Sei sulla pagina 1di 360

Gauge

Mechanics
This page is intentionally left blank
Gauge
Mechanics
L. Mangiarotti
University of Camerino, Italy
G. Sardanashvily
Moscow State University, Russia
World Scientific
Singapore *New Jersey London 'HongKong
Published by
World Scientific Publishing Co. Pte. Ltd
P O Box 128, Farcer Road, Singapore 912805
USA office: Suite IB, 1060 Main Street, River Edge, NJ 07661
UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE
British Library Cataloguing-in-Publication Data
A catalogue record for this book is available from the British Library.
GAUGE MECHANICS
Copyright 1998 by World Scientific Publishing Co Pte. Ltd
All rights reserved. This book, or parts thereof, may not he reproduced in any form or by any means,
electronic or mechanical, including photocopying, recording or any information storage and retrieval
system now known or to be invented, without written permission from the Publisher.
For photocopying of material in this volume, please pay a copying fee through the Copyright
Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to
photocopy is not required from the publisher
ISBN 981-02-3603-4
This book is printed on acid-free paper.
Printed in Singapore by Uto-Print
Preface
This book presents in a unified way modern geometric methods in analytical
mechanics, based on the application of jet manifolds and connections. As is well
known, the technique of Poisson and symplectic spaces provide the adequate Hamil-
tonian formulation of conservative mechanics. This formulation, however, cannot be
extended to time-dependent mechanics subject to time-dependent transformations.
We will formulate non-relativistic time-dependent mechanics as a particular field
theory on fibre bundles over a time axis.
The geometric approach to field theory is based on the identification of classical
fields with sections of fibred manifolds. J et manifolds provide the adequate mathe-
matical language for Lagrangian field theory, while the Hamiltonian one is phrased
in terms of a polysymplectic structure. The 1-dimensional reduction of Lagrangian
field theory leads us in a straightforward manner to Lagrangian time-dependent
mechanics. At the same time, the canonical polysymplectic form on a momentum
phase space of time-dependent mechanics reduces to the canonical exterior 3-form
which plays the role similar to a symplectic form in conservative mechanics. With
this canonical 3-form, we introduce the canonical Poisson structure and formulate
Hamiltonian time-dependent mechanics in terms of Hamiltonian connections and
Hamiltonian forms.
Note that the theory of non-linear differential operators and the calculus of vari-
ations are conventionally phrased in terms of jet manifolds. On the other hand, jet
formalism provides the contemporary language of differential geometry to deal with
non-linear connections, represented by sections of jet bundles. Only jet spaces enable
us to treat connections, Lagrangian and Hamiltonian dynamics simultaneously.
In fact, the concept of connection is the main link throughout the book. Con-
nections on a configuration space of time-dependent mechanics are reference frames.
Holonomic connections on a velocity phase space define non-relativistic dynamic
equations which are also related to other types of connections, and can be writ-
ten as non-relativistic geodesic equations. Hamiltonian time-dependent mechanics
deals with Hamiltonian connections whose geodesies are solutions of the Hamilton
v
PREFACE
vi
equations.
The presence of a reference frame, expressed in terms of connections, is the main
peculiarity of time-dependent mechanics. In particular, each reference frame defines
an energy function, and quantizations with respect to different reference frames are
not equivalent.
Another important peculiarity is that a Hamiltonian fails to be a scalar function
under time-dependent transformations. As a consequence, many well-known con-
structions of conservative mechanics fail to be valid for time-dependent mechanics,
and one should follow methods of field theory.
At the same time, there is the essential difference between field theory and time-
dependent mechanics. In contrast with gauge potentials in field theory, connections
on a configuration space of time-dependent mechanics fail to be dynamic variables
since their curvature vanishes identically. Following geometric methods of field the-
ory, we obtain the frame-covariant formulation of time-dependent mechanics. By
analogy with gauge field theory, one may speak about gauge time-dependent me-
chanics.
In comparison with non-relativistic time-dependent mechanics, a configuration
space of relativistic mechanics does not imply any preferable fibration over a time.
To construct the velocity phase space of relativistic mechanics, we therefore use
formalism of jets of submanifolds. At the same time, Hamiltonian relativistic me-
chanics is seen as an autonomous Hamiltonian system on the constraint space of
relativistic hyperboloids.
With respect to mathematical prerequisites, the reader is expected to be familiar
with the basics of differential geometry of fibre bundles. For the convenience of the
reader, several mathematical facts and notions are included as an Interlude, thus
making our exposition self-contained.
Contents
Preface v
Introduction 1
1 Interlude: bundl es, J ets, Connecti ons 9
1.1 Fibre bundles 9
1.2 Multivector fields and differential forms 20
1.3 J et manifolds 35
1.4 Connections 42
1.5 Bundles with symmetries 46
1.6 Composite fibre bundles 53
2 Geomet ry of Poi sson Manifolds 57
2.1 J acobi structure 57
2.2 Contact structure 61
2.3 Poisson structure 66
2.4 Symplectic structure 73
2.5 Presymplectic structure 81
2.6 Reduction of symplectic and Poisson structures 84
2.7 Appendi x. Poisson homology and cohomology 89
2.8 Appendi x. More brackets 96
2.9 Appendi x. Multisymplectic structures 100
3 Hami l toni an Systems 105
3.1 Dynamic equations 106
3.2 Poisson Hamiltonian systems I l l
3.3 Symplectic Hamiltonian systems 114
3.4 Presymplectic Hamiltonian systems 118
3.5 Dirac Hamiltonian systems 123
3.6 Dirac constraint systems 129
vii
viii
CONTENTS
3.7 Hamiltonian systems with symmetries 133
3.8 Appendi x. Hamiltonian field theory 139
4 Lagrangian time-dependent mechanics 151
4.1 Fibre bundles over K 152
4.2 Dynamic equations 159
4.3 Dynamic connections 162
4.4 Non-relativistic geodesic equations 170
4.5 Reference frames 175
4.6 Free motion equations 179
4.7 Relative acceleration 182
4.8 Lagrangian systems 186
4.9 Newtonian systems 195
4.10 Holonomic constraints 206
4.11 Non-holonomic constraints 211
4.12 Lagrangian conservation laws 218
5 Hamiltonian time-dependent mechanics 227
5.1 Canonical Poisson structure 228
5.2 Hamiltonian connections and Hamiltonian forms 231
5.3 Canonical transformations 242
5.4 The evolution equation 247
5.5 Degenerate systems 248
5.6 Quadratic degenerate systems 262
5.7 Hamiltonian conservation laws 269
5.8 Time-dependent systems with symmetries 271
5.9 Systems with time-dependent parameters 274
5.10 Unified Lagrangian and Hamiltonian formalism 282
5.11 Vertical extension of Hamiltonian formalism 285
5.12 Appendi x. Time-reparametrized mechanics 296
6 Relativistic mechanics 299
6.1 J ets of submanifolds 299
6.2 Relativistic velocity and momentum phase spaces 303
6.3 Relativistic dynamics 307
6.4 Relativistic geodesic equations 311
CONTENTS
ix
Appendix A. Geometry of BRST mechanics 317
Appendix B. On quantum time-dependent mechanics 327
Bibliography 332
I ndex 347
Introducti on
The present book deals with first order mechanical systems, governed by the sec-
ond order differential equations in coordinates or the first order ones in coordinates
and momenta. Our goal is the description of non-conservative mechanical systems
subject to time-dependent transformations, including inertial and non-inertial frame
transformations and phase transformations.
Symplectic technique is well known to provide the adequate Hamiltonian for-
mulation of conservative (i.e., time-independent) mechanics where Hamiltonians are
independent of time [2, 6, 72, 116, 126]. The familiar example is a mechanical sys-
tem whose momentum phase space is the cotangent bundle T'M of a configuration
space M. This fibre bundle is provided with the canonical symplectic form
Q = dpi A dq',
(1)
written with respect to the holonomic coordinates (g',p, =<?,) on T'M. A Hamil-
tonian H of a conservative mechanical system is defined as a real function on the
momentum phase space T'M. Then a motion of this system is an integral curve of
the Hamiltonian vector field
d = ^d' + d%
on T'M which fulfills the Hamilton equations
ti\Q = -dH,
& = sht,
di = -diH.
Lagrangian conservative mechanics is usually seen as a particular Hamiltonian me-
chanics on the tangent bundle TM of a configuration space M, which is endowed
with the presymplectic form defined by a Lagrangian.
The Hamiltonian formulation of conservative mechanics cannot be extended in a
straightforward manner to time-dependent mechanics because the symplectic form
1
2
INTRODUCTION
(1) is not invariant under time-dependent transformations, including the inertial
frame transformations.
The existent formulation of time-dependent mechanics implies a preliminary
splitting of a configuration space
Q =1 x M,
(2)
where M is a manifold, while 1 is a time axis (see [29, 31, 50, 110, 136, 146, 166] and
references therein). FYom the physical viewpoint, it means that a certain reference
frame is chosen. Then we have the corresponding splitting of the velocity phase
space
R x T M (3)
and that of the momentum phase space
R x T'M. (4)
The momentum phase space (4) is provided with the presymplectic form
prjQ =dp, A dq
x
(5)
which is the pull-back of the canonical symplectic form Q. (1) on the cotangent
bundle T'M [27]. By a time-dependent Hamiltonian H is meant a real function
on the momentum phase space R x T'M, while trajectories of motion are integral
curves of the time-dependent vector field
d : R x T'M -> TT'M
which satisfies the Hamilton equations
0" = d'H, A = -diH.
The problem is that the splittings (2) - (4) are broken by any time-dependent
transformation, and so is the presymplectic form (5). Therefore the familiar methods
of conservative mechanics and their extensions to the product spaces (2) - (4) fail
to be valid for mechanical systems subject to time-dependent transformations.
We will formulate non-relativistic time-dependent mechanics as a particular field
theory whose configuration space is a fibred manifold over a time axis R [14, 55, 57,
106, 114, 132, 159, 161].
INTRODUCTION 3
Geometric formalism of field theory is based on the identification of classical
fields with sections of a fibred manifold Y >X. The corresponding velocity phase
space is the first order jet manifold J
1
Y of sections of Y * X, while the momentum
one is the Legendre bundle
U = VYCAT*X), n =dim X,
(6)
over Y [28, 56, 57, 73, 96, 158, 159].
In the case of X = R of time-dependent mechanics, its configuration space is a
fibred manifold
7T : Q R,
equipped with fibred coordinates (t,q
l
). The base R is parameterized by the Carte-
sian coordinates t with the transition functions t' = +const. Relative to these
coordinates, the time axis R is provided with the standard vector field d
t
and the
standard 1-form dt which is also the volume element on R. Of course, this is not
the case of relativistic mechanics (see Chapter 6) nor of the models with a time
reparametrization (see Section 5.12).
The velocity phase space of non-relativistic time-dependent mechanics is the first
order jet manifold J
l
Q of sections of the fibred manifold Q R. It is equipped
with the adapted coordinates (t,q
%
,q\). There is the canonical imbedding
X:J
l
Q^TQ,
(7)
A =d
t
+ q\d
u
of the velocity phase space J
X
Q into the tangent bundle TQ of the configuration
space Q. From now on we will identify J
1
Q with its image in TQ given by the
coordinate conditions
t = l,
<? = <? ;
(8)
This is an affine subbundle of TQ * Q modelled over the vertical tangent bundle
VQ ->Q of the fibred manifold Q R.
The morphism (7) plays a prominent role in our formulation of time-dependent
mechanics. It enables us to treat the jet manifold J
l
Q as a velocity phase space
4
INTRODUCTION
of a mechanical system. Due to this morphism, every connection T on the fibred
manifold Q ->R can be identified with the nowhere vanishing vector field
r : Q - J
l
Q C TQ,
(9)
r = a
t
+ r%,
on Q which is the horizontal lift of the standard vector field d
t
on R by means
of this connection T. We will continue to call (9) a connection in order to refer
to the standard properties of connections without additional explanation. From
the physical viewpoint, a connection (9) sets a tangent vector at each point of the
configuration space Q, which characterizes the velocity of an "observer" at this
point. It follows that a connection T on the fibred manifold Q R defines a
reference frame [57, 132, 161]. In particular, one can think of the difference q\ - P
as being the relative velocity with respect to the reference frame T, whereas the
notion of a relative acceleration is more intricate (see Section 4.7).
The momentum phase space of non-relativistic time-dependent mechanics is the
Legendre bundle (6) where X = R. This phase space is isomorphic to the vertical
cotangent bundle II =V'Q of the fibred manifold Q R, and is equipped with the
holonomic coordinates (t,q
l
,p, = <ft). It should be emphasized that this is not the
most general case of a momentum phase space of time-dependent mechanics, which
is defined as a fibred manifold II R provided with a Poisson structure such that
the corresponding symplectic foliation belongs to the fibration II R [74]. In fact,
putting n = V'Q, we restrict our consideration to Hamiltonian systems which have
the Lagrangian counterparts.
Note that Lagrangian and Hamiltonian formalisms are equivalent only if a La-
grangian is hyperregular, i.e., the Legendre map from the velocity phase space to the
momentum one is a diffeomorphism. In general, a degenerate Lagrangian involves a
set of associated Hamiltonians in order to exhaust solutions of the Lagrange equa-
tions (see Section 5.5). Nevertheless, there are physically interesting systems whose
phase spaces fail to be the cotangent bundles of configuration spaces, and they do
not admit any Lagrangian description [168]. The unified Lagrangian-Hamiltonian
formalism of the joint velocity-momentum phase space
n =V V ' QS J
l
V"Q
enables us to relate a Lagrangian system to any Hamiltonian one (see Section 5.10).
INTRODUCTION 5
Let us turn to the momentum phase space V'Q of time-dependent mechanics.
It is endowed with the canonical exterior 3-form
il = dpi A dq
x
A dt (10)
which is the particular case of the canonical polysymplectic form on the Legendre
bundle (6), when X = R [57, 161]. The exterior form (10) is invariant under all
holonomic transformations of the momentum phase space V'Q.
In time-dependent mechanics, the canonical 3-form $7 (10) plays a role similar to
the canonical symplectic form (1) in conservative symplectic mechanics. The form
(10) yields the canonical Poisson structure on the momentum phase space V'Q,
and provides the Hamiltonian formulation of time-dependent mechanics in terms
of Hamiltonian connections and Hamiltonian forms. This formulation is compatible
with the Lagrangian formulation of time-dependent mechanics on the velocity phase
space J
l
Q, and is equivalent to the Lagrangian one in the case of hyperregular
Lagrangians.
The following peculiarities of Hamiltonian time-dependent mechanics should be
emphasized.
The canonical Poisson structure defined by the 3-form fl (10) on the momen-
tum phase space V'Q of time-dependent mechanics is degenerate.
A Hamiltonian on a momentum phase space of time-dependent mechanics fails
to be a scalar function, but reads
H =p,P + H
T
,
(11)
where T is a connection on the fibred manifold Q IR, while Hr is a Hamil-
tonian function which is also an energy density with respect to the reference
frame T.
A Poisson bracket of a Hamiltonian (11) with functions on a momentum phase
space is defined only locally. Being equal to zero with respect to some coordi-
nates, it does not necessarily vanish with respect to other ones.
As a consequence, the evolution equation in time-dependent mechanics is not
reduced to a Poisson bracket, and integrals of motion are not functions in invo-
lution with a Hamiltonian. For the same reason, the familiar Dirac-Bergmann
6
INTRODUCTION
algorithm for describing constraint systems is not extended to time-dependent
mechanics.
Quantizations with respect to different reference frames are not equivalent.
For the sake of simplicity, throughout the book, a configuration space of time-
dependent mechanics Q >R is assumed to be a fibre bundle with a typical fibre
M. We will call it a configuration bundle. It is always trivial. Note that, although
such a configuration space Q is diffeomorphic to a direct product R x M, in general,
it cannot be canonically identified to R x M. Its different trivializations
ip: Q ^R x M
(12)
differ from each other in fibrations Q >M, while the fibration Q >R is once for
all. Given a trivialization (12) of a configuration space, there are the corresponding
splittings of velocity and momentum phase spaces
J
l
Q ^R x TM,
V'Q =R x T'M.
We show that every trivialization (12) of the fibre bundle Q R defines a (complete)
connection (9) on this fibre bundle, and vice versa. Consequently, every such a
trivialization corresponds to a (complete) reference frame.
Note that a reference frame is one of the main ingredients in our picture of
time-dependent mechanics. For instance, each reference frame defines an energy
function. Following geometric methods of field theory, we obtain the formulation of
time-dependent mechanics which involves necessarily connections T on a configura-
tion bundle Q R, and thus is covariant under reference frame transformations.
By analogy with the gauge field theory, one may speak about gauge time-dependent
mechanics, though the term "gauge mechanics" also stands for mechanics of de-
formable bodies [120] (see Section 5.11).
At the same time, there is an essential difference between field theory and time-
dependent mechanics. Since a configuration space of time-dependent mechanics is a
fibre bundle Q R over a 1-dimensional base, the curvature of any connection on a
configuration bundle vanishes identically. In contrast with gauge potentials of gauge
theories, these connections fail to be dynamic quantities because they can always be
brought into the standard vector field r = d
t
by time-dependent transformations.
INTRODUCTION 7
Therefore, Lagrangians in time-dependent mechanics are covariant, but not invariant
under reference frame transformations.
Connections play a prominent role in our formulation of time-dependent mechan-
ics. As was mentioned above, connections on a configuration bundle Q R describe
non-relativistic reference frames. Holonomic connections on the jet bundle J
l
Q K
define non-relativistic dynamic equation which, in turn, are associated with connec-
tions on the affine jet bundle J
l
Q >Q and the tangent bundle TQ >Q. As a
result, every non-relativistic dynamic equation can be seen as a geodesic equation on
the tangent bundle TQ Q that furnishes the relationship between non-relativistic
and relativistic dynamics (see Section 6.4). Hamiltonian time-dependent mechanics
deals with Hamiltonian connections whose geodesies are solutions of the Hamilton
equations.
In comparison with non-relativistic mechanics, if a configuration space of a me-
chanical system has no preferable fibration Q >R, we obtain the general formula-
tion of relativistic mechanics, including Special Relativity on the Minkowski space
Q = R
4
. The velocity phase space of relativistic mechanics is the first order jet
manifold J\Q of 1-dimensional submanifolds of the configuration space Q [57, 161].
This notion of jets generalizes that of jets of sections of fibre bundles which we have
utilized in field theory and non-relativistic mechanics. The jet bundle J\Q Q is
projective, and one can think of its fibres as being spaces of the 3-velocities of a rela-
tivistic system. The 4-velocities of a relativistic system are represented by elements
of the tangent bundle TQ of the configuration space Q, while the cotangent bundle
T'Q, endowed with the canonical symplectic form, plays the role of the momentum
phase space of relativistic theory. As a result, Hamiltonian relativistic mechanics
can be seen as a constraint Dirac system on the hyperboloids of relativistic momenta
in the momentum phase space T'Q.
Formalism of jets of submanifolds provides the common description of non-
relativistic mechanics and relativistic theory. In particular, the tangent bundle TQ
of a configuration space Q plays the role of the space of the 4-velocities both in non-
relativistic and relativistic mechanics. The difference is only that, given a fibration
Q >R, the 4-velocities of a non-relativistic system live in the subbundle (8) of TQ,
whereas the 4-velocities of a relativistic theory belong to the hyperboloids
g^ifq" =1,
(13)
where g is an admissible pseudo-Riemannian metric in TQ. Moreover, as was men-
tioned above, both relativistic and non-relativistic equations of motion can be seen
8 INTRODUCTION
as geodesic equations on the tangent bundle TQ, but their solutions live in its dif-
ferent subbundles (13) and (8).
Unless otherwise stated, we believe that all quantities are physically dimension-
less. Following field theory, we will sometimes refer to the universal unit system
where the velocity of light c and the Planck constant h are equal to 1, while the
length unit is the Planck one
(Gftc-
3
)
1/2
=G
1/2
= 1.616-10-
M
cm,
where G is the Newtonian gravitational constant. Relative to the universal unit
system, the physical dimension of the spatial and temporal Cartesian coordinates
is the [length], the physical dimension of a mass is the [length]
-1
, while an action
functional and a metric tensor are physically dimensionless.
Chapter 1
Interlude: bundles, jets,
connecti ons
This Chapter does not claim to be a survey on modern differential geometry. The
