+r,k+l )f l (n) = τ k+r,k+l ∗ (X (l) f l (n) )<br />

= τ ∗ k+r,k+l (Xf) l [n]<br />

(4.14)<br />

= (Xf) r [n] ,<br />

what proves (4.12)<br />

In the particular case of X being the restriction of a vector field Y ∈ X(M),<br />

X = Y ◦ τ k,0 , the vector field X (r) reduces to the restriction of the complete lift<br />

Y c,r ∈ X(T r M), X (r) = Y c,r ◦ τ k+r,r . This is the reason why we call X (r) ∈<br />

X(τ k+r,r ) the generalized complete lift of X ∈ X(τ k,0 ).<br />

If we consider a local chart (V, q) of M and the induced charts ( ¯V k , ¯q k ) of T k M,<br />

then, if the coordinate expression for X is<br />

the corresponding expression for X (r) is<br />

r∑<br />

X (r) = (η i ) r ( ∂<br />

[n]<br />

X = η i( ∂<br />

∂q i ◦ τ) , (4.15)<br />

n=0<br />

∂q i (n)<br />

5. The total time derivative operator<br />

◦ τ k+r,r<br />

)<br />

. (4.16)<br />

There exists a canonical vector field along τ k+1,k [8] given by the canonical<br />

immersion (4.2) for r = 1, namely<br />

i k,1 : T k+1 M −→ T (T k M)<br />

[ρ] k+1 ↦−→ [˜ρ k ] 1 = ˙˜ρ k .<br />

Its action on functions F ∈ C ∞ (T k M) is given by<br />

(<br />

dik,1 F )( [ρ] k+1) = i k,1<br />

(<br />

[ρ]<br />

k+1 ) F<br />

(5.1)<br />

= ˙˜ρ k F<br />

= d dt<br />

{<br />

F ◦ ˜ρ<br />

k } ∣ ∣<br />

t=0<br />

,<br />

(5.2)<br />

showing that d ik,1 F = F[1] 1 .<br />

When k = 0, the map i 0,1 is but the identity in T M, so that i 0,1 = T (0) , the<br />

vector field along τ 1,0 described in Section 3. T (0) can be lifted to a vector field<br />

T (k) along τ k+1,k . The resulting vector field T (k) is but i k,1 . Indeed, if F = H k−m<br />

[n]<br />

for H ∈ C ∞ (T m M) then<br />

so that<br />

(<br />

dik,1 H k−m )(<br />

[n] [ρ]<br />

k+1 ) = d {H k−m<br />

dt<br />

[n]<br />

◦ ˜ρ k} ∣<br />

∣t=0<br />

= d { d<br />

n<br />

dt ds n {H ◦ ˜ρm } ∣ } ∣∣t=0<br />

s=t<br />

d ik,1 H k−m<br />

[n]<br />

= dn+1<br />

dt n+1 {H ◦ ˜ρm } ∣ t=0<br />

= ( H k−m+1 )(<br />

[n+1] [ρ]<br />

k+1 ) ,<br />

(5.3)<br />

= H k−m+1<br />

[n+1]<br />

. (5.4)


10 JOSÉ F. CARIÑENA⋆ , CARLOS LÓPEZ† , AND EDUARDO MARTÍNEZ⋆<br />

When taking m = 0 in (5.4) we have that for any function f ∈ C ∞ (M)<br />

d ik,1 f k (n) = f k+1<br />

(n+1) = (T (0) f) k [n] (5.5)<br />

and taking into account that the different liftings of T (0) are characterized by this<br />

condition, we can assure that i k,i = T (k) . The derivation d T (k) is called the total<br />

time derivative operator in T k M.<br />

The lifting X (r) of a vector field X ∈ X(τ k,0 ) satisfies the following ‘commutation’<br />

rule<br />

d X (r+1) ◦ d T (r) − d T (k+r) ◦ d X (r) = 0 (5.6)<br />

Indeed, if f ∈ C ∞ (M), then using (5.5) and (4.5)<br />

(<br />

dX (r+1) ◦ d T (r))<br />

(f<br />

r<br />

(n) ) = d X (r+1)(f r+1<br />

(n+1) ) = (Xf)r+1 [n+1] , (5.7)<br />

and ( ) )<br />

dT (k+r) ◦ d X (r) (f<br />

r<br />

(n) ) = d T (k+r)(<br />

(Xf)<br />

r<br />

[n] = (Xf)<br />

r+1<br />

[n+1] , (5.8)<br />

what proves (5.6).<br />

The property (5.6) is equivalent to<br />

i X (r+1) ◦ d T (r) − d T (k+r) ◦ i X (r) = 0 (5.9)<br />

because on C ∞ (M) both terms vanish, and on exact forms df<br />

i X (r+1) ◦ d T (r)df − d T (k+r) ◦ i X (r)df =<br />

= ( i X (r+1) ◦ d ◦ d T (r) − d T (k+r) ◦ i X (r) ◦ d ) f<br />

= ( (5.10)<br />

)<br />

d X (r+1) ◦ d T (r) − d T (k+r) ◦ d X (r) f = 0.<br />

Moreover, property (5.6) ( or equivalently (5.9)) characterizes prolongations of vector<br />

fields along τ k,0 .<br />

As a consequence of this property we have that if α ∈ ∧ (τ r,l ) then<br />

if r ≤ k + l, and<br />

if r > k + l.<br />

i X (l+1)(<br />

dT (r)α ) = d T (k+l)(<br />

iX lα ) ∈ ∧ (τ k+l+1,l+1 ), (5.11)<br />

i X (r+1)(<br />

dT (r)α ) = d T (r)(<br />

iX lα ) ∈ ∧ (τ r+1,l+1 ), (5.12)<br />

6. The vertical endomorphism and the variational operator<br />

There exists a canonical type (1, 1) tensor field S k on the k-th order tangent bundle<br />

T k M called the vertical endomorphism [8]. It is defined by means of the vertical<br />

lifting map ξ V Q : T τ k,k−1 (Q)(T k−1 M) → T Q (T k M), whose image is the set V (τ k,0 )<br />

of the τ k,0 -vertical tangent vectors to T k M, and the differental of the projection<br />