relevant material is presented in a fairly informal way. For details, we refer the
reader to [57, 100, 157, 164, 170, 185].
Throughout the book, all maps are smooth, i.e., of class C, while manifolds
are real, finite-dimensional, second-countable and, hence, paracompact. Unless oth-
erwise stated, we assume that manifolds are connected.
We use the standard symbols , V, and A for the tensor, symmetric, and exterior
products, respectively. The interior product (contraction) of vectors and forms is
denoted by J . By dg are meant the partial derivatives with respect to the coordinates
with indices g. The symbol o stands for a composition of maps.
1.1 Fibre bundles
Subsections: Fibre bundles, 9; Vector bundles, 12; A/fine bundies, 14; Tangent and
cotangent bundles, 15; Tangent and cotangent bundles of fibre bundles, 16; Sheaves,
18.
Fibre bundl es
By a fibre bundle is meant a locally trivial fibred manifold
K:Y ->X (1.1.1)
9
10
CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
where a fibration (or a projection) n is a surjective submersion from a manifold Y,
called a totai space, onto a base X. Unless otherwise stated, we put di mX =n. By
definition, a base X of a fibre bundle (1.1.1) admits an open covering {/J so that
Y is locally isomorphic to the splittings
^ : T T
1
^) - U
(
x V,
called local bundJ e trivializations, together with the transition functions
P
K
{u
i
n tr
c
) x v -(t/j n p
c
) x v,
Ve() = (p^c)(y).
ye *-'((/<n/
c
),
where V is the typicaj fibre of the fibre bundle (1.1.1). The bundle trivializations
(U(,ilf() constitute an atlas
* = {(t/
t
,V).P}
of a fibre bundle. Given an atlas ty, a fibre bundle Y is provided with the associated
atlas of fibred coordinates (x
x
,y
l
), where
x
x
{y) = (x
x
on){y), yeY,
are coordinates on the base X, and
y'(y) = (y*
o
P^2^)(v)
are coordinates on the typical fibre V.
A fibre bundle Y X is called trivial if V is diffeomorphic to the product X x V.
Different trivializations of a fibre bundle differ from each other in projections Y V
of the total space Y onto the typical fibre V.
THEOREM 1.1.1. [170]. Each fibre bundle over a contractible base is trivial.
By a section (or a global section) of a fibre bundle (111) is meant a manifold
morphism s : X Y such that -n o 5 = I dX. A section s is an imbedding, i.e.,
s(X) C Y is a submanifold of a total space Y which is also a topological subspace
of Y. Similarly, a section s of a fibre bundle Y >X over a submanifold N C X is
a morphism s : N Y, such that
n o s =i
N
: N > Y.
1.1. FIBRE BUNDLES 11
A section of a fibre bundle over an open subset of its base will be called simply a
(local) section. A fibre bundle, by definition, admits a local section over an open
neighbourhood of each point of its base.
THEOREM 1.1.2. [170]. A fibre bundle Y -* X whose typical fibre is diffeomorphic
to R
m
has a global section. A (smooth) section over a closed subset of X can always
be extended to a global section. D
A Bbred morphism of two bundles n : Y X and 7r' : Y' >X' is a pair of maps
$ : Y ->Y' and / : X -* X' such that the diagram
Y -*->r
i I
X >X'
/
(1.1.2)
is commutative, i.e., $ sends fibres to fibres. In brief, we will say that (1.1.2) is a
fibred morphism
$ : Y Y'
!
over /. If f = ldX, then
$ : Y >Y'
x
is called a fibred morphism over X.
Remark 1.1.1. Unless otherwise stated, by the rank of a fibred morphism $ (1.1.2)
over a diffeomorphism / is meant its rank minus di mX.
A fibred morphism $ (1.1.2) over X (or its image $(Y)) is said to be a subbundle
of the fibre bundle Y' ->X if $(Y) is a submanifold of Y'.
We deduce from the implicit function theorem the following useful criteria for
an image and a pre-image of a fibred morphism to be a subbundle [149, 185].
THEOREM 1.1.3. Let # : Y Y' be a fibred morphism over X. Given a global
section s' of the fibre bundle V" >X such that s'(X) C I m$, by the Jcernel of the
fibred morphism $ with respect to the section s' is meant the pre-image
Ker
s
.$ = $-
1
(s'(^))
12 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
of s(X) by $. If $ : Y V" is a fibred morphism of constant rank over X, then
I m$ and Ker
s
/$ are subbundles of Y' and Y, respectively.
An isomorphism of fibre bundles is a fibred morphism (1.1.2) such that $ is a
diffeomorphism. A fibred morphism [isomorphism] of a fibre bundle V - >X to itself
is called an endomorphism (automorphism). An automorphism over I dX is said to
be a vertical automorphism. Following physical terminology, automorphisms of fibre
bundles will also be called gauge transformations.
Given a fibre bundle 7r : Y > X and a manifold map / : X' X, the pull-back
f'Y of Y by / is the fibre bundle over X' whose total space is
f'Y^{(x\y)eX'xY : n(y) = /(*')}
together with the natural projection (x
1
, y) i- x'. Roughly speaking, the fibre of
the pull-back f'Y over a point x' X' is that of Y over the point f(x') e X. If
X' C X is a submanifold of X and i
x
> is the corresponding natural injection, then
the pull-back
i'
x
,Y = Y \
x
,
is called the restriction of a fibre bundle Y to the submanifold X' C X.
Let 7T : Y >X and ir' : V" > X be fibre bundles over the same base X. Then-
fibred product
YxY'
over X is defined as the pull-back
YxY' = n'Y' or V x Y' = Tr'V
A- X
together with the natural projection onto X.
Vector bundles
A vector bundle is a fibre bundle Y * X such that:
its typical fibre V and all the fibres Y
x
= ir '(x), x X, are real finite-
dimensional vector spaces;
LI . FIBRE BUNDLES 13
there is a bundle atlas <P = {(U
a
,ip
a
)} of Y >X whose trivialization mor-
phisms 4>
a
restrict to linear isomorphisms
xP
a
(x) :Y
X
^V,
Vz 6 U
a
.
Dealing with a vector bundle Y, we always use linear bundle coordinates (I
A
,J /')
associated with the above-mentioned bundle atlas 9. We have
(pr
2
o ip
a
)(y) = y'a,
y = y'e,(x) = y
x
ip
a
(x)
l
(e<),
where {e,} is a fixed basis for the typical fibre V of Y, while {e,(z)} is the fibre
basis (or the frame) for the fibre Y
x
of Y, which is associated with the bundle atlas
9.
By virtue of Theorem 1.1.2, vector bundles have global sections, e.g., the global
zero section 0(X). If there is no risk of confusion, we write 0, instead of 0(X).
A morphism of vector bundles #: Y Y' is defined as a fibred morphism over
/ : X X' whose restriction $
x
: Y
x
Y^^ to each fibre of Y is a linear map. It
is called a linear bundle morphism over /.
The following assertion is a corollary of Theorem 1.1.3.
PROPOSI TI ON 1.1.4. If Y X and Y' - X are vector bundles and $ : Y ->V" is a
linear bundle morphism of constant rank over X, then the image of $ and the kernel
Kerjj$ of $ with respect to the zero section 0 of Y' >X are vector subbundles of
Y' * X and Y > X, respectively. Note that a vector subbundle of a vector bundle
is a closed imbedded submanifold.
Unless otherwise stated, by Ker $ of a linear bundle morphism $ is meant its
kernel with respect to the zero section 0.
There are the following standard constructions of new vector bundles from old.
Let Y >X be a vector bundle with a typical fibre V. By Y" X is meant
the dual vector bundle with the typical fibre V, dual of V. The interior
product of Y and Y' is defined as a fibred morphism
\:YY" > X x R .
J
x
14 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
Let Y > X and Y' X be vector bundles with typical fibres V and V,
respectively. Their Whitney sum Y Y' is a vector bundle over X with the
typical fibre V V.
Let V X and Y' X be vector bundles with typical fibres V and V,
respectively. Their tensor product Y Y' is a vector bundle over X with the
typical fibre V V". Similarly, the exterior product Y AY is denned.
In particular, let $ : Y\ * Y
2
be an injection of a vector bundle iT\ : Y\ Xi to
a vector bundle 7r
2
: Y
2
X
2
over a diffeomorphism f : X\ * X
2
. Then, there is
the dual surjection
$: y
2
* y;,
<*(),> =<,*()>, v^eTrrHr'o^^cy,, u e y
2
-, (1.1.3)
over the diffeomorphism / '.
AfRne bundles
Let W : y * X be a vector bundle with a typical fibre V. An afSne bundle
modelled over the vector bundle Y X is a fibre bundle n : Y -+X whose typical
fibre V is an affine space modelled over V, while the following conditions hold.
All the fibres Y
x
of Y are affine spaces over the corresponding fibres Y
x
of the
vector bundle Y.
There is a bundle atlas * = {(U
a
,ip
a
)} of Y X whose local trivializations
restrict to affine isomorphisms
i Mi ) : Y* - V,
Vx e [4.
In particular, every vector bundle has a natural structure of an affine bundle.
Dealing with an affine bundle, we use only affine bundle coordinates {x
x
,y')
associated with the above-mentioned bundle atlas *. We have the fibred morphisms
YxY >Y,
x x
(y\V')^y' + y',
YxY >Y,
x x
(y
i
,j /'
<
)- 2/'- y'
i
,
where (y
1
) are linear coordinates on the vector bundle Y.
1.1. FIBRE BUNDLES 15
By virtue of Theorem 1.1.2, affine bundles have global sections.
A morphism of affine bundles $ ; Y Y' is a fibred morphism over / whose
restriction $
z
: Y
x
Vi. , to each fibre of Y is an affine map. It is called an affine
bundle morphism over / .
Every affine bundle morphism $ : Y Y' from an affine bundle Y modelled
over a vector bundle Y to an affine bundle Y' modelled over a vector bundle Y
determines uniquely the linear bundle morphism
<l>: Y - Y * ,
3$'
called the linear derivative of $>.
Let Y x Y' be the fibred product of two affine bundles Y X and Y' X
x
which are modelled over the vector bundles Y X and Y X, respectively. This
product, called the Whitney sum, is also an affine bundle modelled over Y(BY.
Furthermore, let Y' >X be an affine bundle modelled over a vector bundle Y X.
Let Y C Y' be an affine subbundle modelled over a vector bundle Y >X. Assume
that Y is the Whitney sum of Y and a complementary vector bundle Z < X. Then
one can easily verify that the affine bundle Y' X decomposes in the Whitney sum
Y' = YZ.
x
Tangent and cotangent bundles
The fibres of the tangent bundle
Kz : TZ Z
of a manifold Z are tangent spaces to Z. Given a coordinate atlas
* z = { (%*e)}
of a manifold Z, the tangent bundle is provided with the (holonomic) atlas
* ={( 7r;
I
[ /
?
) , ^=T ^) },
16 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
where by Tfy is meant the tangent map to <j>^. The associated linear bundle coor-
dinates (z
x
) on TZ are called the induced (or holonomic) coordinates with respect
to the frames {d\} in tangent spaces T
Z
Z. Their transition functions read
dz
Every manifold morphism / : Z * Z' generates the fibred morphism of the tangent
bundles
Tf :TZ >TZ',
I
i*oT/ = *.
called the tangent map to /.
The cotangent bundle of a manifold Z is the dual
K.Z :T'Z Z
of the tangent bundle TZ >Z. It is equipped with the (holonomic) coordinates
(z
x
, ix) with respect to the coframes {dz
x
} dual of {d\}. Their transition functions
read
., _ dz .
Zx
~ dz^
2
"-
Tangent and cotangent bundles of fibre bundles
Let TTY : TY Y be the tangent bundle of a fibre bundle ir : Y > X. Given
fibred coordinates (x
A
,y') on Y, the tangent bundle TY is equipped with the coor-
dinates (x
x
,y',x
x
,y').
The tangent bundle TY is a fibre bundle
ix o Try : TY - X
over X, while the tangent map TIT to 7r defines the fibration
Tn:TY T X
There is the commutative diagram
TY ^TX
I I
y x
Tf
1.1. FIBRE BUNDLES
17
The tangent bundle TY ->Y of a fibre bundle Y ->X has the vertical tangent
subbundle
VY
d
=KerTn
of TV, given by the coordinate relation i
A
= 0. This subbundle consists of the
vectors tangent to fibres of Y. The vertical tangent bundle VY is provided with the
coordinates {x
x
,y
x
,y') with respect to the frames {d,}.
Let T$ be the tangent map to a fibred morphism $ : Y Y'. Its restriction
V<f> = T$\
VY
: VY - VY',
i/"oV$ = ^$ ' =^$ ' ,
(1.1.4)
to VY is a linear bundle morphism of the vertical tangent bundle VY to the vertical
tangent bundle VY', called the vertical tangent map to $.
Every vector bundle Y >X admits the canonical vertical splitting of the vertical
tangent bundle
VY = Y x Y
x
(1.1.5)
because the coordinates y' on VY have the same transformation law as the linear
coordinates y' on Y.
An affine bundle Y > X modelled over a vector bundle Y * X also admits the
canonical vertical splitting of the vertical tangent bundle
VY^YxY
x
(1.1.6)
because the coordinates y' on VY have the same transformation law as the linear
coordinates y
1
on the vector bundle Y.
The cotangent bundle T'Y of a fibre bundle Y X is equipped with the
coordinates (x
x
,y
1
,i\,yi). There is its natural fibration T'Y X over X, but not
over T'X.
The vertical cotangent bundle VY V of a fibre bundle Y X is defined
as the vector bundle dual of the vertical tangent bundle VY >Y. It should be
emphasized that there is no canonical injection of VY into the cotangent bundle
T'Y of Y, but we have the canonical projection
C : T'Y - VY,
* Y
(1.1.7)
C : x
x
dx
x
+ i/idy* i->jfidj/*,
18 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
where (dy'} are the bases for fibres of VY, which are dual of the frames {9,} in
the vertical tangent bundle VY.
With VY and V'Y, we have the following two exact sequences of vector bundles
over Y:
0 VY TY ->Y x TX 0,
(1.1.8a)
0 - Y x T'J T T*r - v*y - 0.
(1.1.8b)
For the sake of simplicity, we will denote the pull-backs
YxTX,
x
YxT'X
x
simply by TX and T'X.
Exampl e 1.1.2. Let us consider the tangent bundle TT'X of T'X and the cotan-
gent bundle T'TX of TX. Relative to coordinates (x
x
, p\ = \) on T'X and (x
x
, x
x
)
on TX, these fibre bundles are provided with the coordinates (x
x
,p\, x
x
,p\) and
(x
x
,x
x
,\,xx), respectively. By inspection of the coordinate transformation laws,
one can show that there is the isomorphism
a : TT'X s T'TX,
P\ < Xx, Px *> i\
(1.1.9)
of these bundles over TX [43, 96].
Given a fibre bundle Y X, there is the similar isomorphism
a
v
: VV'Y S W r ,
ft <>j/i, p, < in
(1.1.10)
over VY, where (x
x
,y
t
,p
l
,y
i
,p
l
) and (x
x
,y
,
,y\y
i
,y\) are coordinates on V V V and
V'VY, respectively.
Sheaves
There are several equivalent definitions of sheaves [16, 80). We will start from
the following. A sheaf on a topological space X is a topological fibre bundle 5 >
X whose fibres, called the stalks, are Abelian groups S
x
provided with discrete
topology.
A presheaf on a topological space X is defined if an Abelian group Sy corresponds
to every open subset U C X (Sj = 0) and, for any pair of open subsets V C U,
there is the homomorphism
r
v
: Su > Sy
1.1. FIBRE BUNDLES 19
such that
r
v
= Id Su,
r
u _
r
v V
T
w ~w~v>
W C V C U.
Exampl e 1.1.3. Let X be & topological space, Su the additive Abelian group of
all continuous functions on U C X, while the homomorphism
r
v
: Su * Sy
is the restriction of these functions to V C X. Then {S
v
, ry} is a presheaf.
Every presheaf {Su, ry} on a topological space X yields a sheaf on X whose
stalk S
x
at a point x X is the direct limit of the Abelian groups Su, x e U, with
respect to the homomorphisms ry. It means that, for each open neighbourhood U
of a point x, every element s e Su determines an element s
x
S
x
, called the germ
of s at x. Two elements s Su and s' S'
v
define the same germ at x if and only
if there is an open neighbourhood W 3 x such that
r
u . _
r
v I
r
w
s r
w
s .
For instance, two real functions s and s' on X define the same germ s
x
if they
coincide on an open neighbourhood of x. The sheaf generated by the presheaf in
Example 1.1.3 is called the sheaf of continuous functions. The sheaf of smooth
functions on a manifold X is defined in a similar way.
Two different presheaves may generate the same sheaf. Conversely, a sheaf de-
fines a presheaf of Abelian groups T(U, S) of local sections of the sheaf S. This
presheaf {T(U,S),ry} is called the canonical presheaf of the sheaf S. It is eas-
ily seen that the sheaf generated by the canonical presheaf {r(U, S),ry} of the
sheaf S coincides with S. Therefore, we will further identify sheaves and canonical
presheaves.
Exampl e 1.1.4. Let Y * X be a vector bundle. The germs of its sections make
up the sheaf S(Y) of sections of Y X. The stalk S
X
(Y) of this sheaf at a point
x e X consists of the germs of sections of Y >X in a neighbourhood of x X.
The stalk S
X
(Y) is a module over the ring C(X) of the germs at x G X of smooth
functions on X. If we deal with a tangent bundle TX X, the stalk S
X
(TX) is a
Lie algebra with respect to the Lie bracket of vector fields.
20 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
1.2 Multivector fields and differential forms
Subsections. Vector fields, 20; Vector fields on fibre bundles, 21; Multivector fields,
22; The Schouten-Nijenhuis bracket, 24; Exterior forms, 25; Exterior forms on fibre
bundles, 26; Interior products, 28; Bivector fields and 2-forms, 29; The Lie derivative,
31; Tangent-valued forms, 31; Distributions, 32; Foliations, 34.
Vector fields
A vector field on a manifold Z is defined as a global section of the tangent bundle
TZ - Z. The set 7~(Z) of vector fields on Z is both a module over the ring C{Z)
of smooth functions on Z and a real Lie algebra with respect to the Lie bracket
[v, u] = (v
x
d
x
u" - v^dxv^d^, u = u
x
d\, v = v
x
d\.
A curve c : () Z, () C R, in Z is said to be an integraJ curve of a vector field
u on Z if
c =u o c,
c
A
(t) =u
A
(c(t)), te().
Recall that, for every point z Z, there exists a unique integral curve
c : ( - e, e) - Z,
( >0,
of a vector field u through z = c(0).
A vector field u on an imbedded submanifold N c Z is said to be a section of
the tangent bundle TZ Z over N. It should be emphasized that this is not a
vector field on a manifold N since u(N) does not belong to TN C TX in general.
A vector field on a submanifold N C Z is called tangent to the submanifold iV if
u(N) C TAT.
Let U C 2 be an open subset and t > 0. By a JocaJ 1-parameter group of local
diffeomorphisms of Z defined on (c, e) x (/ is meant a mapping
G :(-e,() xU3{t,z)^G,{z)eZ
which possesses the following properties:
for each t e (e, e), the mapping G
(
is a diffeomorphism of / onto the open
subset G
t
(l7) C Z;
1.2. MULTIVECTOR FIELDS AND DIFFERENTIAL FORMS 21
G
t+t
,(z) = (G
t
o G,)(z) if t + H e (-, e).
If such a mapping G is defined on 1 x Z, it is called a 1-parameter group of diffeo-
morphisms of Z.
THEOREM 1.2.1. [100]. Each local 1-parameter group of local diffeomorphisms G
on U C Z defines a local vector field u on U by setting u{z) to be the tangent vector
to the curve s(t) = G
t
{z) at ( =0. Conversely, let u be a vector field on a manifold
Z. For each z G Z, there exist a number e > 0, a neighbourhood (/ of 2 and a unique
local 1-parameter group of local diffeomorphisms on (e, e) x U, which determines
u. D
In brief, every vector field u on a manifold Z is the generator of a local 1-
parameter group of local diffeomorphisms. In particular, every exterior form (p on
a manifold Z is invariant under a local 1-parameter group of local diffeomorphisms
G
u
with the generator u, i.e.,
g'4> = 0, V
5
e G,
if and only if its Lie derivative L
u
4> along u vanishes.
If a vector field u on a manifold Z is induced by a 1-parameter group of diffeo-
morphisms of Z, then u is called a complete vector Held.
Vector fields on fibre bundl es
A vector field u on a fibre bundle Y > X is said to be projectable if it projects
over a vector field ux on X, i.e., if the following diagram
Y
JL^TY
X *TX
is commutative. A projectable vector field has the coordinate expression
U =T l V ) & + W ) 3 i . i t* = u
x
d
x
-
A vector field r = T
A
&A on a base X of a fibre bundle V X can give rise to
a vector field on Y, projectable over r, by means of some connection on this fibre
bundle (see (1.4.7) below). Nevertheless, a tensor bundle
T=(TX)(T
m
X),
22 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
admits the canonical lift
r =T*0 +[d
u
T^x^;
0
a
- + ...
/3i
T x
v0i-0
k
'J o-a, a
m
(1.2.1)
of any vector field r o n l . In particular, there exist the canonical lift
f =T<^+3T
a
"A
(1.2.2)
of r onto the tangent bundle TX, and its canonical lift
ri
f = r^dy, - d
0
T
v
x
u
-
oxff
(1.2.3)
onto the cotangent bundle T'X. Hereafter, we will use the compact notation
dx =
dx
x
(1.2.4)
A projectable vector field u =u'<9, on a fibre bundle Y * X is said to be vertical
if it projects over the zero vector field ux = 0 on X.
Let Y X be a vector bundle. Using the canonical vertical splitting (1.1.5),
we obtain the canonical vertical vector field
u
Y
= y'd,
(1.2.5)
on Y, called theLiouville vector Held. For instance, the Liouville vector field on the
tangent bundle TX reads
UTX = i
x
d\. (1.2.6)
Accordingly, any vector field r =r
x
d\ on a manifold X has the canonical vertical
lift
TV = T
X
d
X
(1.2.7)
onto the tangent bundle TX.
Multivector fields
A multivector field d of degree | t? |=r (or simply an r-vector field) on a manifold
Z is a section
4= V
1 Xr
d
Xl
A---Ad
Xr
r!
1.2. MULTIVECTOR FIELDS AND DIFFERENTIAL FORMS 23
of the exterior product l\TZ >Z. Let us denote by %{Z) the vector space of
r-vector fields on Z. In particular, T\{Z) is the space of vector fields on Z (denoted
by T(Z) for the sake of simplicity), while T
Q
(Z) is the vector space C
ca
{Z) of smooth
functions on Z. All multivector fields on a manifold Z make up the real Z-graded
vector space T,{Z) which is also a Z-graded exterior algebra with respect to the
exterior product of multivector fields.
Given a manifold Z, the tangent lift tf onto TZ of an r-vector field d on Z is
denned bv the relation
d{F,...,<j')=d{a\...^) (1.2.8)
where: (i) a
k
=a
k
dx
x
are arbitrary 1-forms on the manifold Z, (ii) by
a
k
=xdvO$dx
x
+ a
k
x
dx
x
are meant their tangent lifts (1.2.24) onto the tangent bundle TZ of Z, and (iii)
the right-hand side of the equality (1.2.8) is the tangent lift (1.2.22) onto TZ of the
function -d(o
r
,..., a
1
) on Z [67]. We then have the coordinate expression
t? = - rtf
A l V
d
A l
A - - A d
A r
,
r:
7"!
(1.2.9)
tf
Al
"*'&, A ASA, A A&
r
].
In particular, if r is a vector field on a manifold Z, its tangent lift (1.2.9) coincides
with the canonical lift (1.2.2). If an r-vector field #is simple, i.e.,
tf = T ' A - - - A T ' ,
its tangent lift (1.2.9) reads
T
d= ^ T y A - - - A f
i
- - - A r ( ; ,
i = l
where T is the vertical lift (1.2.7) onto TZ of the vector field r*.
Exampl e 1.2.1. The tangent lift of a bivector field
w^^w^d^hdv
24 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
is
w = - ( i
A
a
A
w^4 A 3 +I O""^A d
v
+ ufdn A d
v
)

Schouten-Ni jenhui s bracket
The exterior algebra of multivector fields on a manifold Z is provided with the
Schouten-Nijenhuis bracket which generalizes the Lie bracket of vector fields as
Mows [13, 181]:
[., .]SN : %{M) x T
S
(M) - T
r+S
_,(M),
(1.2.10)
r! SI
[tf,u]sK =* * t+(- l )"t;* tf ,
* * "
=
^j ( ^
2 A r
^
a
'
a
*^
2
A A 3
Ar
A d
ol
A A d
a
.).
There following relations hold:
[Ma, = (-1)
WM
M]SN.
(1.2.11)
[/, 0 A V]SH = [u, tf]
SN
A v +(-1){M-DII^
A
[v, u]
SN
,
(-i j MCM-D^(
tfi U]SN]SN
+(_i )W0"l -^[, ]
SN
]
SN
+
(-l)M(ll-i)[
t
,
j
[
V|tf
]
ss
]
SN = 0
.
(1.2.12)
(1.2.13)
Example 1.2.2. Let
w = -ui
MI/
d
M
A d
v
be a bivector field. Its Schouten-Nijenhuis bracket reads
[U>,W]SN = w^d^w^d^ A d
x
, A d
X3
.
The Schouten-Nijenhuis bracket commutes with the tangent lift (1.2.9) of mul-
tivectors [67], i.e.,
[tf,v]
SN
= [I MSN-
(1.2.14)
1.2. MULTIVECTOR FIELDS AND DIFFERENTIAL FORMS 25
Remark 1.2.3. Let us point out another sign convention used in the definition of
the Schouten-Nijenhuis bracket [125]. This bracket, denoted by [., .]SN-, is
[IMSN' = -(-1)'"[1MSN.
(1.2.15)
The relation (1.2.11) for this bracket reads
[MSN< = (-i r'-'X'^Ml sN'.
(1.2.16)
The relation (1.2.12) keeps its form, i.e.,
[v, t? A v]
SN
, = [u, tf]
SN
<At; +(_l)CM-DW
tf A
[,
v
]
m
,
t (1.2.17)
while the relation (1.2.13) is replaced by
(
_
1)
(M-.)(M-i ,
[l /] [dMsN
,
]sN
, + ( ^M- i X W- Dp,
[Vf v]sN
,
]sN
,
+
(
_
1 )
( M- i ) ( W- i )
[ w[ j /
^
s N
,
] s N
,
= 0
(1.2.18)
The equalities (1.2.16) and (1.2.18) show that, with the modified Schouten-Nijenhuis
bracket (1.2.15), the Z-graded vector space T.{Z) of multivector fields on a manifold
Z is a graded Lie algebra, where the Lie degree of a multivector field $ is | d \ 1.
In particular,
a&d{v)^[d,v\sw
(1.2.19)
is a graded endomorphism of degree | & | 1 of the graded Lie algebra T,(Z). If &
is a vector field, the endomorphism (1.2.19) is the Lie derivative
adtf(u) = L#v (1.2.20)
of the multivector field v along ti.
Exterior forms
An exterior r-form on a manifold Z is a section
cj>^-d>x
1
...x
r
dz
Xl
A---Adz
Xr
r!
of the exterior product A T'Z Z. We denote by D
r
(Z) the vector space of exterior
r-forms on a manifold Z. This is also a module over the ring D(Z) = C{Z).
26 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
From now on we will use the notation C{Z) for the ring of smooth functions on a
manifold Z, while D(Z) stands for the vector space of these functions as a rule.
All exterior forms on Z constitute the exterior Z-graded algebra 0"{Z) with
respect to the exterior product. The exterior differential is the first order differential
operator
d: D
r
(Z)- D
r +1
(Z),
d<t> = -S0
A l Ar
dz" A dz
x
> A dz
x
',
r!
on D'{Z). It obeys the relations
dod = Q,
d{4> A a) = d{4>) A a + (-l )
w
< A d(a),
where | </ >| is the degree of cj>.
Given a manifold map / : Z * Z*, by f'(j> is meant the pull-back on Z of an
r-form <j> on Z' by /, which is defined by the condition
r4>{v\ .y){z) = (PiTfiv
1
),..., Tf(v
r
))(f(z)),
Vv\ / eT
z
Z.
We have the relations
/ (^Aa) = / >A / V ,
df'4> = r{d4>).
For instance, if in N Z is a submanifold, the pull-back i j ^onto TV is called
the restriction of an exterior form<j> to N.
Exterior forms on fibre bundles
Let n : Y X be a fibre bundle with fibred coordinates (i
A
,y'). The pull-back
on Y of exterior forms on X by n provides the inclusion
it* : 0*(X) - D*(y).
Exterior forms
<&: y - A r *x,
<A =-7^A,...A
r
<ii:
Al
A---Adx
Xr
,
r!
1.2. MULTIVECTOR FIELDS AND DIFFERENTIAL FORMS 27
on Y such that d\<j> = 0 for arbitrary vertical vector field d on. Y are said to be
horizontal forms. A horizontal n-form is called a horizontal density. We will use the
notation
w = dx
1
A Adz", ^A = d\\w.
(1.2.21)
In the case of the tangent bundle TX X, there is a different way, besides
the pull-back, to lift onto TX the exterior forms on X [67, 110, 189]. Let / be a
function on X. Its tangent lift onto TX is defined as the function
/ =
X
dxJ-
(1.2.22)
Let a be an r-form on X. Its tangent lift onto TX is said to be the r-form a given
by the relation
^ ( n , . . . , f
r
) = Cr(Ti, . . . , T
r
), (1.2.23)
where r
4
are arbitrary vector fields on TX, and ?<are their canonical lifts (1.2.2
onto TX. We have the coordinate expression
a = -a
Xv
.\
r
dx
Xi
A- - - Adx
v
,
r!
a = [x
}i
d
li
ax
l
...x
r
dx
Xi
A A dx
Ar
+
r!
r
5^CV- A
r
dz
Al
A A dx
A
' A A dx
Ar
].
t = l
(1.2.24)
The following equality holds:
da = da.
Exampl e 1.2.4. Given a 2-form
fi = -Sl^dx* A dx"
on a manifold X, its tangent lift (1.2.24) onto TX reads
0 = hx^xSl^dx
11
A dx" +il^dx" A dx" + fi^dx" A dx").
(1.2.25)