τ k,k−1 : T k M → T k−1 M at the point Q; for v ∈ T Q (T k M)<br />

Its local expression in natural coordinates is<br />

and it satisfies<br />

(1) S k+1<br />

k<br />

= 0<br />

(2) Kern(Sk n) = V (τ k,k−n)<br />

S k (v) = ξ V Q(τ k,k−1∗Q v). (6.1)<br />

k−1<br />

∑<br />

S k = (n + 1)<br />

n=0<br />

∂<br />

∂q i (n+1)<br />

⊗dq i (n) , (6.2)


<strong>SECTIONS</strong> ALNG A <strong>MAP</strong> IN MECHANICS 11<br />

(3) Im(S n k ) = V (τ k,n−1)<br />

(4) The Nijenhuis tensor [S k , S k ] vanishes<br />

(5) L ∆ S k = −S k<br />

(6) S ∗ k (df (n)) = n df (n−1)<br />

where n = 0, . . . , k and ∆ is the Liouville vector field on T k M defined by ∆(Q) =<br />

ξ V Q (T (k−1) (Q)).<br />

Proposition 6.1. Let X be a vector field along τ l,0 . The operator [X (k) , S n k ] mapping<br />

∧1 (T k M) in ∧1 (T k+l M) and defined by<br />

for n = 1, . . . , k + l, satisfies<br />

where F is a function on T k M.<br />

[X (k) , S n k ] = d X (k) ◦ S ∗n<br />

k<br />

− S ∗n<br />

k+l ◦ d X (k), (6.3)<br />

[X (k) , S n k ](F α) = (τ k+l,k ∗ F )[X (k) , S n k ](α), (6.4)<br />

Proof. Using that d X (k)<br />

is a derivation along τ k+l,k we have<br />

(F α) =<br />

because τ k+l,k ∗ ◦ S ∗n<br />

k<br />

= d X (k)(F Sk<br />

∗n (α)) − Sk+l(τ ∗n ∗ k+l,k F d X (k)α + d X (k)F τ ∗ k+l,k α)<br />

= d X (k)F τ ∗ k+l,k (Sk<br />

∗n (α)) + τ ∗ k+l,k F d X (k)(Sk<br />

∗n (α))<br />

− τ ∗ k+l,k F Sk+l(d ∗n<br />

X (k)α) − d X (k)F Sk+l(τ n ∗ k+l,k (α))<br />

= τ k+l,k ∗ F [X (k) , S n k ](α),<br />

= S ∗n<br />

k+l ◦ τ k+l,k ∗ .<br />

(6.5)<br />

□<br />

In particular, if X = T (0) , i.e. the canonical vector field along τ 1,0 , we have [8]<br />

[T (k) , S n k ] = −nτ k+1,k ∗ ◦ S ∗n−1<br />

k<br />

. (6.6)<br />

It is easy to see from this equation (by induction, for instance) that<br />

min{m,n}<br />

∑<br />

d m ◦ S ∗n<br />

T (k) k =<br />

r=0<br />

(−1) r ( m<br />

r<br />

) n!<br />

(n − r)! τ k+m,k+m−r ∗ ◦ S ∗n−r<br />

k+m−r ◦ dm−r T (k) , (6.7)<br />

(where we have written d n for d<br />

T (m) T (m+n−1) ◦ · · · ◦ d T (m)) a relation that we will use<br />

later in this section.<br />

Proposition 6.2. Let X be a vector field along τ l,0 and f a function in T l M.<br />

Then<br />

k∑<br />

(fX) (k) 1<br />

=<br />

n! (τ k+l,l+n ∗ d n f)S n T (l) k (X (k) ). (6.8)<br />

n=0<br />

Proof. Let us write Y k for Y k = ∑ k<br />

n=0 1 n! (τ k+l,l+n ∗ d n T (l) f)S n k (X(k) ). First we prove<br />

that Y k commutes with T (k) ; for α ∈ ∧1 (T k M) taking into account (5.9) and (6.6)


12 JOSÉ F. CARIÑENA⋆ , CARLOS LÓPEZ† , AND EDUARDO MARTÍNEZ⋆<br />

we see that<br />

(d T (k+l) ◦ i Y k − i Y k+1 ◦ d T (k))α =<br />

( k∑<br />

)<br />

1<br />

= d T (k+l)<br />

n! (τ k+l,l+n ∗ d n f)i<br />

T (l) X (k)Sk<br />

∗n (α)<br />

n=0<br />

=<br />

=<br />

k+1<br />

∑<br />

−<br />

n=0<br />

n=0<br />

1<br />

n! (τ k+l,l+n ∗ d n f)i<br />

T (l) X (k+1)Sk<br />

∗n (d T (k)α)<br />

k∑ 1<br />

n! (τ k+l+1,l+n+1 ∗ d n+1 f)τ ∗ T (l) k+l+1,k+l i X (k)Sk<br />

∗n (α)<br />

k+1<br />

∑ 1<br />

+<br />

n! (τ k+l+1,l+n ∗ d n f)i<br />

T (l) X (k+1)[T (k) , Sk n ](α)<br />

n=0<br />

= 0.<br />

n=0<br />

k∑ 1<br />

n! (τ k+l+1,l+n+1 ∗ d n+1 f)τ ∗ T (l) k+l+1,k+l i X (k)Sk<br />

∗n (α)<br />

k+1<br />

∑<br />

−<br />

n=1<br />

1<br />

n! (τ k+l+1,l+n ∗ d n f)i<br />

T (l) X (k+1)τ ∗ k+1,k S ∗n−1<br />

k<br />

(α)<br />

(6.9)<br />

This implies that Y k is the prolongation of a vector field along τ l,0 . Since τ k,0∗ ◦Y k =<br />

fX we conclude Y k = (fX) (k) .<br />

□<br />

Let α be a 1-form on T k M. Then, for s = 1, . . . , k we can define operators<br />