28 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
Interior products
The interior product (or the contraction) of a vector field u = u^d^ and an
exterior r-form <j> is given by the coordinate expression
u
\<t> =
{
~\-
uA
*<K.-K...A,fa*' A A d/ ' A A dz
A
' =
*=l
r
-
7 7TT
M
0^,
2
...a
r
dz
a
* A A dz
Q
'.
( r - 1)!
(1.2.26)
It satisfies the relations
0(ui ,...,U
r
) =Ur\ Ui\4>,
(1.2.27)
u\{<j>Aa) = u\4> A a + {-\)
W
(j> Au\a,
(1.2.28)
[U,U']J 0 =u\d{u'\4>) - u'\d{u\4>) - u'\u\d<j>, 0 6O'(Z). (1.2.29)
The generalization of the interior product (1.2.26) for multivector fields is the
left interior product
0\4> = W),
l*!<IH
^ eD' ( Z) , tf e r.(z),
of multivector fields and exterior forms, which is derived from the equality
(j>(u
{
A--- A Mr) = ^(Ui ,...,^),
* 6 0'(Z),
u, e T(Z),
for simple multivector fields. We have the relation
d\v\<j> =(t)A #)J 0 = (-l )
|u||
%J tf|0,
*eom
i9,w<=T.(Z).
Example 1.2.5. The formula (1.2.29) can be generalized for multivector fields as
follows [13]:
[ I MSNJ 0 = (- l )
M (
"
|
-
I ,
0J dM0) +(-l)l*ljd(tfj) - ujdjd*,
where | 0 | =| i? | + | u | - 1.
The right interior product
*!.* = *(*).
Ul<l^l,
tfeO'(Z), i? 6 T.(Z),
1.2. MULTIVECTOR FIELDS AND DIFFERENTIAL FORMS 29
of exterior forms and multivector fields is given by the equalities
0( &, . . . , 0
r
) =0| VL*i .
AeO'CZ), 0 e r
r
(Z),
^ = ( T Z^^- ^- ' ^A, A A d
0r
_
0 6D'(Z).
It satisfies the relations
(tf A u)| > = tf A (v|>) + (-1)
M
(T?L<W A U, ^efl'iZ),
tf(<M<r) = tfk[<, 4>,aeO'(Z).
In particular, if | d |=| 0 |, we have the natural pairing
(,):T
r
(Z)xD
r
(Z)-C(Z),
(0, 0) =tf|0 =T?[0 =#(0) =W (1.2.30)
Bi vector fields and 2-forms
Each bivector field
w =-w^df,. A d
on a manifold Z defines the linear fibred morphism
W
t-.T'Z -*TZ,
z
w
,
(a) = w(z)[a,
a e r;z,
(1.2.31)
IW'(Q) = w
)w
{z)a
iL
d
l
,,
which fulfills the relation
to(z)(a,/J ) =w'(a)J /8 =w(z)[/?|a,
zeZ, a,(3eT'
z
Z.
One says that a bivector field w is of rani r at a point z Z if the morphism
(1.2.31) has rank r at 2. If this morphism is an isomorphism at all the points z Z,
the bivector field w is said to be non-degenerate. Such a bivector field can exist only
on an even-dimensional manifold.
The morphism (1.2.31) can be generalized to the homomorphism of graded al-
gebras >
m
(Z) T,(Z) in accordance with the relation
W*(^)(<Tl, - , <7
r
) =( - l )
r
( ^( t U
J
( f f
1
) , . . . , iu'((7r)).
(1.2.32)
0 e o m
ai CO
1
^).
30 CHAPTER 1, INTERLUDE: BUNDLES, JETS, CONNECTIONS
This is clearly an isomorphism if the bivector field w is non-degenerate.
Each 2-form
ft = ^zQ.^dz'
1
A dz
v
on a manifold Z defines the linear fibred morphism
ft
b
: TZ T'Z,
n\v) = -n^iz^dz".
(1.2.33)
One says that a 2-form ft is of rank r at a point z 6 Z if the morphism (1.2.33) has
rank r at z. This is the maximal number Ik such that A Q(z) ^0.
The kernel of a 2-form ft is defined as the kernel
Kerft
d
= \J{v e T
Z
Z : ujujft =0, Vu 6 T
Z
Z)
zez
(1.2.34)
of the morphism (1.2.33). Its fibre Ker
2
ft at a point z e Z is a vector subspace of
the tangent space T
Z
Z whose codimension equals the rank of ft at z. If a 2-form ft
is of constant rank, its kernel (1.2.34) is a subbundle of the tangent bundle TZ in
accordance with Proposition 1.1.4.
A 2-form ft is called non-degenerate if its rank is equal to dimZ at all points
z G Z. A non-degenerate 2-form ft can exist only on a 2m-dimensional manifold.
Then A ft is nowhere vanishing, and can play the role of a volume element on Z.
On a 2m-dimensional manifold Z, there is one-to-one correspondence between
the non-degenerate 2-forms ft
m
and the non-degenerate bivector fields WQ in accor-
dance with the equalities
wn(<t>,) =n,(wJi(?l)),wJi((7)), (1.2.35a)
n
w
(-d,v) = w
n
(ni(ti),{t(v)), (1.2.35b)
0,aeD'(2), tf,i/eT(Z),
where the morphisms WQ (1.2.31) and ftj (1.2.33) obey the relations
4 = (Ar\
ftiua/jU'n
=
3<
i.e.,
wtittm = 0, <(&(*)) = <t>-
1.2. MULTIVECTOR FIELDS AND DIFFERENTIAL FORMS 31
The Lie derivative
The Lie derivative of an exterior form 4> along a vector field u is given by the
equality
L
u
< = u\d(j> + d{u\4>).
In particular, if / is a function, then
Uf = u{}) = u\d}.
The relation
L
u
(< A a) =L
u
cj> A a + <j> A L
u
cr
is fulfilled. Given the tangent lift <j> (1.2.24) of an exterior form 4>, we have
L
u
(0) =u'4>
[67, 147]. The Lie derivative (1.2.20) of a multivector field v along a vector field u
is
L
U
V = [U,V]SN' = [u,f]sN,
and it obeys the equality
L(i9 Au) =L
u
tf A +t)A L
u
w
in accordance with the relation (1.2.17).
Tangent-val ued forms
Elements of the tensor product O
T
(Z) T(Z) are called the tangent-valued
r-forms
<b:Z-^AT
,
ZTZ,
<j, = - ^
Xr
dz
Xl
A A dz
Xr
d.
There is one-to-one correspondence between the tangent-valued 1-forms <j> on a ma-
nifold Z and the linear bundle endomorphisms over Z:
4>:TZ ->TZ,
0:T
2
Z3~?;J tf >(z)eT
2
Z, (1.2.36)
32 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
and
$' : T'Z T'Z,
$' : T'
Z
Z 3 B ' H <t>{z)\v" e T ; Z.
(1.2.37)
In particular, the canonical tangent-valued J -form
9
Z
= dz
x
<g>d
x
on Z corresponds to the identity morphisms (1.2.36) and (1.2.37).
Let Z =TX. There is the fibred endomorphism J of the tangent bundle TTX
of TX such that, for every vector field r on X, we have
J o f =TV , J o TV =0,
where r is the canonical lift (1.2.2) and Tv is the vertical lift (1.2.7) onto TTX of a
vector field r on TX. This endomorphism reads
J W0 = a*,
J0
X
) = 0.
(1.2.38)
It is readily observed that J o J =0, and the rank of J equals n. The endomorphism
J (1.2.38), called an almost tangent structure [110, 189], corresponds to the tangent-
valued form
<j>j = dx
x
d
x
(1.2.39)
on the tangent bundle TX.
Distributions
An n-dimensional smooth distribution on a /c-dimensional manifold Z is an n-
dimensional subbundle T of the tangent bundle TZ. We will say that a vector field
v on Z is subordinate to a distribution T if it is a section of T Z. An integral
curve of a vector field, subordinate to a distribution T, is called admissible with
respect to T.
A distribution T is said to be involutive if the Lie bracket [u, v!\ is a section of
T >Z, whenever u and u' are sections of the distribution T Z.
A connected submanifold TV of a manifold Z is called an integral manifold of a
distribution T on Z if the tangent spaces to N belong to the fibres of this distribution
at each point of N. Unless otherwise stated, by an integral manifold we mean an
1.2. MULTIVECTOR FIELDS AND DIFFERENTIAL FORMS 33
integral manifold of maximal dimension, equal to dimension of the distribution T.
An integral manifold N is called maxima] if there is no other integral manifold which
contains N.
THEOREM 1.2.2. [185]. Let T be a smooth involutive distribution on a manifold Z.
For any point z Z, there exists a unique maximal integral manifold of T passing
through z.
In view of this fact, involutive distributions are also called completely integrable
distributions.
If a distribution T is not involutive, there are no integral submanifolds of di-
mension equal to the dimension of a distribution. However, integral submanifolds
always exist, e.g., the integral curves of vector fields, subordinate to T.
We refer the reader to [68] for a detailed exposition of differential and Pfaffian
systems.
A differential system S on a manifold Z is said to be a subbundle of the sheaf
S(TZ) of vector fields on Z whose fibre S
z
at each point z Z is a submodule
of the Cf>(Z)-module S
Z
{TZ) (see Example (1.1.4)). The germs of sections of a
distribution T obviously make up a differential system S(T).
The Bag of a differential system S is the sequence of differential systems
S, =S, S
2
= [S,S], S, =[Si-i,S].
Here [S,S']
2
is the Cf (Z)-module generated by [v,u], v S
2
, u e S^. Let S(T) be
a differential system associated with a distribution T, and let
S(T) = S , c S j C -
be its flag. In general, S, is not associated with a distribution. If this is the case for
all i, we may define the Sag of a distribution
T =Ti C T
2
C .
(1.2.40)
A distribution is called regular if its flag (1.2.40) is well defined. The sequence
(1.2.40) stabilizes, i.e., there exists an integer r such that T
r
_i ^T
r
=T
r+
i [184];
moreover T
r
is involutive. In particular, if r = 1, we are dealing with the integrable
case. If T
r
=TZ, the distribution T is called totally non-holonomic.
A codistribution T* on a manifold Z is a subbundle of the cotangent bundle.
For instance, the anniMator Ann T of an n-dimensional distribution T is a (k n)-
dimensional codistribution.
34
CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
THEOREM 1.2.3. [185]. Let T be a distribution and AnnT its annihilator. Let
A Ann T be the ideal of the exterior algebra O'(Z) which is generated by elements of
Ann T. A distribution T is involutive if and only if the ideal AAnn T is a differential
ideal, i.e., d(AAnnT) C AAnnT. D
COROLLARY 1.2.4. Let T be a smooth involutive r-dimensional distribution on a
fc-dimensional manifold Z. Every point z Z has an open neighbourhood U B z
which is a domain of a coordinate chart (z\ ..., z
k
) such that the restrictions of the
distribution T and its annihilator Ann T to U are generated by the r vector fields
dz
1
''''' dz
r
and the (k - r) 1-forms dz
k r +1
,..., dz
k
, respectively. It follows that integral man-
ifolds of an involutive distribution make up a foliation.
Example 1.2.6. Every 1-dimensional distribution on a manifold Z is integrable.
Its section is a nowhere vanishing vector field u on Z, while its integral manifolds are
the integral curves of u. By virtue of Corollary 1.2.4, there exist local coordinates
(z ,..., z
k
) around each point z Z such that u is given by
u = d/dz
1
.

A Pfaffian system S' is asubmodule of the C(Z)-module D
1
(Z). In particular,
sections of a codistribution constitute a Pfaffian system. Any Pfaffian system S"
defines the ideal AS" of the exterior algebra D'(Z) which is generated by elements
of 5*. Given a flag (1.2.40) of a regular distribution, one can introduce the coBag
of the codistribution
Ann (T) D Ann (T
2
) D .
(1.2.41)
The coflag (1.2.41) stabilizes. In particular, a distribution T is totally non-holonomic
if and only if its coflag (1.2.41) shrinks to zero.
Fol i ati ons
An r-dimensional (regular) foliation on a k-dimensional manifold Z is said to be
a partition of Z into connected leaves F
L
with the following property. Every point
2.3. JET MANIFOLDS 35
of Z has an open neighbourhood U which is a domain of a coordinate chart (z
a
)
such that, for every leaf F
t
, the connected components F
t
D U are described by the
equations
z
r+1
= const., z
k
= const.
[90, 150]. Note that leaves of a foliation fail to be imbedded submanifolds, i.e.,
topological subspaces in general.
Example 1.2.7. Submersions n : Y X and, in particular, fibre bundles are
foliations with the leaves 7r"'(x), x G 7r(V) C X. A foliation is called simple if it is
a fibre bundle. Any foliation is locally simple.
Exampl e 1.2.8. Every real function / on a manifold Z with nowhere vanishing
differential df is a submersion Z R. It defines a 1-codimensional foliation whose
leaves are given by the equations
f(z) = c,
c e f{Z) c R.
This is the foliation of level surfaces of the function /, called a generating function.
Every 1-codimensional foliation is locally a foliation of level surfaces of some function
on Z.
The level surfaces of arbitrary function / ^ const, on a manifold Z define a sin-
gular foliation F on Z [90]. Its leaves are not submanifolds in general. Nevertheless
if df(z) ^0, the restriction of F to some open neighbourhood U of z is a foliation
with the generating function f\y-
1.3 J et mani fol ds
Subsections: J et manifolds, 35; Canonical horizontal splittings, 38; Second order jet
manifolds, 38; The total derivative, 40; Higher order jet manifolds, 40; Differential
operators and differential equations, 41.
Jet manifolds
Given a fibre bundle Y X with bundle coordinates (x
x
,y
l
), let us consider
the equivalence classes j^s, x X, of its sections s, which are identified by their
36 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
values s'(z) and the values of their first derivatives 3
M
s'(i) at points i e X. The
equivalence class j^s is called the first order jet of sections s at the point i l
The set J
X
Y of first order jets is provided with a manifold structure with respect to
the adapted coordinate atlas
(z\<A</i),
{x\y\tf
x
){3ls) = {x\s>{x),d
x
s\x))
}
^ = S
(a
"
+
^
)2/
"-
(1.3.1)
It is called the jet manifold of sections of the fibre bundle Y X (or simply the jet
manifold of the fibre bundle Y > X).
The jet manifold admits the natural fibrations
TT
1
: J
l
Y 3 jls - x 6 X,
(1.3.2)
*\;J
l
Y 3jls~s(x)eY, (1.3.3)
where (1.3.3) is an affine bundle modelled over the vector bundle
T'XVY - K
Y
For the sake of convenience, the fibration J
l
Y >X is further called a jet bundie,
while the fibration J
l
Y Y is an affine jet bundle.
There are the following two canonical monomorphisms of the jet manifold J
l
Y
over y:
A: J
1
Y '-->T*X7T, (1.3.4)
A =dx
x
d
A
=dx
x
(d
A
+j/^a,),
where d\ is called the total derivative, and
0! : J'Y ^T'YVY,
Y
(1.3.5)
S
l=
pd, = (dy
l
- y\dx
x
) d
where
9' =dy' - y\dx
x
(1.3.6)
1.3. JET MANIFOLDS 37
is called thecontact form. In accordance with these monomorphisms, every element
of the jet manifold J
X
Y can be represented by the tangent-valued forms
dx
x
(d
x
+ y\di) and (dy' - y\dx
x
) d
x
.
Each fibred morphism $ : F Y' over a diffeomorphism / is extended to the
fibred morphism of the corresponding jet manifolds
J^:J
X
Y J
1
?",
*
8(f~
x
Y
y'\ o J
1
* =(d,<DV
M
+ 9
M
* ' ) ^J _ ,
called the jet prolongation of the morphism $.
Each section s of a fibre bundle Y X has the jet prolongation to the section
(2/
,
,2/l )-/
1
s=(s'(x),^5
,
(x)),
of the jet bundle J
l
Y * X. A section s of the jet bundle J
l
Y >X is said to be
hoionomic if this is the jet prolongation of some section of the fibre bundle Y * X.
Any projectable vector field
u = u
x
(x")d
x
+u'(x
>
',y
]
)d,
on a fibre bundle Y X admits the jet prolongation to the vector field
u = n o j
l
u . j 'y -j'ry ->rj'y,
G =u
A
a
A
+u% + (d
x
u> - fidxurffi, (1.3.7)
on the jet manifold J
l
Y. One can show that the jet prolongation of vector fields
u i u is the morphism of Lie algebras, i.e.,
[u,u'] = [u,lf}.
In order to obtain (1.3.7), we have used the canonical fibred morphism
r, : J
l
TY - TJ
l
Y,
yin = {y')
x
- yi
x
-
*
J^:jis^j}
{x)
{^o
S
of-%
y'\ o J
1
* = (d^X +
9
M * ' )