: ∧1 (T k M) → ∧1 (T 2k−s M) be means of<br />

〈<br />

〉<br />

∑s−1<br />

∗<br />

τ 2k+l−s,k+l α, (fX) (k) 1<br />

−<br />

n! (τ k+l,l+n ∗ d n f)S n T (l) k (X (k) )<br />

D (k)<br />

s<br />

= (τ 2k+l−s,s+l ∗ d s T (l) f)<br />

n=0<br />

〈<br />

D (k)<br />

s (α), X (2k−s)〉 mod d T ,<br />

(6.10)<br />

where mod d T means that the equality holds except for the addition of the total<br />

time derivative of a function. A straightforward calculation shows that<br />

k∑<br />

D (k)<br />

s = (−1) n+s 1 n! τ 2k−s,k+n−s ∗ ◦ d n−s ◦ S ∗n<br />

T (k) k . (6.11)<br />

n=s<br />

From this expression it is easy to see that the image of D (k)<br />

s<br />

forms, and that the following relations holds<br />

are τ 2k−s,k−s -semibasic<br />

D (k)<br />

s−1 = 1 s! τ 2k−s+1,k ∗ ◦ S ∗s−1<br />

k<br />

− d T (2k−s) ◦ D (k)<br />

s (6.12)<br />

S 2k−s ◦ D (k)<br />

s<br />

= sτ 2k−s,2k−s−1 ∗ ◦ D (k)<br />

s+1 . (6.13)<br />

Equation (6.13) implies that only D (k)<br />

0 and D (k)<br />

1 are relevant and they are called<br />

the variational operator E (k) and the Cartan operator S (k) , respectively. They are<br />

related by<br />

E (k) = τ 2k,k ∗ − d T (2k−1) ◦ S (k) , (6.14)<br />

as the equation (6.12) claims.


<strong>SECTIONS</strong> ALNG A <strong>MAP</strong> IN MECHANICS 13<br />

Another usefull expression for D (k) is given by<br />

D (k)<br />

s =<br />

k−s<br />

∑<br />

n,m=0<br />

(−1) n 1 m<br />

(n + s)!(<br />

n<br />

)<br />

τ 2k−s,k+n ∗ ◦ S ∗n+s<br />

k+n ◦ dn T (k) , (6.15)<br />

as it can be easily proved using the expression (6.7). In this sum we have assumed<br />

the convention ( m<br />

n)<br />

= 0 for m < n, otherwise the sum must be only for k − s ≥<br />

m ≥ n ≥ 0.<br />

Proposition 6.3. Let X be a vector field along τ l,0 and define a map [X (k) , S (k) ]<br />

mapping ∧1 (T k M) in ∧1 (T 2k+2l−1 M) by means of<br />

[X (k) , S (k) ] = τ 2k+2l−1,2k+l−1 ∗ ◦ d X (2k−1) ◦ S (k) − S (k+l) ◦ d X (k). (6.16)<br />

Then if α is a τ k,0 -semibasic 1-form on T k M, we have<br />

k+l−1<br />

∑<br />

[X (k) , S (k) ](α) = { (−1) n 1<br />

n,m=0<br />

(n + 1)!( m<br />

n<br />

)<br />

×<br />

τ 2k+2l−1,k+l+n ∗ ◦ [X (k+n) , S n+1<br />

k+n ] ◦ dn T (k) }(α). (6.17)<br />

Proof. Indeed, using the expressions (6.15) and (5.6) we have<br />

[X (k) , S (k) ] =<br />

=<br />

=<br />

k−1<br />

∑<br />

n,m=0<br />

+<br />

k+l−1<br />

∑<br />

n,m=0<br />

k+l−1<br />

∑<br />

n,m=0<br />

(−1) n 1 m<br />

(n + 1)!(<br />

n<br />

(−1) n 1 m<br />

(n + 1)!(<br />

n<br />

(−1) n 1 m<br />

(n + 1)!(<br />

n<br />

− ∑ ∗(−1) n 1<br />

(n + 1)!( m<br />

n<br />

)<br />

τ ∗ 2k+2l−1,k+n+l ◦ d X (k+n) ◦ S ∗n+1<br />

k+n<br />

◦ dn T (k)<br />

)<br />

τ 2k+2l−1,k+n+l ∗ ◦ S ∗n+1<br />

k+l+n ◦ d X (k+n) ◦ dn T (k)<br />

)<br />

τ 2k+2l−1,k+l+n ∗ ◦ [X (k+n) , S n+1<br />

k+n ]∗ ◦ d n T (k)<br />

)<br />

τ 2k+2l−1,k+n+l ∗ ◦ d X (k+n) ◦ S ∗n+1<br />

k+n ◦ dn T (k) (6.18)<br />

where ∑ ∗ means sum for k − 1 ≤ n ≤ k + l − 1, 0 ≤ m ≤ k + l − 1 and k − 1 ≤ m ≤<br />

k +l −1, 0 ≤ n ≤ k −1. But this sum vanishes on τ k,0 -semibasic forms, because if α<br />

is τ k,0 -semibasic then d n α is τ<br />

T (k) k+n,n -semibasic and thus S ∗n+1<br />

k+n (dn α) = 0. □<br />

T (k)<br />

What it is important in this relation is that the right hand side is a sum of<br />

‘linear’ operators acting on total time derivatives of α.<br />

In particular if X = T (0) , then<br />

[T (k) , S (k) ] = τ 2k+1,2k ∗ ◦ E (k) (6.19)<br />

because S (k) ◦ d T (k−1) = τ ∗ 2k−1,k−1 and (6.14). Moreover, this implies that<br />

[X (2k) , S (2k) ] ◦ E (k) = τ 4k+2l−1,2k+2l−1 ∗ [X (k) , S (k) ]. (6.20)


14 JOSÉ F. CARIÑENA⋆ , CARLOS LÓPEZ† , AND EDUARDO MARTÍNEZ⋆<br />

Indeed,<br />

[X (2k) , S (2k) ] ◦ E (k) =<br />

= [X (2k) , S (2k) ] ◦ (τ ∗ 2k,k − d T (2k−1) ◦ S (k) )<br />

= [X (2k) , S (2k) ] ◦ τ ∗ 2k,k + S (2k+l) ◦ d T (2k+l−1) ◦ d X (2k−1) ◦ S (k)<br />

− τ ∗ 4k+2l−1,4k+l−1 ◦ d X (4k−1) ◦ S (2k) ◦ d T (2k−1) ◦ S (k)<br />

= τ ∗ 4k+2l−1,2k+2l−1 ◦ [X (k) , S (k) ] + τ ∗ 4k+2l−1,2k+l−1 ◦ d X (2k−1) ◦ S (k)<br />