^T - .
(^ )(x)^ i i a,
[u,u'J =[u.JZ'].
vi J-I = (y' h - y^v
(1.3.7)
(y\y\)oJ
l
s = (s>(x),d
x
s>(x)),
38 CHAPTER 1, INTERLUDE: BUNDLES, JETS, CONNECTIONS
In particular, there is the canonical isomorphism
VJ
l
Y = J
X
VY,
(1.3.8)
y'x = (y')x-
Canonical horizontal splittings
The canonical morphisms (1.3.4) and (1.3.5) can be viewed as the morphisms
A : J
l
Y x TX B d
x
>- d
x
=d
x
\X J
l
Y x TY
X Y
(1.3.9)
and
0i : J
l
Y x V'Y 9 dy* H-. 0>=9
x
\dy' J
l
Y x T'Y, (1.3.10)
where {dy'} are the bases for the fibres of the vertical cotangent bundle V'Y. These
morphisms determine the canonical horizontal splittings of the pull-backs
J'Y xTY = X(TX) 0 VY,
Y j i y
(1.3.11)
x
x
8
x
+y% = x
x
(d
x
+y\d
x
) + (y* - x
x
y\)d
and
J
l
Y x T'Y =T'X 0 0i(V*y),
Y j i y
(1.3.12)
x
x
dx
x
+ y,dy
{
= (i
A
+y,y
x
)dx
x
+y^dtf - y\dx
x
).
Second order jet manifolds
Taking the first order jet manifold of the jet bundle J 'V * X, we come to the
repeated jet manifold J
l
J
l
Y, provided with the adapted coordinates
(* ,y\y
x
,y
l
M
>2/J.A).
dx
a
y'\
x
= d^(
d
+ yfa)
d
i + yi
a
dj)y'y
There exist two different affine fibrations of J
1
J*Y over J
l
Y:
1.3. JET MANIFOLDS 39
the familiar affine jet bundle (1.3.3)
*ii : J'J'Y - J 'y,
y\Tn = y\,
(1.3.13)
modelled over the vector bundle
T'X VJ
l
Y J
l
Y, (1.3.14)
and the affine bundle
J
1
^ : J
x
J
l
Y -> / F ,
y* J^l = vU>
(1.3.15)
whose underlying vector bundle
J
l
(T'XVY)^J
1
Y (1.3.16)
differs from (1.3.14).
In general, there is no canonical identification of these fibrations, but it can be made
by means of a symmetric linear connection on X [57].
The points q J
1
J
1
Y, where n
n
(q) = ./'^(g), make up the affine subbundle
J*Y >J
1
Y of J^J^Y, called the sesquiholonomic jet manifold. This is given by
the coordinate conditions
V\x) = y\<
and is coordinated by (x
x
, y\j&,jfj,^).
The second order jet manifold J
2
Y of a fibre bundle Y >X is the affine sub-
bundle 7!-? : J
2
Y - J
l
Y of the fibre bundle pY J
1
^, given by the coordinate
conditions
vU = yU
and coordinated by (x
x
,y',y\,y\
ll
= y'^)- It
i s
modelled over the vector bundle
VT'X
The second order jet manifold PY can also be seen as the set of the equivalence
classes j
2
s of sections s of the fibre bundle Y > X, which are identified by their
yy - J
1
Y.
40 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
values and the values of their first and second order partial derivatives at points
xX:
y\(f
x
s) = d
x
s\x), VUJIS) = d^s'ix).
Let s be a section of a fibre bundle Y X and J*s its jet prolongation to a
section of the jet bundle J
l
Y X. The latter gives rise to the section J
l
J
l
s of
the repeated jet bundle J
l
J
l
Y >X. This section takes its values into the second
order jet manifold f*Y. It is called the second order jet prolongation of the section
s, and is denoted by J
2
s.
PROPOSI TI ON 1.3.1. Let s be a section of the jet bundle J
l
Y >X and J
l
s its jet
prolongation to the section of the repeated jet bundle J
1
J
i
Y > X. The following
three facts are equivalent:
s =J
l
s where s is a section of the fibre bundle Y >X;
./'s takes its values into J^Y;
J^s takes its values into J
2
Y.
D
The total derivative
We will use the total derivative operator
d^d^ + yidi + yifl.
It satisfies the equalities
d\{4> A a) = d\(4>)
A a
+ <t> A d
x
(a),
d
x
{d<t>) = d(d
x
(4>)).
Higher order jet manifolds
The k-orderjet manifold J
k
Y of a fibre bundle Y ~* X comprises the equivalence
classes j *s, x 6 X, of sections s of Y identified by the k + 1 terms of their Tailor
1.3. JET MANIFOLDS 41
series at the points x e X. The jet manifold J
k
Y is provided with the adapted
coordinates
(x
x
,y\yl...,y\
k
...
Xi
),
y\
l
.x
l
U*
s
) =
d
xr--dx
l
s
i
(x), 0<l<k.
Every section s of a fibre bundle Y > X gives rise to the section J
k
s of the fibre
bundle J
k
Y -> X such that
j / A , . A
t
^= a
A
,...a
Al S
i
, 0 <I < k.
Differential operators and differential equations
Let J
k
Y be the fc-order jet manifold of a fibre bundle V X and E * X a
vector bundle over X.
DEFI NI TI ON 1.3.2. A fibred morphism
:J
k
Y^E
x
(1.3.17)
is called a fc-order differential operator on the fibre bundle Y X. It sends each
section s(x) of Y >X onto the section ( o J
k
s)(x) of the vector bundle - !

The kernel of a differential operator is the subset
Ker = -
1
(0(X))cJ
k
Y, (1.3.18)
where 0 is the zero section of the vector bundle E X, and we assume that
Q(X) C {J
k
Y).
DEFI NI TI ON 1.3.3. A system of Ar-order partial differential equations (or simply a
differential equation) on a fibre bundle V X is defined as a closed subbundle
of the jet bundle J
k
Y - X [20, 57, 104]. D
Its (classical) solution is a (local) section s of the fibre bundle Y > X such that
its A;-order jet prolongation J
k
s lives in .
For instance, if the kernel (1.3.18) of a differential operator is a closed sub-
bundle of the fibre bundle J
k
Y ->X, it defines a differential equation
oJ
k
s = 0.
42 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
The following condition is sufficient for a kernel of a differential operator to be
a differential equation.
PROPOSI TI ON 1.3.4. Let the morphism (1.3.17) be of constant rank. By virtue of
Theorem 1.1.3, its kernel (1.3.18) is a closed subbundle of the fibre bundle J
k
Y > X
and, consequently, is a fc-order differential equation.
1.4 Connections
Subsections: Connections, 42; The curvature of connections, 44; Linear connections,
44; Affine connections, 44; Flat connections, 45.
Connecti ons
A connection on a fibre bundle Y X is defined as a global section
r : Y -* J
l
Y,
T = dx
x
{d
x
+ T\{x,y*)d
i
),
of the affine jet bundle J
l
Y Y. Combining a connection T and the morphisms
(1.3.9) and (1.3.10) gives the splittings
X o T : TX - TY,
0i o r : V'Y - T'Y
of the exact sequences (1.1.8a) and (1.1.8b), respectively. Accordingly, substitution
of the section y\ r\ into the expressions (1.3.11) and (1.3.12) leads to the familiar
splittings of the tangent bundle
TY =r(TX) VY,
Y

x
d, + y% = \d
x
+ r\d,) + (
y
< - i
x
r\)d
t
,
(1.4.1)
and the cotangent bundle
T'Y = T'X(BT(V'Y),
Y
x
x
dx
x
+ iudtf = (&
x
+T\ili)dx
x
+
yi
{dy' - r\dx
x
),
(1.4.2)
1.4. CONNECTIONS 43
of a fibre bundle Y X with respect to the connection T. In an equivalent way,
the connection V defines the corresponding projection
T:TY3x
x
d
x
+ y% -> (y* - i
x
r\)di G VY (1.4.3)
and the corresponding section
T = (dy
l
- T\dx
x
) <S> di (1.4.4)
of the fibre bundle T'Y VY -> Y.
Y
Connections on a fibre bundle Y X constitute an affine space modelled over
the linear space of soldering forms
a:Y ->T'X VY,
Y
(T = a\dx
x
(8) di-
Any connection T on a fibre bundle Y >X defines the first order differential
operator on Y
D
r
:J
l
Y3z^[z- Tiirliz))] G T'X VY,
D
r
= (y\-r\)dx
x
d
i
,
(1.4.5)
called the covarianfc differential. Its action on sections s of the fibre bundle Y reads
V
r
s = D
r
o J
l
s = [dxs* - (r o s)\]dx
x
a,. (1.4.6)
For instance, a section s is said to be an integral section for a connection T, if
V
r
s =0, i.e., T o s = J
J
s. For any section s of a fibre bundle Y > X, there exists
a connection r on Y >X such that s is its integral section. This connection is an
extension of the section s(x) i- J
l
s(x) of the affine jet bundle J
1
Y >Y over the
closed submanifold s(X) C V in accordance with Theorem 1.1.2.
A connection T on a fibre bundle Y X defines the horizontal lift
rr = r\d
x
+ r\d
{
) (1.4.7)
onto Y of each vector field T = r
x
d\ on X.
44 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
The curvature of connections
The curvature of a connection T on a fibre bundle Y X is said to be the
2-form on Y
R:Y ->AT'XVY,
Y
R= -R\
ti
dx
x
Adxdi,
R\ =^r; - d,r\ + r^r; - r^rj , (1.4.8)
Linear connections
Let Y X be a vector bundle. A linear connection on Y >X reads
r =dx
A
[a
A
+r
A
i
J
(xya,].
It defines the duai linear connection
r =dx
x
[5
A
- r^
A
(x)
%
a']
on the dual vector bundle Y* X. For instance, a linear connection .ft" on the
tangent bundle TX, and the dual linear connection K* on the cotangent bundle
T'X are given by the expressions
K =dx
x
<8> (3
A
+ ^
a
(i )i "a
Q
), (1.4.9)
A" = dx
A
(5
A
- Kf

(x)*Jr)-
(1.4.10)
Affine connections
Let Y > X be an affine bundle modelled over a vector bundle Y X. An
affine connection on Y X reads
r =dx
A
[d
A
+(r*'j (*y +r\(x))di}.
It defines the linear connection
r = dx
x
[d
x
+ r
x
l
J
(x)y
i
-?-}
on the vector bundle Y X.
1.4. CONNECTIONS 45
Flat connecti ons
Each connection T on a fibre bundle Y >X, by definition, yields the horizontal
distribution T(TX) C TY on Y, generated by the horizontal vector fields (1.4.7).
The following assertions are equivalent.
The horizontal distribution is involutive.
The connection T is flat (curvature-free), i.e., its curvature is equal to zero
everywhere.
There is an integral section for the connection T through any point y e Y.
Hence, a fiat connection r on Y * X yields the integrable horizontal distribution,
i.e., the horizontal foliation on Y, transversal to the fibration Y > X. Its leaf
through a point y 6 Y is defined locally by an integral section s
y
for the connection
T through y. Conversely, let a fibre bundle Y > X admit a horizontal foliation such
that, for each point y e Y, the leaf of this foliation through y is locally defined by
a section s
y
of Y X through y. Then the map
r : Y - J
l
Y,
r(?/) = j
l
x
s
v
,
n(y) = x,
introduces a flat connection on Y >X. Thus, there is one-to-one correspondence
between the flat connections and the horizontal foliations on a fibre bundle Y X.
Given a horizontal foliation on a fibre bundle Y > X, there exists the associated
atlas of bundle coordinates (x\ y') of Y such that every leaf of this foliation is locally
generated by the equations y
l
= const., and the transition functions y
l
> 2/"(j/
J
) are
independent of the base coordinates x
A
[23, 57]. This is called the atJ as of constant
local trivializations. Two such atlases are said to be equivalent if their union is
also an atlas of constant local trivializations. They are associated with the same
horizontal foliation. Thus, we come to the following assertion.
PROPOSI TI ON 1.4.1. There is one-to-one correspondence between the flat connec-
tions r on a fibre bundle Y >X and the equivalence classes of atlases of constant
local trivializations of Y such that T\ = 0 relative to these atlases.
46 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
1.5 Bundles with symmetries
Subsections: Tangent and cotangent bundles of Lie groups, 46; Principal bundles,
48; The linear frame bundle, 52.
Tangent and cotangent bundles of Lie groups
Let G be a real Lie group with dimG > 0 and flj [fl
r
] its left [right] Lie algebra
of ieft-invariant vector fields i(g) = TL
g
(i(e)) [right-invariant vector Selds
r
(<?) =
TRg(
r
{e))} on the group G. Here, e is the unit element of G, while L
g
and R
g
denote the action of G on itself on the left and on the right, respectively. Every
left-invariant vector field i(g) [right-invariant vector field ,(<?)] corresponds to the
element v = i(e) [v = f
r
(e)] of the tangent space T
e
G provided with both left and
right Lie algebra structures. For instance, given v T
e
G, let Vi(g) and v
r
(g) be the
corresponding left-invariant and right-invariant vector fields. There is the relation
v
l
(g) = TL
g
oTR-
g
\v
T
{g)).
Let {e
m
= e
m
(e)} [{^m = m(e)}] denote the basis for the left [right] Lie algebra,
and let c^
nn
be the right structure constants:
[
m
,e
n
] = c^
n
e
k
.
The mapping J H J ' yields the isomorphism
P fll 3 (m >- m = -m 6 0r
(1.5.1)
of left and right Lie algebras. For instance, we have
[t
m
,e
n
} = -c^
n
e
k
.
The tangent bundle TTQ : TG G of the Lie group G is trivial. There are the
isomorphisms
ft : TG 3 q - (g = n
G
(q),TL-
1
(q)) 6 G x
ft>
P r
: T G 3
?
( j =n
G
(q),TR-
i
(q)) 6 G x
flr
.
The left action L
9
of a Lie group G on itself defines its adjoint representation g t-*
Adg in the right Lie algebra g
T
and its identity representation in the left Lie algebra
gj. Correspondingly, there is the adjoint representation
e' : e i ade'() = [e',e],
ade
m
(e
n
) = c^
n
e
k
,
1.5. BUNDLES WITH SYMMETRIES 47
of the right Lie algebra g
r
in itself.
An action
G x Z B (g,z)>-> gz e Z
of a Lie group G on a manifold Z on the left yields the homomorphism
8, 3 e - & 6 T(Z)
of the right Lie algebra g
r
of G into the Lie algebra of vector fields on Z such that
Udg(s) =Tgo
e
og
l
(1.5.2)
[100]. Vector fields
m
are said to be the generators of a representation of the Lie
group G in Z.
Let g" =T*G be the vector space dual of the tangent space T
e
G. It is called the
dual Lie algebra (or the Lie coalgebra), and is provided with the basis {e
m
} dual of
the basis {e
m
} for T
e
G. The group G and the right Lie algebra g
r
act on g* by the
coadjoint representation
(Ad'g(e'),e)^{e\Adg-
1
(e)),
* G 9*, ee0r,
(1.5.3)
(adV(e*),e) = -(

*,[

',

]),
e' e ft.,
ad*
m
(e") = -c
n
mk
e
k
.
Remark 1.5.1. In the literature (see, e.g., [2]), one can meet another definition of
the coadjoint representation in accordance with the relation
{Ad'g(e'), ) = <*, Ad g(e)).
An exterior form<f> on the group G is said to be left-invariant [right-invariant] if
4>{e) =L'((j>{g)) \<j){e) = R
m
((p(g))]. The exterior differential of a left-invariant [right-
invariant] form is left-invariant [right-invariant]. In particular, the left-invariant
1-forms satisfy the Maurer-Cartan equations
<^(,6') = - ^( M ) , e,e'eg,.
There is the canonical Qi-valued left-invariant 1-form
er-TcGBe^eegi
48 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
on a Lie group G. The components #p of its decomposition 0; =0,
m
e
m
with respect
to the basis for the left Lie algebra g
(
make up the basis for the space of left-invariant
exterior 1-forms on G:
tmW = C
The Maurer-Cartan equation, written with respect to this basis, reads
<*r = ^M
n
A0f.
Accordingly, the canonical g
r
-valued right-invariant 1-form
9
T
:T
c
GBe^e&Q
T
on the group G is defined. There are the relations
e
l
{v
g
) = e
l
{TL-
g
\v
g
)) = TL-
g
\v
g
), Vg G TgG,
8
T
(v
9
) = 8
T
(TR;\v
9
)) = TR;>(v
g
),
p{0,(v
9
)) = -TL
g
oTR
9
l
9
r
(v
9
) = -Adg(e
r
(v
9
)),
where p is the isomorphism (1.5.1).
Principal bundles
We refer the reader to [100, 170, 192] for the general theory of principal bundles.
Let iT
P
: P X be a principal bundle with a real structure Lie group G. There
is the canonical free transitive action
Rc-.PxG-^P,
x
(1.5.4)
R
g
P-> pg,
peP, geG,
of G on P on the right.
A principal bundle P is equipped with a bundle atlas *p = {(U
a
,^)} whose
trivialization morphisms
V
a
T P W -> U
a
x G
obey the condition
pi
2
oip[oR
g
= go pr
2
o t/,
VpeC.
1.5. BUNDLES WITH SYMMETRIES 49
Due to this property, every trivialization morphismtp uniquely determines a local
section z
a
:U
a
-*P such that
pr
2
o v o z
a
= e.
The transformation rules for z
a
read
z
0
(x) = z
a
{x)p
a0
(x),
x e U
a
n Up, (1.5.5)
where p
a0
are the transition functions of the atlas tfp. Conversely, the family
{(U
a
,z
a
)} of local sections of P, which obey (1.5.5), uniquely determines a bundle
atlas *
P
of P.
A principal bundle P X admits the canonical trivial vertical splitting
a:VP^ PxQ
t
such that a~
l
(e
m
) are fundamental vector fields on P corresponding to the basis
elements e
m
for the left Lie algebra gj.
Taking the quotient of the tangent bundle TP * P and the vertical tangent
bundle V P of P by the tangent map TR
g
, we obtain the vector bundles
T
G
P = TP/G and V
G
P = VP/G (1.5.6)
over X. Sections of T
G
P * X are G-invariant vector fields on P, while sections of
VQP X are G-invariant vertical vector fields on P. Hence, the typical fibre of
VQP X is the right Lie algebra g
r
of the right-invariant vector fields on the group
G. The group G acts on this typical fibre by the adjoint representation.
The Lie bracket of vector fields on P goes to the quotients (1.5.6) and defines the
Lie bracket of sections of the vector bundles TQP X and VGP * X. It follows
that V
G
P >X is a fibre bundle of Lie algebras (the gauge algebra bundle in the
terminology of gauge theories) whose fibres are isomorphic to the right Lie algebra
g
r
of the group G.
Exampl e 1.5.2. When P =X x G is trivial, we have
V
G
P = X x TG/G =X x
9 r
.
Exampl e 1.5.3. Given a local bundle splitting of P, there are the corresponding
local bundle splitting of T
G
P and V
G
P. Given the basis {e
p
} for the Lie algebra g
r
,
50 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
we obtain the local fibre bases {d\,e
p
} for the fibre bundle T
G
P X and {e
p
} for
the fibre bundle V
C
P X. If
t, i;: X - T
G
P,
e=^ +^ %,
77 =r/'d,, + r/%,
are sections, the coordinate expression of their bracket is
&?] =K"fl^- V^
A
)a
A
+(^a^
r
-
v
x
d
x
e +
C
;,VK-
Let J
l
P be the first order jet manifold of a principal bundle P X with a
structure Lie group G. Bearing in mind that the jet bundle J
l
P P is an affine
bundle modelled over the vector bundle
T'XVP^ P,
p
let us consider the quotient of the jet bundle J
X
P P by the jet prolongation J
l
R
g
of the canonical action (1.5.4). We obtain the affine bundle
C = J
1
P/G-^X (1.5.7)
modelled over the vector bundle
C = T'X V
G
P -> X.
Hence, there is the canonical vertical splitting
VC^CxC.
x
In the case of a principal bundle P X, the exact sequence (1.1.8a) reduces to
the exact sequence
0 V
G
P -^T
G
P TX - 0.
x
(1.5.8)
A principal connection A on a principal bundle P X is defined as a section
A : P * J
l
P which is equivariant under the action (1.5.4) of the group G on P,
i.e.,
J
1
R
g
oA = AoR
g
, VpeG.
(1.5.9)
1.5. BUNDLES WITH SYMMETRIES 51
Turning now to the quotients (1.5.6), such a connection defines the splitting of the
exact sequence (1.5.8). It is represented by the tangent-valued form
A =dx
x
(dx + Ale,), (1.5.10)
where A
p
x
are local functions on X.
On the other hand, due to the property (1.5.9), there is obviously one-to-one
correspondence between the principal connection on a principal bundle P >X and
the global sections of the fibre bundle C -* X (1.5.7), called the bundie of principal
connections.
Let a principal connection on the principal bundle P * X be represented by
the vertical-valued form A (1.4.4). Then the form
A:P -^T'PVP
l
'^$T*PQ
l
p
is the familiar g
r
valued connection form on the principal bundle P. Given a local
bundle splitting (U^,z^) of P, this form reads
A = dp - ~A\dx
x