− τ ∗ 4k+2l−1,4k+l−1 ◦ d X (4k−1) ◦ τ ∗ 4k−1,2k−1 ◦ S (k)<br />

= τ ∗ 4k+2l−1,2k+2l−1 ◦ [X (k) , S (k) ].<br />

(6.21)<br />

7. Ordinary differential equations<br />

The geometrical interpretation of a system of (N + 1)st-order differential equations<br />

(hereafter shortened to (N + 1)-ode)<br />

(<br />

φ i q, dq<br />

)<br />

dt , . . . , dN+1 q<br />

dt N+1 = 0 (7.1)<br />

is given by a closed imbeded submanifold S of T N+1 M satisfying [19]<br />

τ N+1,N −1 (τ N+1,N S) ≠ S. (7.2)<br />

The solutions of this system are the curves σ : R → M whose lift to T N+1 M,<br />

˜ρ N+1 : R → T N+1 M, takes values in S.<br />

We are interested on (N + 1)-ode that can be presented in normal form:<br />

d N+1 (<br />

q<br />

dt N+1 = f i q, dq )<br />

dt , . . . , dN q<br />

dt N , (7.3)<br />

which corresponds to the image of a section γ of τ N+1,N , S = Im(γ). Then, a<br />

solution of this (N + 1)-ode is a curve ρ in M such that ˜ρ N+1 = γ ◦ ˜ρ N . The<br />

composition map Γ = i 1,N ◦ γ, or equivalently Γ = T (N) ◦ γ, is a vector field on<br />

T N M that is also called a (N + 1)-ode because its integral curves are natural<br />

liftings of solutions of γ. The associated derivations d Γ and d T (N) are related by<br />

d Γ = γ ∗ ◦ d T (N). (7.4)<br />

We can associate with every (N + 1)-ode a sequence of differential equations of<br />

order greater than (N + 1).<br />

Proposition 7.1. There exists exactly one (N + 2)-ode ˜Γ (1) in M satisfying<br />

This ode is said to be the first ode-prolongation of Γ.<br />

d˜Γ(1) ◦ d T (N) = d T (N) ◦ d Γ . (7.5)<br />

Proof. In order to prove this we consider and atlas an a partition of unity subordinate<br />

to this atlas. In a local chart the condition (7.5) imposes<br />

˜Γ (1) q i (n) = qi (n+1)<br />

˜Γ (1) q i (N+1) = d T (N)fi ,<br />

n = 1, . . . , N<br />

(7.6)


<strong>SECTIONS</strong> ALNG A <strong>MAP</strong> IN MECHANICS 15<br />

where f i are given by f i = Γq(N) i . Since ˜Γ (1) must be a ode, its expression in this<br />

chart is<br />

N∑<br />

˜Γ (1) = q i ∂<br />

(n+1)<br />

∂q i + d T (N)f i ∂<br />

n=0<br />

(n)<br />

∂q(N+1)<br />

i . (7.7)<br />

Now using the partition of unity we obtain a global (N + 2)-ode.<br />

□<br />

The l-th ode-prolongation ˜Γ (l) of Γ is defined by recurrence as the first odeprolongation<br />

of the (l − 1)-th ode-prolongation. The section of the projection<br />

τ N+l+1,N+l associated to ˜Γ (l) will be denoted ˜γ (l) . We can construct from they the<br />

sections γ (l) : T N M → T N+l+1 M of τ N+l+1,N defined by γ (l) = ˜γ (l) ◦ ˜γ (l−1) ◦· · ·◦γ.<br />

We call γ (l) the l-th prolongation of γ because they satisfy<br />

τ N+l+1,N+m+1 ◦ γ (l) = γ (m) , for l > m > 0, (7.8)<br />

and a curve ρ in M is a solution of γ, ˜ρ N+1 = γ ◦ ˜ρ N , if and only if ρ is also a<br />

solution of γ (l) : ˜ρ N+l+1 = γ (l) ◦ ˜ρ N . The equation (7.8) also holds for m = −1, 0 if<br />

we assume the convention γ (−1) = id T N M and γ (0) = γ.<br />

Lemma 7.2. The prolongations of a (N + 1)-ode γ satisfy<br />

for l ≥ 0, and<br />

for l ≥ r.<br />

d Γ ◦ γ (l)∗ = γ (l)∗ ◦ d˜Γ(l+1), (7.9)<br />

γ (l)∗ ◦ d l−r+1 = d l−r+1<br />

T (N+r) Γ<br />

◦ γ (r−1)∗ , (7.10)<br />

Proof. We prove (7.9) by induction on l. Since the definition of ˜Γ (1) we have<br />

d˜Γ(1) ◦ d T (N) − d T (N) ◦ d Γ = 0 and composing with γ ∗<br />

γ ∗ ◦ (d˜Γ(1) ◦ d T (N) − d T (N) ◦ d Γ ) = γ ∗ ◦ d˜Γ(1) ◦ d T (N) − d Γ ◦ d Γ<br />

= (γ ∗ ◦ d˜Γ(1) − d Γ ◦ γ ∗ ) ◦ d T (N),<br />

so that γ ∗ ◦ d˜Γ(1) − d Γ ◦ γ ∗ vanishes on functions of the type f (N+1) . Moreover<br />

(7.11)<br />

(γ ∗ ◦ d˜Γ(1) − d Γ ◦ γ ∗ ) ◦ τ N+1,N ∗ = γ ∗ ◦ d T (N) − d Γ = 0, (7.12)<br />

and then γ ∗ −d ◦d˜Γ(1) Γ ◦γ ∗ vanishes on functions of the type f (n) for all n ≥ 0, what<br />

implies that it vanishes on every function. This proves (7.9) for l = 0. Note that<br />

this implies ◦ ˜γ (l) = ˜γ (l) ◦ since ˜Γ (l+1) is the first ode-prolongation of<br />

d˜Γ(l) d˜Γ(l+1)<br />

˜Γ (l) . Then assuming that d Γ ◦ γ (l−1)∗ = γ (l−1)∗ ◦ ˜Γ (l) holds and taking into account<br />

that γ (l) = ˜γ (l) ◦ γ (l−1) we have:<br />

d Γ ◦ γ (l)∗ = d Γ ◦ γ (l−1)∗ ◦ ˜γ (l)∗<br />

= γ (l−1)∗ ◦ d˜Γ(l) ◦ ˜γ (l)∗<br />

= γ (l−1)∗ ◦ ˜γ (l) ◦ d˜Γ(l+1)<br />

(7.13)<br />

= γ (l)∗ ◦ d˜Γ(l+1).<br />

Now, we also prove (7.10) by induction on l. For l = r since (7.4):<br />

γ (r)∗ ◦ d T (N+r) = γ (r−1)∗ ◦ ˜γ (r)∗<br />

= γ (r−1)∗ ◦ d˜Γ(r)<br />

= d Γ ◦ γ (r−1)∗ .<br />

(7.14)


16 JOSÉ F. CARIÑENA⋆ , CARLOS LÓPEZ† , AND EDUARDO MARTÍNEZ⋆<br />

Asuming that<br />

holds, we have<br />

γ (l−1)∗ ◦ d T (N+l−1) ◦ · · · ◦ d T (N+r)<br />

= d l−r<br />

Γ<br />

◦ γ (r−1)∗ (7.15)<br />

γ (l)∗ ◦ d T (N+l) ◦ d T (N+l−1) ◦ · · · ◦ d T (N+r) =<br />

= γ (l−1)∗ ◦ ˜γ (l)∗ ◦ d T (N+l) ◦ d T (N+l−1) ◦ · · · ◦ d T (N)<br />

= γ (l−1)∗ ◦ d˜Γ(l) ◦ d T (N+l−1) ◦ · · · ◦ d T (N+r)<br />

= d Γ ◦ γ (l−1)∗ ◦ d T (N+l−1) ◦ · · · ◦ d T (N)<br />

= d Γ ◦ d l−r<br />

Γ<br />

◦ γ (r−1)∗<br />

= d l−r+1<br />

Γ<br />

◦ γ (r−1)∗ .<br />

(7.16)<br />

□<br />

The following proposition gives us the relation between the operators S (N) and<br />

E (N) here defined and the operators σ Γ and E Γ defined by Crampin et. al. [8]:<br />

σ Γ =<br />

E Γ =<br />

N∑<br />

(−1) n+1 1 n! dn−1<br />

n=1<br />

Γ<br />

◦ S ∗n<br />

N ,<br />

N∑<br />

(−1) n 1 n! dn Γ ◦ S ∗n<br />

n=0<br />

N .<br />

(7.17)<br />

Proposition 7.3. If Γ is a (N + 1)-ode and γ is the section associated to Γ, then<br />

(1) γ (N−2)∗ ◦ S (N) = σ Γ ,<br />

(2) γ (N−1)∗ ◦ E (N) = E Γ .<br />

Proof. Using (7.10), (7.8) and the definition of S (N) we have<br />

γ (N−2)∗ ◦ S (N) =<br />

=<br />

=<br />

that proves (1), and<br />

N∑<br />

(−1) n+1 1 n! γ(N−2)∗ ◦ τ ∗ 2N−1,N+n−1 ◦ d n−1 ◦ S ∗n<br />

T (N) N<br />

n=1<br />

N∑<br />

(−1) n+1 1 n! γ(n−2)∗ ◦ d n−1 ◦ S ∗n<br />

T (N) N<br />

n=1<br />

N∑<br />

(−1) n+1 1 n! dn−1<br />

n=1<br />

γ (N−1)∗ ◦ E (N) =<br />

that proves (2).<br />

=<br />

=<br />

Γ<br />

◦ S ∗n<br />

N ,<br />

N∑<br />

(−1) n+1 1 n! γ(N−1)∗ ◦ τ ∗ 2N,N+n ◦ d n ◦ S ∗n<br />

T (N)<br />

n=0<br />

N∑<br />

(−1) n+1 1 n! γ(n−1)∗ ◦ d n ◦ S ∗n<br />

T (N)<br />

n=0<br />

N∑<br />

(−1) n+1 1 n! dn Γ ◦ S ∗n<br />

n=0<br />

N ,<br />

N<br />

N<br />

(7.18)<br />

(7.19)<br />


<strong>SECTIONS</strong> ALNG A <strong>MAP</strong> IN MECHANICS 17<br />

Associated to a (N + 1)-ode Γ there is a set X Γ of vector fields [18] on T N M<br />

such that its flow maps Γ into another (N + 1)-ode. Then if Y ∈ X Γ it satisfies<br />

S N ([Γ, Y ]) = 0, or equivalently τ N,N−1∗ ◦ [Γ, Y ] = 0. The fact that all symmetries<br />

of Γ are elements of X Γ makes this set to be very interesting. The local expression<br />

of an element Y in X Γ is<br />

Y = η i<br />

∂<br />

∂q i (0)<br />

+<br />

N∑<br />

d n Γη i<br />

n=1<br />

∂<br />

∂q i (n)<br />

(7.20)<br />

with η i functions on T N M.<br />

Proposition 7.4. There exists a one-to-one correspondence I Γ : X(τ N,0 ) → X Γ<br />

between the set of vector fields along τ N,0 and the set X Γ given by X ↦→ X (N) ◦<br />

γ (N−1) . The inverse of this map is τ N,0∗<br />

∣<br />

∣XΓ<br />

.<br />

Proof. First we prove that X(γ) = X (N) ◦ γ (N−1) is an element of X Γ , i. e., it<br />

satisfies d [Γ,X(γ)] ◦ τ N,N−1 ∗ = 0. Indeed,<br />

d [Γ,X(γ)] ◦ τ ∗ N,N−1 = d Γ ◦ d X(γ) ◦ τ ∗ ∗<br />

N,N−1 − d X(γ) ◦ d Γ ◦ τ N,N−1<br />

= d Γ ◦ γ (N−1)∗ ∗<br />

◦ d X (N) ◦ τ N,N−1<br />

− γ (N−1)∗ ◦ d X (N) ◦ γ ∗ ∗<br />

◦ d T (N) ◦ τ N,N−1<br />

= d Γ ◦ γ (N−1)∗ ◦ τ ∗ 2N,2N−1 ◦ d X (N−1)<br />

(7.21)<br />

− γ (N−1)∗ ◦ d X (N) ◦ γ ∗ ◦ τ N+1,N ∗ ◦ d T (N−1)<br />

= {d Γ ◦ γ (N−2)∗ − γ (N−1)∗ ◦ d T (2N−1)} ◦ d X (N−1)<br />

= 0.<br />

∣<br />

In order to see that the inverse of I Γ is τ ∣XΓ N,0∗ , we must note that two vector<br />

fields Y and Y ′ in X Γ projecting on the same vector field along τ N,0 , τ N,0∗ (Y −Y ′ ) =<br />