where Op is the canonical g;-valued 1-form on P, {t
p
} is the basis of gj, and A
p
x
are
local functions on P such that
AUpg)c
q
= Al(p)Adg-\e
q
).
The pull-back z^A of A over U^ is the well-known local connection 1-form
At = -A
q
x
dx
x
e (1.5.11)
where A
q
x
= A
9
x
o z^ are local functions on X.
It is readily observed that the coefficients A\ of this form are precisely the co-
efficients of the form (1.5.10). Moreover, given a bundle atlas of P, the bundle of
principal connections C is equipped with the associated bundle coordinates (x
x
, a
x
)
such that, for any section A of C X, the local functions
A\ = a\oA
are again the coefficients of the local connection 1-form (1.5.11). In gauge theory,
these coefficients are treated as gauge potentials. We will use this term to refer to
sections A of the fibre bundle C * X.
52 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
Let now
Y =(P x V)/G
(1.5.12)
be a fibre bundle associated with the principal bundle P >X whose structure
group G acts on the typical fibre V of Y on the left. Let us recall that the quotient
in (1.5.12) is defined by identification of the elements (p, v) and (pg,g~
l
v) for all
g 6 G. Briefly, we will say that (1.5.12) is a P-associated Rbre bundle.
As is well known, the principal connection A (1.5.10) induces the corresponding
connection on the P-associated fibre bundle (1.5.12). If Y is a vector bundle, this
connection takes the form
A = dx
x
(d
x
+ A
v
x
I'
p
d
t
),
where I
p
are generators of the representation of the right Lie algebra g
r
on V. This
is called the associated principal connection or simply a principal connection on
Y-+X.
The linear frame bundle
Let X be an n-dimensional connected oriented manifold. Let
n
LX
LX >X
be the principal bundle of oriented linear frames {s
a
} in the tangent spaces to X
(or simply the frame bundle). Its structure group is GL
+
(n,R). The frame bundle
is associated with the tangent bundle TX and with the cotangent bundle T'X of
X. Given frames {<9
M
} in the tangent bundle TX, every element {s
a
} of the frame
bundle LX takes the form
SQ S a C/^i,
where s^a are matrix elements of the group GL
+
(n, R). These matrix elements
constitute the bundle coordinates
\X , s
a
j,
S
dx* "'
on the frame bundle LX. With respect to these coordinates, the canonical action
of GL
+
(n,R) on LX on the right reads
R
g
: s*
0
H-> s"
b
g
b
a
, gGL
+
{n,R).
1.6. COMPOSITE FIBRE BUNDLES 53
The frame bundle LX is equipped with the canonical R
n
-valued 1-form
OLX = s^dx" <a, (1.5.13)
where {t
a
} is a fixed basis for R
n
, while s"^are elements of the inverse matrix of
o a-
The frame bundle, like tensor bundles, admits the canonical lift of any diffeo-
morphism / of its base X to the automorphism
/ : ( x
A
, 5
A
0
) ^( /
A
( x) , ^/ V
a
) .
These automorphisms and the corresponding automorphisms of associated bundles
are called generai covariant transformations or holonomic automorphisms. For in-
stance, in the case of the tangent bundle TX, the holonomic automorphisms f = Tf
are the tangent maps to the diffeomorphisms /. Generators of general covariant
transformations of tensor bundles are the canonical lifts r (1.2.1) of vector fields r
on X.
1.6 Composite fibre bundles
Let us consider the composition of fibre bundles
Y ->E - X, (1.6.1)
where
HYT. - Y - E
(1.6.2)
and
7T*-> v * ? , * X
(1.6.3)
are fibre bundles. This is called a composite fibre bundle. The composite fibre bundle
(1.6.1) is provided with an atlas of fibred coordinates (x
x
,a
m
, y'), where (x
M
,(r
m
)
are fibred coordinates on the fibre bundle (1.6.3) and the transition functions a
m

a
m
(x
x
,a
k
) are independent of the coordinates y
x
.
The following two assertions on composite fibre bundles are useful in applications
to field theory and mechanics [57, 159].
54 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
PROPOSI TI ON 1.6.1. Given a section h of the fibre bundle E - X and a section s
E
of the fibre bundle Y E, their composition
s = S o /i
(1.6.4)
is a section of the composite fibre bundle Y * X (1.6.1). Conversely, every section
s of the fibre bundle Y X is the composition (1.6.4) of the section h = nys o s
of the fibre bundle E X and some section s^ of the fibre bundle Y E over the
submanifold h(X) C E.
PROPOSI TI ON 1.6.2. Given a composite fibre bundle (1.6.1), let h be a global
section of the fibre bundle E X. Then the restriction
Y
h
= h'Y
(1.6.5)
of the fibre bundle Y > E to h(X) C E is a subbundle
i
h
Y
K
^ Y
of the fibre bundle Y ->X. a
Let us consider the jet manifolds J 'E, J^Y, and J
X
Y of the fibre bundles E X,
Y >E and Y X, respectively. They are parameterized by the coordinates
(
A _.m _.fn\
(x\<7
m
,2A&,24), (x^.a"
1
,?/
1
,^,^),
respectively. There is the canonical map
p : J 'E x j ' y j ' y,
r
E v
(1.6.6)
2/w = 2/>r+-
In particular, let
^=dx
x
{d
x
+ A\di) + da
m
(9
m
+4,3,) (1.6.7)
be a connection on the fibre bundle Y * E, and
r =dx
x
{d
x
+ r?d
m
)
a connection on the fibre bundle E X. Then the connection
B =p o {A o TT* x r o TT
1
) =dx
x
\d
x
+r7<9
m
+(>i ^r^ +A'
x
)d,}
(1.6.8)
1.6. COMPOSITE FIBRE BUNDLES 55
on the composite fibre bundle Y X is defined. This is called the composite
connection. In brief, we will write
B =A o r.
For instance, let us consider a vector field r on the base X, its horizontal lift IV
onto E by means of the connection V and, in turn, the horizontal lift A{Vr) of TT
onto Y by means of the connection A. Then A(FT) coincides with the horizontal
lift BT of r onto Y by means of the composite connection (1.6.8).
Given a composite fibre bundle Y (1.6.1), there are the following exact sequences
of vector bundles over Y:
0 v
E
y <->VY ->Y x VE - 0,
E
(1.6.9a)
o -> y x V*E -v*y - >y?y -o,
E *-
(1.6.9b)
where Vj;y and VY are vertical tangent and cotangent bundles of the fibre bundle
y E, respectively. Every connection A (1.6.7) on the fibre bundle Y E
determines the splittings
VY =V
E
y A(Y xK E),
K E
y-a, +a
m
d
m
= (y> - A^d)* +a
m
(d
m
+ A^d,),
VY = (Y x V"E) A(VY),
yidtf + a
m
da
m
= y,(dy
l
~ A^da) + [a
m
+ A^da,
(1.6.10a)
(1.6.10b)
of exact sequences (1.6.9a) and (1.6.9b), respectively. Using the splitting (1.6.10a),
one can construct the first order differential operator
D: J
l
Y ^T'XV
E
Y,
Y
D = dx
x
(y\ - 4 - 4 ^R
(1.6.11)
called the verticai covariant differential, on the composite fibre bundle 7- >X . This
operator can also be seen as the composition
D = pr, o D
B
: J
l
Y ~* T'X VY ->T'X Vy
E
,
r Y
where D
B
is the covariant differential (1.4.5) relative to some composite connection
(1.6.8), but D does not depend on T.
56 CHAPTER 1. INTERLUDE: BUNDLES, JETS, CONNECTIONS
The vertical covariant differential (1.6.11) possesses the following important
property. Let ft be a section of the fibre bundle E X and V), the subbundle
(1.6.5) of the composite fibre bundle Y X, which is the restriction of the fibre
bundle Y to h(X). Then the restriction of the vertical covariant differential
D (1.6.11) to JH
h
(J
l
Y
h
) C J
l
Y coincides with the familiar covariant differential on
Y
h
relative to the connection
A
h
= dx
x
[ft +(A\
n
d
x
h
m
+ (A o h)
x
)di]
on Yh, which is the restriction of the connection A to h(X) in accordance with the
commutative diagram
j i y,
J
2% j i y
A
k
! | A
E
Y
k
Y
Exampl e 1.6.1. Let r : Y J
l
Y be a connection on a fibre bundle Y X.
In accordance with the canonical isomorphism VJ 'Y = J
l
VY (1.3.8), the vertical
tangent map VT : VY VJ
l
Y to V defines the connection
VT-.VY ^ J
l
VY,
vr = dx
x
(d
x
+r\d, + djTitfdi),
(1.6.12)
on the composite vertical tangent bundle VY >X. The dual connection on the
composite vertical cotangent bundle VY >X reads
VT : VY - j ' vy,
v*r =dx
x
(^+rift - arift#).
(1.6.13)
If y X is an affine bundle, the connection VT (1.6.12) can be seen as the
composite connection (1.6.8) generated by the connection T on Y X and the
linear connection
r =dx
x
(ft +dX\y>d
x
) +dy* ft
(1.6.14)
on the vertical tangent bundle VY F.
Chapter 2
Geometry of Poisson manifolds
This Chapter is devoted to the basic geometric structures on a manifold, which we
meet in mechanics. We start from a J acobi structure whose particular case is a
Poisson structure.
Throughout this Chapter, by Z, unless otherwise stated, is meant a Ar-dimensional
manifold with coordinates (z
x
).
For the sake of convenience, the Schouten-Nijenhuis bracket [., .]SN is denoted
simply by [.,.].
2.1 J acobi structure
A Jacobi bracket (or a Jacobi structure) on a manifold Z is defined as a bilinear
map
D(Z) x D(Z) 9 (f,g) - {/,<?} G O
0
(Z)
on the vector space D(Z) of real functions on Z. This map, by definition, satisfies
the following conditions:
(Al) {g,f} = ~{f,g} (skew-symmetry),
(A2) {/, {g, ft}} + {(?, {ft, /}} +{ft, {/, <?}} =0 (Jacobi identity),
(A3) the support of {/, g} is contained in the intersection of the supports of /
and g.
A manifold Z endowed with a J acobi structure is called a Jacobi manifold.
A J acobi bracket provides the space O
0
(Z) with a structure of a Lie algebra
because it is expressed by a bidifferential operator of not more than first order in
each of its arguments [71, 98].
57
58 CHAPTER 2. GEOMETRY OF POISSON MANIFOLDS
PROPOSI TI ON 2.1.1. [98, 118, 123]. Every J acobi bracket on a manifold Z is
uniquely defined in accordance with the relation
{/. 9} =w(
rf
/>dg) + u\ {fdg - gdf)
(2.1.1)
by a pair (tu, u) of a bivector field w and a vector field u on Z such that
[u, w] = 0, [w, w] 2u A w.
(2.1.2)

Example 2. 1. 1. Taking w = 0 in (2.1.1), we find that every vector field u on a
manifold Z provides the J acobi bracket
{f,g}=u\(fdg-gdf).
The relations (2.1.2) are obviously satisfied.
Example 2.1.2. The J acobi bracket (2.1.1) with u = 0 is said to be a Poisson
bracket. According to Proposition 2.1.1, a bivector field w on a manifold Z yields a
Poisson bracket if it meets the condition
[w, w] = w^d^^dx, A d
X2
Adx,=0-
It is called a Poisson bivector Geld.
Let us consider the following examples of J acobi manifolds which are not Poisson
ones.
Example 2. 1. 3. Odd-dimensional contact manifolds are J acobi manifolds in ac-
cordance with Proposition 2.2.6.
Example 2.1.4. Let fi be a non-degenerate 2-form and cj> a closed 1-form on an
even-dimensional manifold Z such that
dfi =<j> A fi.
The triple (Z, fi, <j>) is called a locally conformally symplectic manifold. This is a
J acobi manifold characterized by the bivector field wn (1.2.35a) and the vector field
u = {nW<j>) [98, 118, 123]. If <t> = 0, we have a symplectic manifold.
2.1. JACOBI STRUCTURE 59
Let Z\ and Z2 be J acobi manifolds. A manifold map Q : Z\ >Z
2
is said to be a
Jacobi morphism if, for every pair (f,g) of functions on Z
2
, we have
{fQ,ge}i = {f,g}2e-
In particular, a vector field v on a J acobi manifold (Z;w,u) is the generator of a
local 1-parameter group of J acobi morphisms (or an infinitesimal Jacobi morphism)
if and only if
Lw =0, Lu =0.
DEFI NI TI ON 2.1.2. Given a function / D{Z) on a J acobi manifold (Z;w,u), the
vector field
d,^w\df) + }u
is called the Hamiltonian vector Held for /. D
We have the relation
[*/..] = w
(2.1.3)
It follows that the map / - 1?/ is the Lie algebra homomorphism.
The values of all Hamiltonian vector fields at all points of Z constitute the char-
acteristic distribution T on the J acobi manifold (Z;w,u). A glance at the formula
(2.1.3) shows that this distribution is involutive, but it has different dimensions at
different points z Z in general. Therefore, this distribution is not a subbundle of
the tangent bundle TZ and, consequently, is not a distribution as that in Section
1.2. A J acobi manifold Z is said to be transitive if its characteristic distribution
coincides with the tangent bundle TZ. Transitive J acobi manifolds are proved to
be either locally conformally symplectic manifolds, if dimZ is even, or the contact
ones, if dimZ is odd [98, 118].
THEOREM 2.1.3. [44, 98]. The characteristic distribution on a J acobi manifold is
completely integrable in the sense of Sussmann [173] and defines on Z a singular
Stefan foliation [171]. Each leaf of this foliation is endowed with a unique J acobi
structure such that its canonical injection into Z is a J acobi morphism.
The following definition generalizes for J acobi manifolds the notion of coisotropic
and Lagrangian submanifolds of a Poisson manifold [83] (cf. Definition 2.3.5).
60 CHAPTER 2. GEOMETRY OF POISSON MANIFOLDS
DEFI NI TI ON 2.1.4. Let (Z\w,u) be a J acobi manifold, with the characteristic
distribution T, and N a submanifold of Z. The submanifold N is said to be:
coisotropic if
w(Ann7yV)C
T
*
N
> zN,
where AnnT
z
N C T'Z is the annihilator o(T
z
N, and
Lagrangian if
w
l
(AnnT
z
N)=T
I
NnT
z
, ze N.
D
Different J acobi structures can lead to the same characteristic distribution as
follows. Let (Z\ w, u) be a J acobi manifold and a a nowhere vanishing function on
Z. Let us consider the bivector field w
a
and the vector field u
a
on Z given by
w
a
= aw, u
a
= w
i
(da) + au. (2.1.4)
Then the pair (w
a
,u
a
) (2.1.4) is a J acobi structure on Z. It is called conformally
equivalent to the J acobi structure {w,u) because of the relation
{/,0}a = ~{af,ag}.
a
It is readily observed that the Hamiltonian vector field for a function / with respect
to the J acobi structure (u, u) is also the Hamiltonian vector field for the function
i / with respect to the conformally equivalent J acobi structure (2.1.4). It follows
that conformally equivalent J acobi structures on a manifold Z define the same char-
acteristic distribution on Z.
A J acobi morphism from a J acobi manifold to a conformally equivalent J acobi
manifold is said to be a conformal Jacobi morphism. Such a morphism preserves
the characteristic distribution on a J acobi manifold.
In particular, a vector field v on a J acobi manifold (Z; w, u) is an infinitesimal
conformal J acobi isomorphism if and only if there exists a function b on Z such that
Lw = bw, h
v
u = w'(6) +bu.
2.2. CONTACT STRUCTURE 61
2.2 Contact structure
We will consider contact structures on odd-dimensional manifolds. By a contact
structure is meant a strictly contact structure in the sense of [116].
DEFI NI TI ON 2.2.1. An exterior 1-form6 on a (2m +l)-dimensional manifold Z is
said to be a contact form if
8 A (d6)
m
5* 0 (2.2.1)
everywhere on Z. O
If 6 is a contact form, so is ad where a is a nowhere vanishing function on Z.
The pair (Z, 8) (or equivalently the pair (Z,a8)) is termed a contact manifold (we
refer the reader to [15, 116] for the details of terminology).
If a manifold admits a contact structure, this manifold is orientable, and has the
volume element (2.2.1). It follows from (2.2.1) that the exterior differential d9 of a
contact form 8 is a presymplectic form of rank 2m.
There is the following variant of well-known Darboux's theorem [116].
THEOREM 2.2.2. Every point z of a (2m +l)-dimensional contact manifold (Z, 8)
has an open neighbourhood U which is the domain of a coordinate chart (z,..., z
2m
)
such that the contact form8 on U has the local expression
m
8 = dz- 5>
m+,
dz\
(2.2.2)
These coordinates are called the Darboux coordinates.
A contact form8 on a manifold Z defines the isomorphism
\>(u)
d
=u\d8+(u\8)6, u T(Z), (2.2.3)
of the C(Z)-module T(Z) of vector fields on a manifold Z to the C(Z)-module
>' (Z) of exterior 1-forms on Z [83]. This isomorphism can be extended to a mapping
from the exterior algebra T,(Z) of multivector fields to the exterior algebra D'(Z)
of exterior forms by putting
b(i A Au
r
)
d
=b(ui ) A Abfa), u
{
6 T(Z). (2.2.4)
62 CHAPTER 2. GEOMETRY OF POISSON MANIFOLDS
PROPOSI TI ON 2.2.3. Let $ be a contact form on a manifold Z. There exists the
unique nowhere vanishing vector field
E = b~
1
{9)
(2.2.5)
on Z, called the Reeb vector field, such that
E\6=l, E\de = Q
[116]. It is readily observed that, relative to the local Darboux coordinates, this
vector field reads E =d
0
. O
Example 2. 2. 1. Let M be a manifold with coordinates (q
l
) and Z the odd-
dimensional manifold R x T'M, equipped with the coordinates {t,q',Pi = ft). The
manifold R x T'M is well known to admit the contact form
6 =p,dq
{
- dt. (2.2.6)
For the sake of convenience, this contact form is chosen to differ in minus sign from
the Darboux one (2.2.2).
Exampl e 2.2.2. To generalize Example 2.2.1, let Q R be a fibre bundle, with
fibred coordinates (t, q'), and Z = V'Q the vertical tangent bundle of Q, equipped
with the coordinates (t,q\p
l
= q,). As mentioned above, V'Q is a phase space of
time-dependent mechanics. Given a connection
r = dt g> {d
t
+ Pdi)
on the fibre bundle Q R, the manifold V'Q is provided with the contact form
6
r
= pi{dq
{
- T'dt) - dt
(see Proposition 5.2.11 below). The corresponding Reeb vector field reads
E = -{d
t
+ Fd
i
-p
j
d
i
Fd
x
).
Since a connection T on a fibre bundle Q >R is flat, it defines the corresponding
atlas of local constant trivializations such that, with respect to this atlas, T = dtd
t
and the contact form 9r is brought into the form (2.2.6).
2.2. CONTACT STRUCTURE 63
Given a (2m + l)-dimensional contact manifold (Z,6), the tangent bundle TZ
of Z admits the splitting
TZ = Kerd6KeiO,
where KerdO is the 1-dimensional vector subbundle generated by the Reeb vector
field E [116]. In particular, every vector field u on Z is decomposed in a unique
fashion into
u = (u\e)E+(u-(u\6)E).
By duality, the cotangent bundle T'Z of the contact manifold Z is found to have
the splitting
T'Z =6 e Ker E,
where 0 is the 1-dimensional vector subbundle generated by the contact form6. As
a consequence, every 1-form<j) on the contact manifold Z is decomposed into
4>=(E\4>)9 + (4>-(E\4>)e)-
There is the corresponding splitting of the isomorphism (2.2.3). This sends Kerd6
onto and Ker 6 onto KerE. The restriction of b to Ker 8 coincides with the
morphism (d@f (1.2.33) defined by the presymplectic 2-form d8.
The 2m-dimensional subbundle Ker0 of the tangent bundle TZ of a contact
manifold Z is called the contact distribution on Z. Any contact form a9, where a
is a nowhere vanishing function on Z, defines the same contact distribution on Z.
As is well known, there exist integral submanifolds of the contact distribution Ker 6
of dimension m, but none of higher dimension [15]. It is readily observed that the
contact forms 6 and a6 have the same contact distribution.
DEFI NI TI ON 2.2.4. A submanifold N oia. (2m +l)-dimensional contact manifold Z
is said to be a Legendre submanifold if it is an m-dimensional integral submanifold
of the contact distribution on Z.
Let (Z
u
0i) and (Z
2
,0
2
)
be
contact manifolds. A manifold map Q : Z
x
- Z
2
is
said to be a contact transformation if
e'0
2
= o0i
64
CHAPTER 2. GEOMETRY OF POISSON MANIFOLDS
where a is a nowhere vanishing function on Z [116]. A contact transformation
preserves the contact distribution, but not a contact form in general. We consider
especially the contact automorphisms of a contact manifold (Z,9), which keeps the
contact form 9.
By definition, the Reeb vector field E (2.2.5) is the generator of a local 1-
parameter group of local contact automorphisms (or an infinitesimal contact au-
tomorphism) of the contact manifold {Z,9), i.e.,
L
E
9 =0.
It is readily observed that, if a vector field u on Z is an infinitesimal contact auto-
morphism of the contact manifold (Z, 6), then u is a Hamiltonian vector field with
respect to the presymplectic formd6 on Z, i.e., the exterior formu\d0 is exact (see
Definition 2.5.2 below). The converse assertion is the following.
PROPOSI TI ON 2.2.5. Given a contact manifold (Z, 6), let d
s
be a Hamiltonian
vector field on Z for a function / with respect to the presymplectic form d.9, i.e.,
d,\d9 = -df.
Then the vector field
#/ = [f-{*/m]E+*f
(2.2.7)
is an infinitesimal contact automorphism of the contact manifold Z, i.e.,
L
h
e = o.

Proof. It follows from direct computation. QED
Note that the vector field (2.2.7), like dj, is a Hamiltonian vector field for the
function / with respect to the presymplectic form d9.
A contact manifold is a J acobi one as follows.
PROPOSI TI ON 2.2.6. [123]. Each contact form9 on a (2m+l)-dimensional manifold
Z yields the associated J acobi bracket (2.1.2) on Z which is defined by the Reeb
vector field E (2.2.5) and the bivector field w such that
w>(4>)\9 = 0, w
t
(<t>)\d9=-(<f,-(E\(f>)9) (2.2.8)
2.2. CONTACT STRUCTURE 65
for any 1-form<f> on Z.
Relative to the local Darboux coordinates for the contact form 9, the above-
mentioned J acobi bracket (2.1.2) reads
-B+-"! > * &=.
-
m
{/.} = E(
9
m
+
,ffa,/ - a^ifdtg) + ({
9
}d
0
f - [f}d
og
),
where
m
i =l
m
[g}=^z
m
^d
m+i
g~g.
i =i
In particular, let (Z,9) be a contact manifold and {w,u) the associated J acobi
structure on Z. A submanifold N of Z is a Legendre submanifold of the contact
manifold (Z, 9) if and only if it is the Lagrangian submanifold of the J acobi manifold
(Z;w,u) [83]. Note that, in this case, the characteristic distribution of the J acobi
structure coincides with TZ, and we have
w{(AnnT
z
N) = T
z
N.
Let (Z, 0) be a contact manifold, v a vector field on Z, and b a function on Z.
One can show [83] that the pair (v, b) is an infinitesimal conformal J acobi morphism
if and only if the pair (t>, 6) is an infinitesimal contact transformation, i.e.,
L
v
9 = -66>.
Exampl e 2.2.3. Let M be a manifold with coordinates (q') and Z the odd-
dimensional manifold K x T"M, equipped with the holonomic coordinates (, q
l
,Pi =
Qi). This manifold is provided with the contact form (2.2.6). Hence, it is a J acobi
manifold where
w = -{di + p,d
t
) Ad', E = -d
t
. (2.2.9)

66 CHAPTER 2. GEOMETRY OF POISSON MANIFOLDS
2.3 Poisson structure
In accordance with Example 2.1.2, a bivector field
w = -w^dp A d
v
on a manifold Z defines a Poisson bracket (or a Poisson structure)
{f,g}=w(df,dg) = w^d
ll
fd
l/
g
(2.3.1)
on Z if and only if
[w, w] =0, w"
k,
d^w
X2X3
+ w^dpW*
3
* + w^d^w^
2
= 0.
Besides the conditions (Al - A3), the Poisson bracket (2.3.1) satisfies the Leibniz
rule
{h,fg} = {h,f}g + f{h,g}. (2.3.2)
A manifold Z endowed with a Poisson bivector field w is termed a Poisson manifold,
while a real vector space D(Z) of functions on Z forms a Poisson algebra, i.e., O
0
(Z)
is a real associative and commutative algebra with unit with respect to pointwise
multiplication, a real Lie algebra with respect to the Poisson bracket, and these two
operations intertwine via the Leibniz rule (2.3.2).
Example 2. 3. 1. Each manifold admits a zero Poisson structure characterized by
the zero bivector field w =0.
Example 2.3.2. Let u and v be vector fields on a manifold Z such that [u,v\ =0
everywhere on Z. It is readily observed from the relation (1.2.12) that w = u A v is
a Poisson bivector field.
The Poisson structure (2.3.1) defined by a Poisson bivector field w is said to be
regular if the associated morphism w' : T'Z * TZ (1.2.31) is of constant rank.
Hereafter, by a Poisson structure we mean only a regular Poisson structure.
The regular Poisson structure (2.3.1) defined by a Poisson bivector field w is
said to be non-degenerate if the associated morphism w' : T'Z TZ (1.2.31) is
of maximal rank. A non-degenerate Poisson structure can exist only on an even-
dimensional manifold.
2.3. POISSON STRUCTURE 67
Exampl e 2.3.3. A non-degenerate Poisson manifold (Z, w) is a symplectic manifold
with the symplectic form fi
w
(1.2.35b) (see Proposition 2.4.4 below).
Two functions / and g on a Poisson manifold (Z, w) are said to be in involution
with each other if their Poisson bracket {/, g} equals zero. A function C on a Poisson
manifold (Z, w) is called a Casimir function if it is in involution with any function
on Z
y
i.e.,
{<?,/} = 0, V / 6 0(Z).
Casimir functions make up the centre of the Poisson algebra (D(Z),u/). For in-
stance, if a Poisson structure is non-degenerate, all the Casimir functions (on a
connected manifold Z) are constant.
Let {Z\,wi) and {Z<i,wi) be Poisson manifolds. A manifold map g : Z\> Z
2
is
said to be a Poisson morphism if
{fQ,gQ}i = {f,g}2Q,
Vf,geQ(Z
2
),
or
W
2
= TQOWI,
where Tg is the tangent map to g. If g is a Poisson morphism, the rank of W\{z) is
grater or equal to that of W2{g{z)). If a Poisson morphismg is an immersion, tui(z)
and w
2
{g{z)) are of equal rank.
A vector field u on a Poisson manifold [Z, w) is the generator of a local 1-
parameter group of local Poisson automorphisms (or an infinitesimal Poisson auto-
morphism) of (Z, w) if and only if
\J
U
W = [u, w] = 0. (2.3.3)
Such a vector field is called canonical
Note that there are no pull-back or push-forward operations of Poisson structures
by manifold maps in general. The following assertion deals with Poisson projections
[181].
PROPOSI TI ON 2.3.1. Let (Z, w) be a Poisson manifold and i : Z - V a projection.
The following properties are equivalent:
for every pair of functions (/, g) on Y and for each point j i eV , the restriction
of the function {/ o n,g o n} to the fibre TT~
1
(y) is constant;
68 CHAPTER 2. GEOMETRY OF POISSON MANIFOLDS
there exists a Poisson structure w' on Y for which 7r is a Poisson morphism,
i.e.,
{fon,goTr} = {f,g}' OTT.