0, are equal as the coordinate expression (7.20) shows. Then if Y ∈ X Γ and we<br />

take the vector field X = τ N,0∗ ◦ Y along τ N,0 we have that X(γ) ∈ X Γ and<br />

τ N,0∗ ◦ (Y − X(γ)) = 0 what implies X(γ) = Y .<br />

□<br />

8. Lagrangian Mechanics<br />

We have seen in section 3 that the basic canonical objects on both the tangent<br />

and cotangent bundle can be redefined as sections along maps. Once a Lagrangian<br />

function L ∈ C ∞ (T M) is considered, new objects associated to L are defined, as<br />

the Poincaré-Cartan 1-form θ L = S ∗ (dL), the Legendre map FL : T M → T ∗ M<br />

or the time evolution operator [2] K L : C ∞ (T ∗ M) → C ∞ (T M), They can also be<br />

redefined as sections along maps. For example, the 1-form θ L is semibasic, so that<br />

it can be seen as a 1-form along τ, θ L ∈ ∧ (τ), θ L = ∂L<br />

∂v<br />

(dq i ◦ τ), and then, it is<br />

i<br />

identified with the Legendre map<br />

FL≡θ L<br />

T ∗ M<br />

<br />

T M M<br />

τ<br />

π<br />

(8.1)


18 JOSÉ F. CARIÑENA⋆ , CARLOS LÓPEZ† , AND EDUARDO MARTÍNEZ⋆<br />

With L and the isomorphism Ψ : T T ∗ M → T ∗ T M we define the time evolution<br />

operator as a vector field along FL [11], K = Ψ −1 ◦ dL<br />

K<br />

T T ∗ M<br />

<br />

T M T ∗ M<br />

FL<br />

τ T ∗ M<br />

(8.2)<br />

and, as a generalization of the vertical endomorphism S, a (1,1) tensor field along<br />

FL⊗ id can also be defined [3]<br />

T T ∗ M<br />

R L<br />

<br />

T M T ∗ M⊗ M T M<br />

FL⊗ id<br />

π T ∗ M ⊗τ T M<br />

(8.3)<br />

given by R L (v) = (dq i ◦ FL)(v)⊗∂ v i. These redefinitions make some relations<br />

trivial, for example, FL ∗ (θ o ) = θ L , taking into account that ˇθ 0 ≡ id T ∗ M and<br />

FL ≡ ˇθ L .<br />

In higher order Mechanics, the Lagrangian is a function on T k M so that (assuming<br />

that L is regular) the Euler-Lagrange equations define in a unique way a<br />

2k-th order ordinary differential equation. More precisely, the Lagrangian defines<br />

the Cartan form θ (k)<br />

L<br />

= S(k) (dL) wich is a τ 2k−1,k−1 -semibasic 1-form, and the<br />

Euler-Lagrange form δL = E (k) (dL) wich is a τ 2k,0 -semibasic 1-form. They can<br />

(k)<br />

be considered as forms along τ 2k−1,k−1 and τ 2k,0 , respectively, and then ˇθ<br />

L<br />

coincides<br />

whith the Legendre transformation FL: T 2k−1 M → T ∗ (T k−1 M) defined<br />

by L. Since E (k) = τ ∗ 2k,k − d T (2k−1) ◦ S (k) , the relation between θ (k)<br />

L<br />

and δL is<br />

δL = τ ∗ 2k,k dL − d T (2k−1)θ (k)<br />

L<br />

. The Euler-Lagrange equations are the (2k)-ode γ<br />

such that<br />

γ ∗ δL = 0, (8.4)<br />

or equivalently, if Γ is the vector field associated to γ,<br />

d Γ θ (k)<br />

L − τ 2k−1,k ∗ dL = 0. (8.5)<br />

ˇθ<br />

(k)<br />

If we define ω (k)<br />

L<br />

= −dθ(k) L<br />

and E L =<br />

L<br />

(T (k−1) ) − τ ∗ 2k−1,k L then it is easy to<br />

prove that Γ is the Hamiltonian vector field of the Hamiltonian dynamical system<br />

(T 2k−1 M, ω (k)<br />

L<br />

, E L). It is also easy to see that i T (2k−1)ω (k)<br />

L<br />

− τ 2k,2k−1 ∗ dE L = −δL.<br />

Since solutions of γ are solutions of γ (l) we can deduce<br />

γ (l)∗ δL (l) = 0 (8.6)<br />

where δL (l) = d l T (2k) δL. Indeed, from (7.10) for r = 1 and N = 2k − 1 we have<br />

γ (l)∗ d l T (2k) δL = d l Γ ◦ γ∗ δL = 0.<br />

The following proposition gives us the variations of the Lagrangian and the<br />

Cartan form under the prolongation of a vector field along τ 2k−1,0 .<br />

Proposition 8.1. Let X be a vector field along τ 2k−1,0 , X ∈ X(τ 2k−1,0 ). Then the<br />

following properties hold:<br />

d X (k)L = d T (3k−2)(ˇθ L (X (k−1) )) + τ 3k−1,2k ⋆ [(δL) ∨ (X)], (8.7)


<strong>SECTIONS</strong> ALNG A <strong>MAP</strong> IN MECHANICS 19<br />

+<br />

3k−2<br />

∑<br />

n,m=0<br />

τ 8k−3,4k−2 ∗ d X (2k−1)θ (k)<br />

L<br />

(−1) n 1 m<br />

(n + 1)!(<br />

n<br />

= τ 8k−3,6k−3 ∗ θ (3k−1)<br />

d X (k) L +<br />

)<br />

τ 8k−3,4k+n−1 ∗ [X (2k+n) , S n+1<br />

2k+n ](δL(n) ).<br />

(8.8)<br />

Proof. Since δL = τ ∗ 2k,k dL−d T (2k−1)θ (k)<br />

L<br />

and taking into account (4.12) and (5.12)<br />

we have<br />

i X (2k)δL = i X (2k)[τ ∗ 2k,k dL − d T (2k−1)θ (k)<br />

L ]<br />

= τ ⋆ 4k−1,3k−1 d X (k)L − i X (2k)d T (2k−1)θ (k)<br />

(8.9)<br />

L<br />

= τ ⋆ 4k−1,3k−1 d X (k)L − d T (4k−2)i X (2k−1)θ (k)<br />

L .<br />

Now, since δL is semibasic over M and θ L is semibasic over T k−1 M, we can put<br />

and<br />

i X (2k)δL = τ 4k−1,2k ⋆ (δL) ∨ (X) (8.10)<br />

i X (2k−1)θ L = τ 4k−2,3k−2<br />

⋆ ˇθL (X (k−1) ), (8.11)<br />

so that<br />

{ }<br />

⋆<br />

τ 4k−1,3k−1 d X (k)L − d T (3k−2)(ˇθ L (X (k−1) )) − τ ⋆ 3k−1,2k [(δL) ∨ (X)] = 0. (8.12)<br />

and since τ 4k−1,3k−1 is a submersion we get the abovementioned result.<br />

In order to prove (8.8) we apply τ 8k−3,6k−3 ∗ [X (k) , S (k) ] = [X (2k) , S (2k) ] ◦ E (k) to<br />

dL, and so we obtain<br />

τ 8k−3,4k−2 ∗ d X (2k−1)θ (k)<br />

L<br />

= τ 8k−3,6k−3 ∗ θ (3k−1)<br />

d X (k) L + [X(2k) , S (2k) ](δL). (8.13)<br />

Now using the expression (6.17) for [X (k) , S (k) ] we get (8.8).<br />

Lemma 8.2. Let L be a regular Lagrangian and G a constant of the motion for<br />

the dynamics defined by L. Then there exists a vector field X along τ 2k−1,0 such<br />

that<br />

d T (2k−1)G = −δL ∨ (X). (8.14)<br />

Proof. Since L is regular, there exists a vector field Y on T 2k−1 M such that<br />

i Y ω (k)<br />

L<br />

= −dG. Moreover, Y is a symmetry of the Hamiltonian dynamical system<br />

(T 2k−1 M, ω (k)<br />

L<br />

, E L). Choosing a vector field Ȳ on T 2k M projecting on Y and<br />

contracting dG + i Y ω (k)<br />

L<br />

= 0 with T (2k−1) we have<br />

d T (2k−1)G + i T (2k−1)i Y ω (k)<br />

L =<br />

= d T (2k−1)G − i Ȳ i T (2k−1)ω (k)<br />

L<br />

= d T (2k−1)G − i Ȳ {i T (2k−1)ω (k)<br />

L<br />

− τ 2k,2k−1 ∗ dE L } − i Ȳ τ ∗ 2k,2k−1 dE L<br />

= d T (2k−1)G + i Ȳ δL − τ ∗ 2k,2k−1 d Y E L<br />

= d T (2k−1)G + i Ȳ δL,<br />

□<br />

(8.15)<br />

where we have used d Y E L = 0. Now defining X = τ 2k−1,0∗ ◦ Y , since δL is τ 2k,0 -<br />

semibasic we obtain d T (2k−1)G = −δL ∨ (X).<br />

□<br />

Definition 8.3. Let L be a Lagrangian function in T k M and X a vector field<br />

along τ 2k−1,0 . We say that X is a symmetry of the Lagrangian [5] if there exists a<br />

function F ∈ C ∞ (T 3k−2 M) such that d X (k)L = d T (3k−2)F .


20 JOSÉ F. CARIÑENA⋆ , CARLOS LÓPEZ† , AND EDUARDO MARTÍNEZ⋆<br />

Now, we prove the following generalized version of Noether’s Theorem for higherorder<br />

Lagrangians that establishes a one-to-one correspondence between symmetries<br />

of the Lagangian function and first integrals of the dynamics defined by this Lagrangian.<br />

Theorem 8.4. If X is a symmetry of the Lagrangian and F is the associated funtion<br />

to X, then there exists a function G ∈ C ∞ (T 2k−1 M) such that τ ⋆ 3k−2,2k−1 G =<br />

F − ˇθ L (X (k−1) ) and it is a constant of the motion. Conversely, given a constant<br />

of the motion G ∈ C ∞ (T 2k−1 M) there exist a vector field X along τ 2k−1,0 and a<br />

function F = τ ⋆ 3k−2,2k−1 G + ˇθ L (X (k−1) ) in T 3k−2 M such that X is a symmetry of<br />

the Lagrangian with associated function F . Moreover, X(γ) is a symmetry of the<br />

Hamiltonian dynamical system (T 2k−1 M, ω (k)<br />

L<br />

, E L).<br />

Proof. Let us suppose that X ∈ X(τ 2k−1,0 ) is a symmetry of the Lagrangian, namely<br />

Then, from (8.7) we have<br />

d X (k)L = d T (3k−2)F. (8.16)<br />

d T (3k−2)[F − ˇθ L (X (k−1) )] = −τ 3k−1,2k ⋆ [(δL) ∨ (X)] (8.17)<br />

so that, there exists a funtion G ∈ C ∞ (T 2k−1 M) such that τ ⋆ 3k−2,2k−1 G = F −<br />

ˇθ L (X (k−1) ), i. e., d T (2k−1)G = −(δL) ∨ (X), and composing it with γ ⋆ we have that<br />

d Γ G = 0, i. e., G is a constant of the motion.<br />

Conversely, let G ∈ C ∞ (T 2k−1 M) be a constant of the motion and X the vector<br />

field along τ 2k−1,0 such that d T (2k−1)G = −(δL) ∨ (X). Then defining F =<br />

ˇθ (k)<br />

L (X(k−1) ) + τ ⋆ 3k−2,2k−1 G, the equation (8.7) reads d X (k)L = d T (3k−2)F .<br />

Now, we prove that X(γ) is a symmetry of (T 2k−1 M, dθ (k)<br />

L<br />

account (7.8), (8.8) and θ (3k−1)<br />

d T (3k−2) F = τ 6k−2,3k−2 ∗ dF we have<br />

d X(γ) θ (k)<br />

L = d(γ(k−2)∗ F )+<br />

+<br />

3k−2<br />

∑<br />

m,n=0<br />

(−1) n 1 m<br />

(n + 1)!(<br />

n<br />

, E L). Taking into<br />

)<br />

γ (2k+n−1)∗ [X (2k+n) , S n+1<br />

2k+n ](δL(n) ), (8.18)<br />

and since [X (2k+n) , S n−1<br />

2k+n<br />

] are ‘linear’ operators the last term vanishes, so that<br />

d X(γ) θ (k)<br />

L<br />

= d(γ(k−2)∗ F ) and d X(γ) ω (k)<br />

L = 0. □<br />

Note that i X(γ) ω (k)<br />

L<br />

= −d(γ(k−2)∗ F − θ (k)<br />

L<br />

(X(γ)) and since the function F −<br />

ˇθ L (X (k−1) ) is the pull-back τ ⋆ 3k−2,2k−1 G we obtain i X(γ) ω (k)<br />

L<br />

= −dG.<br />

Note also that the equation τ ⋆ 3k−2,2k−1 G = F − ˇθ L (X (k−1) ) imposes severes<br />

restrictions to the form of the function F . For instance F must be a polynomial in<br />

the coordinates q(n) i for n = 2k, . . . , 3k − 2.<br />

Finally we can see why the approach by Marmo and Mukunda [14] is not generalizable<br />

to higher-order Mechanics. The pull-back of the equation d X (k)L = d T (3k−2)F<br />

to T 4k−2 M is d X (2k−1)(τ ∗ 2k−1,k L) = τ 4k−2,3k−1 d T (3k−2)F . Now, if D is 2k-ode and<br />

δ is the corresponding section we have<br />

d X(δ) (τ 2k−1,k ∗ L) = δ (2k−2)∗ τ 4k−2,3k−1 d T (3k−2)F<br />

= δ (k−1)∗ d T (3k−2)F<br />

= d D (δ (k−2)∗ F ).<br />

(8.19)


<strong>SECTIONS</strong> ALNG A <strong>MAP</strong> IN MECHANICS 21<br />

It is easy to see that X is a symmetry of the Lagrangian if and only if we have that<br />

d X(δ) (τ 2k−1,k ∗ L) = d D (δ (k−2)∗ F ) for every 2k-ode D. In first order theories, k=1,<br />

this equation reduces to d X(δ) L = d D F , wich is the condition imposed by Marmo<br />

and Mukunda. Nevertheles in higher order theories, the function that appears on<br />

the right hand side is δ (k−2)∗ F , wich depends on the ode D.<br />

References<br />

[1] Bourbaki N, Varietés differentielles et analytiques (Fasc. XXXVI), Hermann, París, 1971<br />

[2] Cariñena J F and López C, The time evolution operator for singular Lagrangians, Lett.<br />

Math. Phys. 14 (1987) 203.<br />

[3] Cariñena J F, López C and Román-Roy N, Geometric study of the connection between<br />

the Lagrangian and Hamiltonian constraints J. Geom. Phys. 4 (1986) 315.<br />

[4] Cariñena J F, López C and Martínez E, A geometric characterization of Lagrangian second<br />

order differential equations, Inverse Problems 5 (1989) 691.<br />

[5] Cariñena J F, López C and Martínez E, A new approach to the converse of Noether’s<br />

Theorem, J. Phys. A: Math. Gen. 22 (1989) 4777.<br />

[6] Cantrijn F, Crampin M, Sarlet W and Saunders D, The canonical isomorphism between<br />

T k T ∗ M and T ∗ T k M C. R. Acad. Sci. París 309 II (1989) 1509.<br />

[7] Crampin M, On the differential geometry of the Euler-Lagrange equations and the inverse<br />

problem of Lagrangian dynamics, J. Phys. A: Math. Gen. 14 (1981) 2567.<br />

[8] Crampin M, Sarlet W and Cantrijn F, Higher order differential equations and higher<br />

order Lagrangian mechanics, Math. Proc. Camb. Phil. Soc. 99 (1986) 565.<br />

[9] De León M and Rodrigues P, Generalized Classical Mechanics and Field Theory, North-<br />

Holland Math. Stu. 112, Amsterdam, 1985.<br />

[10] Frölicher A and Nijenhuis A, Theory of vector valued differential forms, Nederl. Akad.<br />

Wetensch. Proc.A59 (1956) 338.<br />

[11] Gràcia X and Pons J M, On an evolution operator connecting Lagrangian and Hamiltonian<br />

formalisms, Lett. Math. Phys. 17 (1989) 175.<br />

[12] Kosmann-Schwarzbach Y, Vector fields and generalized vector fields on fibered manifolds,<br />

Lecture Notes in Mathematics 792, p. 307, Springer, New York, 1980.<br />

[13] Kupershmidt B A, Geometry of jet bundles and the structure of Lagrangian and Hamiltonian<br />

formalisms, Lecture Notes in Mathematics 775, p. 162, Springer, New York, 1980.<br />

[14] Marmo G and Mukunda N, Symmetries and constants of the motion in the Lagrangian<br />

formalism on T Q: beyond point transformations, Nuovo Cim. 92 B (1986) 1.<br />

[15] Olver P, Applications of Lie groups to differential equations, Springer, New York, 1986<br />

[16] Pidello G and Tulczyjew W M, Derivations of Differential Forms on Jet Bundles, Annali<br />

di Mathematica Pura ed Aplicata 147 (1987) 249.<br />

[17] Poor W A, Differential Geometric Structures, Mc- Graw-Hill, 1981.<br />

[18] Sarlet W, In Proc. Conf. on Differential Geometry and Applications, Brno 1986, Reidel,<br />

1987, p. 279.<br />

[19] Saunders D J The Geometry of Jet Bundles, Cambridge University Press, 1989<br />

[20] Tulczyjew W M, An intrinsic geometric construction for the Lagrange Derivative, Bull.<br />

Acad. Pol. Sci. XXIII 9 (1975).<br />

⋆ Departamento de Física Teórica, Universidad de Zaragoza, 50009 Zaragoza (Spain)<br />

† Departamento de Física Teórica II, Universidad Complutense de Madrid, 28040 Madrid<br />

(Spain)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!