One says that the Poisson structure w' in Proposition 2.3.1 is coinduced by the
projection IT.
The direct product Z x Z' of Poisson manifolds (Z, w) and {Z',w') has the
Poisson structure defined by the bivector field w + w' on Z x Z'. It is called the
direct product of Poisson structures. Obviously, the projections p^ and pr
2
are
Poisson morphisms.
One can speak of a Poisson submanifold (Z
1
, w') of a Poisson manifold (Z, w) if
Z' is a submanifold of Z and the natural injection Z' >Z is a Poisson morphism.
DEFI NI TI ON 2.3.2. Given a real function / on a Poisson manifold (Z, w), the image
0/ = *'(#),
0, = wTdJd
u
,
(2.3.4)
of its exterior differential d/ by the morphismw
l
(1.2.31) is called the Hamiitom'arj
vector Geld for / with respect to the Poisson structure w. O
Exampl e 2.3.4. The Hamiltonian vector field for a Casimir function equals zero.
It is readily observed that each Hamiltonian vector field is also a canonical vector
field.
The Hamiltonian vector field $/ for a function /, by definition, obeys the rela-
tions
#
f
\dg = {f,g},
[0/.*] = W
V<? e D(Z),
(2.3.5)
Then it follows from (2.3.5) that / i i?y is the Lie algebra homomorphism.
In particular, we have
tf/J d/ =0
2.3. POISSON STRUCTURE 69
and, consequently, dj is tangent to the leaves of the singular foliation of the level
surfaces of the function / at its regular points, where df ^ 0.
Remark 2.3.5. It is clear that not every vector field on a Poisson manifold Z is
a Hamiltonian vector field. Given a vector field u on a manifold Z, one can try to
construct a Poisson bracket and to find a function on Z such that u would be a
Hamiltonian vector field for this function. A closely related subject is the inverse
problem in Hamiltonian mechanics, which consists in trying to represent a given
system of first order dynamic equations
z
x
= u
x
{z) (2.3.6)
on a manifold Z as the Hamilton equation with respect to some Poisson structure,
called the generating Poisson structure on Z. In Section 3.2, we will present a
general technique of constructing a generating Poisson structure for any dynamic
equation (2.3.6) at least locally [59, 82].
The values of all Hamiltonian vector fields at all points of a Poisson manifold
Z constitute an even-dimensional characteristic distribution T on the Poisson ma-
nifold (Z,w). By virtue of the relation (2.3.5), this distribution is involutive and,
consequently, completely integrable.
THEOREM 2.3.3. [181]. A Poisson structure induces a symplectic structure on
leaves of the characteristic foliation on Z, called a symplectic foliation.
In particular, if a Poisson structure is non-degenerate, its characteristic distri-
bution T coincides with TZ, and Z is a symplectic manifold.
Remark 2.3.6. It should be recalled that we are considering a regular Poisson
structure, and the corresponding symplectic foliation is non-singular.
A 2m-dimensional symplectic foliation on a /c-dimensional Poisson manifold Z
admits the adapted coordinates (... ,z
2m+1
,..., z
k
) described in Corollary 1.2.4.
Moreover, one can choose these coordinates in such a way to bring the Poisson
structure into the following canonical form [181, 186].
PROPOSI TI ON 2.3.4. For any point 2 of a Poisson manifold, there exists a coordinate
system
( <
?
i
, . . . , <r , p
1
, . . . ,
P m
, z
2 m + i
, . . . ,
2
f c
)
(2.3.7)
70 CHAPTER 2. GEOMETRY OF POISSON MANIFOLDS
in a neighbourhood of z such that
{Pi,9
>
}=Sh {q\<?} = K f t } = {q\z
A
} = {Pi,z
A
} = {Az
8
} = 0.(2.3.8)
D
The coordinates (2.3.7) are called canonical coordinates. Given canonical coor-
dinates, one says that the coordinates q' and p<are canonicaiiy conjugate.
In canonical coordinates (2.3.7), the Poisson bracket (2.3.1) takes the form
w = d' A di,
{f,9}=d'fd,g-djd\
while the Casimir functions are arbitrary functions of the coordinates z
A
. Accord
ingly, the Hamiltonian vector field for a function / reads
df = <97d, - djd'.
Let (Z, w) be a Poisson manifold with the characteristic distribution T, and N
a submanifold of Z. The familiar definitions of coisotropic and Lagrangian sub-
manifolds of a symplectic manifold are generalized for a Poisson manifold as follows
[181, 187].
DEFI NI TI ON 2.3.5. The submanifold N is said to be
coisotropic if
w(AnnTA0CTAf,
Lagrangian if
"(Ann7W) = TATnT.
a
The following theorems generalize for Poisson manifolds the corresponding re-
sults on symplectic isomorphisms (see [178]).
THEOREM 2.3.6. [187]. Let $ : Z\ Z
2
be a manifold morphism between Poisson
manifolds (Z\,w{) and (Z
2
,11)2). This is a Poisson morphism if and only if its graph
A* = {(z,$(z))ze Zi} C Z, x Z
2
2.3. POISSON STRUCTURE 71
is a coisotropic submanifold of the Poisson manifold [Z\ x Z^, W\ w-i).
THEOREM 2.3.7. [67]. Let u be a vector field on a Poisson manifold (Z, w).
(i) In accordance with the equality (1.2.14), the tangent lift w (1.2.9) of the
Poisson bivector field w is a Poisson bivector field on the tangent bundle TZ of Z.
(ii) A vector field v. is an infinitesimal Poisson automorphism of the Poisson
manifold (Z, w) if and only if u(Z) is a Lagrangian submanifold of the Poisson
manifold (TZ,w).
A Poisson structure on a manifold Z can be defined entirely by its symplectic
foliation instead of by a Poisson bivector field [181].
THEOREM 2.3.8. Let Z be a manifold and F a foliation on Z such that every leaf
F
t
of F is endowed with a symplectic structure. Given a function / 6 Q(Z) on Z,
let d'j be a Hamiltonian vector field on a leaf F
t
for the function / \p
L
with respect
to the symplectic structure on this leaf. Put
0f(')=r
f
{z),
ze F
L
.
If -df is a differentiable vector field on the manifold Z for arbitrary function /, thei
Z has a unique Poisson structure given by the Poisson bracket
{/,</}(*) =*/J<fo
whose symplectic foliation is F.
Exampl e 2.3.7. Let us consider the product Z =R x T'M with the coordinates
if-, 9%Pi)- If the cotangent bundle T'M is provided with the canonical symplectic
structure (see Example 2.4.2 below), the fibres of the projection pr, : Z R
constitute the symplectic foliation on Z which satisfies the conditions of Theorem
2.3.8. Then, the manifold K x T'M is provided with the Poisson structure given by
the Poisson bivector field
w = 9' A di.
(2.3.9)
Obviously, this is the product of the zero Poisson structure on K and that defined
by the canonical symplectic structure on the cotangent bundle T'M.
72 CHAPTER 2. GEOMETRY OF POISSON MANIFOLDS
Example 2.3.8. To generalize Example 2.3.7, let us consider a fibre bundle IT :
Z X, parameterized by fibred coordinates (x
x
;q',p
l
), whose typical fibre V is a
symplectic manifold with the Poisson bracket
{f,g} = d'fd
t
g-d
x
fd\ (2.3.10)
Each trivialization chart
V
{
: rr-^Uf) - U
(
x V
is endowed with the Poisson structure described in Example 2.3.7 with the Poisson
bracket (2.3.10). To generalize construction of symplectic vector bundles [116], we
will say that Z X is a symplectic bundle if it admits a bundle atlas $ whose
transition functions
p
K
(x) : {x} x V - {x} x V,
x e c/
f
n u
(
,
provide isomorphisms of the symplectic manifold V. Then the fibre bundle 2- >X
is equipped with the Poisson structure given by the Poisson bracket (2.3.10).
For instance, let Z = V'Q be the vertical cotangent bundle of a fibre bundle Q
K. The typical fibre of V'Q is the cotangent bundle T'M of the typical fibre M of the
fibre bundle Q >R. The fibre bundle V'Q >R is a symplectic bundle, provided
with the Poisson bracket (2.3.10) written with respect to the coordinates (, q
%
, p
(
) on
V'Q. Indeed, it is readily observed that this Poisson bracket is invariant under all
holonomic coordinate transformations of the vertical cotangent bundle V'Q. This is
the canonical Poisson structure on a phase space V'Q of time-dependent mechanics.
In Section 5.1, we will obtain it in a different way.
The notion of a symplectic bundle is naturally extended to that of a Poisson
bundle. One should distinguish this notion from that of J acobi bundles in the
sense of Kirillov [98, 123] where a J acobi bracket is generalized for sections of a
1-dimensional linear bundle, called a Jacobi bundle.
Example 2.3.9. The Lie coalgebra g* of a Lie group G is provided with the
canonical Poisson structure, called the Lie-Poisson structure (see [2, 116, 181]).
This Poisson structure is given by the bracket
{f,g}^(e'Adf(e'),dg(e')}),
/.seOV).
(2.3.11)
2.4. SYMPLECTIC STRUCTURE
73
on g*, where df(e'),dg(e') 6 fl
r
since we can regard them as linear mappings from
T
e
'2' = Q* to R. The Lie-Poisson bracket (2.3.11) is defined by the Poisson bivector
field
w
mn = C
k
mn
Z
ky
where c,
n
are the structure constants of the Lie algebra 0
r
and z
k
are coordinates
on g* with respect to a basis {e
k
}. We have the coordinate expression
{f,9}=c
k
mn
z
k
Vfd
n
g.
There is the following well-known theorem.
THEOREM 2.3.9. [186]. The symplectic leaves of the Lie-Poisson structure on the
coalgebra g* of a connected Lie group G are the orbits of the coadjoint representation
(1.5.3) of Gong*. D
2.4 Symplectic structure
Non-degenerate Poisson manifolds are symplectic manifolds (see Proposition 2.4.4
below).
DEFI NI TI ON 2.4.1. A non-degenerate exterior 2-form Q on a manifold Z is said to
be symplectic if it is closed, i.e., dQ. = 0. A manifold equipped with a symplectic
form is called a symplectic manifold.
Every symplectic manifold (Z, l) is 2m-dimensional. It is orientable, and
1 m
V =rA f i
m!
(2.4.1)
is the volume element on Z.
DEFI NI TI ON 2.4.2. A morphism (, : Z -> Z' of a. symplectic manifold (Z, Q) to a
symplectic manifold (Z', fi') is called a symplectic morphism if Q = C*ft'.
By very definition, a symplectic morphism is an immersion.
74 CHAPTER 2. GEOMETRY OF POISSON MANIFOLDS
Remark 2. 4. 1. One should distinguish a symplectic morphism from a symplecto-
morphism in the terminology of [116]. The latter is a diffeomorphism.
A vector field u on a symplectic manifold (Z, Q) is said to be a generator of a
local 1-parameter group of symplectic automorphisms (or an infinitesimal symplectic
automorphism) of a symplectic manifold (Z, fi) if and only if
L
u
n =0.
Such a vector field is called canonical
Example 2.4.2. Let M be a manifold with coordinates (q
l
) and TT.M : T'M >
M its cotangent bundle, equipped with the holonomic coordinates (q',Pi). The
cotangent bundle T'M is endowed with the canonical Liouville form
9 = p,dq\
This form is defined by the condition
v\6{p) =Tn.
M
(v)\p, Vu T
P
T'M, p T'M.
Its exterior differential is the canonical symplectic form
n = dd = dp
x
A dq* (2.4.2)
on T'M. All holonomic coordinates on the cotangent bundle T'M are canonical for
this form.
It should be emphasized that the canonical symplectic form (2.4.2) is not a
unique symplectic form on the cotangent bundle T'M. For every closed 2-form </>
on a manifold M, the form
n* = n + ir;
M
<j> (2.4.3)
is also a symplectic form on T'M [116].
The canonical symplectic form (2.4.2) plays the fundamental role in view of
Darboux's theorem [116].
THEOREM 2.4.3. Let (Z, Q) be a symplectic manifold. Each point of Z has an open
neighbourhood U which is the domain of a canonical coordinate chart
(q\...,q
m
,p
l
,...,p
m
)
2.4. SYMPLECTIC STRUCTURE 75
such that the symplectic form fi on U is given by the coordinate expression (2.4.2).
D
This theorem is an immediate consequence of Proposition 2.3.4 and Proposition
2.4.4 below.
We have the following relationship between the non-degenerate Poisson struc-
tures and the symplectic ones.
PROPOSI TI ON 2.4.4. [117]. On a 2m-dimensional manifold Z, there is one-to-one
correspondence between the symplectic forms SI and the non-degenerate Poisson
bivector fields w given by the equalities (1.2.35a) - (1.2.35b).
The canonical coordinates for a symplectic form Q are also canonical for the
corresponding Poisson bivector field w, and we have
fi =dp, A dq
l
, w^&Adi.
With respect to these coordinates, the bundle isomorphism fr (1.2.33) reads
0
k
: v'di +Vid
x
i -Vidq' + v'dp
lt
while the isomorphism w
l
= (r) (1.2.31) is
w
l
: Vidq' + v
l
dpi i v
x
d
x
Vid\
In view of Proposition 2.4.4, a Poisson structure is termed sometimes a cosym-
plectic structure (one should distinguish this notion from the cosymplectic structure
in Remark 4.8.8). A local structure of an arbitrary Poisson manifold in relation to
a symplectic one is described by the following two theorems [181, 186].
THEOREM 2.4.5. Any point of a Poisson manifold has an open neighbourhood which
is Poisson equivalent to a product of a manifold with the zero Poisson structure and
a symplectic manifold.
Let (Z,w) be a Poisson manifold. Its symplectic realization is said to be a
symplectic manifold (Z
1
, 3) together with a projection Z' > Z such that the Poisson
structure w on Z is coinduced from the Poisson structure wn on Z'.
THEOREM 2.4.6. Any point of a Poisson manifold has an open neighbourhood
which is realizable by a symplectic manifold.
76 CHAPTER 2. GEOMETRY OF POISSON MANIFOLDS
In local Darboux coordinates, this symplectic realization is seen as follows. The
Poisson structure given by the Poisson bracket (2.3.8) with respect to the canonical
coordinates is coinduced from the symplectic structure given by the symplectic form
Q = dpi A dq' + dlfy A dz
A
with respect to the coordinates
(q , - - , q ,P\,--,Pm,Z , , - Z , Z . > 2 )
by the projection
Cn
1
n
m
T) n
7
2m+l k
?
2 m+ l
y
k\
(q,...,q ,pi,,,. ,Pm,Z ,...,z,z , ...,z/t
l<? , , ? i P l , - -,Pm,Z , ,Z )
Example 2.4.3. Let V'Q be the vertical cotangent bundle of a fibre bundle Q R
as in Example 2.3.8. It is provided with the Poisson structure given by the Poisson
bracket
{f,9} = d'fd,g-d,fd'g.
Its symplectic realization is the cotangent bundle T'Q of Q endowed with the canon-
ical symplectic form
dE = dp Adt + dp, A dq', (2.4.4)
written with respect to the holonomic coordinates (t, q',p,Pi) (see Section 5.1).
Let (Zi,fti) and (Z
2
, h) be symplectic manifolds equipped with the associated
Poisson structures w-i and w
2
, respectively. The map g : Z\ >Z
2
is a Poisson
isomorphism if it is a symplectic isomorphism. Accordingly, a vector field on a
symplectic manifold is canonical if and only if it is canonical for the associated
Poisson structure. However, if dimZ\ > dimZ
2
and g is a Poisson morphism, then
this is not a symplectic morphism, i.e.,
w
7
=Tgowi, Qi ^
e
'n
2
.
The notion of a Hamiltonian vector field on a Poisson manifold is restated for a
symplectic one as follows.
DEFI NI TI ON 2.4.7. A vector field d on a symplectic manifold (Z, Q) is said to be
locally Hamiltonian [Hamiltonian] if the exterior form tfjft is closed [exact].
As an immediate consequence of Definition 2.4.7, we find the following.
2.4. SYMPLECTIC STRUCTURE
77
A vector field d on a symplectic manifold (Z, il) is locally Hamiltonian if and
only if it is an infinitesimal symplectic automorphism, i.e.,
USl = d(ti\n) = 0.
A vector field i )ona symplectic manifold (Z, fi) is Hamiltonian if and only if
it is a Hamiltonian vector field in accordance with Definition 2.3.2 for some
function / on Z, and then
d;\U= -df. (2.4.5)
The Poisson bracket defined by a symplectic form $1 reads
{/.g} = <Wj n-
In the literature (see, e.g., [2]), one may meet a definition of Hamiltonian vector
fields, which differs in the minus sign from (2.4.5).
Exampl e 2.4.4. Let us consider the cotangent bundle T'M equipped with the
canonical symplectic form fl (2.4.2). Let u = u'd
x
be a vector field on M. Then its
canonical lift
u = u
l
di PjdtvSd'
(1.2.3) onto T'M is a Hamiltonian vector field. We have
u]Q = d(u'pi).
Let us turn now to submanifolds of a symplectic manifold.
Let (Z, Q) be a 2m-dimensional symplectic manifold and w the corresponding
Poisson bivector field on Z. Let ./V be an n-dimensional submanifold of Z. The set
Orth
n
TAr= (J {
v e
T
Z
Z : v\u\Q =0, Vu T
Z
N}, (2.4.6)
is called the orthogonal of TN relative to the symplectic form Q. or simply Q-
orthogonal space ofTN.
78
CHAPTER 2. GEOMETRY OF POISSON MANIFOLDS
Recall the useful formulas
Orth
n
(Orth
n
7W) = TN,
Ann TAT = ft^OrthnTAT),
iu(AnnTAr) = Orth
n
TAr,
Ann(Orth
n
TAr) =n''(TAr).
The inclusion TN
t
C TN
2
is equivalent to OrthnTATi D Oith
n
TN
2
. We have also
Orth
n
(TAf! nTJ V
2
) =OrthnTAr, + Orth
n
TN
2
and, in particular,
TN D Orth
n
TAr =Orth
n
(TN +OrthnTAT).
Note that, in general,
TN n OrthnTAf ^0,
while TZ \
N
is not the sumTN + Orth
n
TN.
Henceforth, when there is no risk of confusion, we will write OrthTN instead of
OrthnTA^.
Let fijv = 'J v^
De
'he restriction of the symplectic form ft to the submanifold
AT. This is a presymplectic form on A^. Its kernel (1.2.34) is
Ker Q
N
= TNH Orth^TN.
The following special submanifolds of a symplectic manifold (Z, ft) are usually
considered. The presymplectic formQ
N
on any of such special manifolds is always
of constant rank.
DEFI NI TI ON 2.4.8. A submanifold N C Z of a 2m-dimensional symplectic manifold
(Z, ft) is said to be (cf. Definition 2.3.5):
coisotropic if OrthnAf C TN, n > m;
isotropic if TN C Orth^A^, n<m;
Lagrangian if OrthnAT =TN, n = m;
symplectic if ft#is a symplectic form on N.
2.4. SYMPLECTIC STRUCTURE 79

As an immediate consequence of this definition, a symplectic form restricted to
isotropic or Lagrangian submanifolds is equal to zero.
One can classify the germs of special submanifolds of a symplectic manifold [48].
Recall that a germ of a submanifold N at a point z G Z is the equivalence class
(N; z) of submanifolds of the manifold Z which pass through z and coincide with
N in an open neighbourhood of z.
With respect to local canonical coordinates (q\p
t
), a Lagrangian submanifold of
a symplectic manifold is given by the equations
q" = d
b
S,
Pa = -d
a
S,
(2-4.7)
where S(q
a
,p
0
) is a function, called the generating function, of the m variable:
{<7,P(,; a 6 A, b E B} for some partition [A, B) of the set (1, . . . , m). Then the
germ of a Lagrangian submanifold (N; z) is symplectically equivalent to the gern
of the subspace
{(q,p)R
2m
: p, =0,t = l ,...,m} (2.4.8)
at the point 0 6 R
2m
of the symplectic space M
2m
. The expression (2.4.8) results
from (2.4.7) by means of the symplectic automorphisms
q
b
>-* ~Pb, Pb<->q
b
and
Pi^Pi + d,S(q
a
,q").
Exampl e 2.4.5. Let the cotangent bundle T'M of a manifold M be equipped
with the canonical symplectic form Ci. Then the image 0(M) of the zero section
6 of T'M M is a Lagrangian submanifold {pi = 0} of the symplectic manifold
(T'M,Q).
The germ of a (2m r)-dimensional coisotropic submanifold (N; z) is symplec-
tically isomorphic to the germ of the subspace
{(g,p)R
2m
: g
i
=0,t = l ,...,r}
at the point 0 R
2m
of the symplectic space R
2m
.
80
CHAPTER 2. GEOMETRY OF POISSON MANIFOLDS
The germ of a 2(m - r)-dimensional symplectic submanifold (N; z) is symplec-
tically isomorphic to the germ of the subspace
{(
g
, p)eR
2m
: q<=
Pl
= 0,l=l,...,r}
at the point 0 e R
2m
of the symplectic space R
2m
.
We will need the following two constructions in the sequel [178].
Example 2.4.6. Let {Z
u
Qi) and (Z
2
,n
2
) be symplectic manifolds. The 2-form
fii 0 &2 =pr'fii prjf22
is clearly a symplectic form on the product Z\ x Zi- Then the graph of a sym-
plectic diffeomorphism of (Z
u
Vl{) onto (Z
2
,^2)
i s a
Lagrangian submanifold of the
symplectic manifold (Z\ x Z
2
, J li 0 f2
2
).
Example 2.4.7. Let Q be the canonical symplectic form (2.4.2) on the cotangent
bundle T'M of an m-dimensional manifold M. Then its tangent lift ft (1.2.23) is
the symplectic form
fi =dpi A dxf + dp, A dq
x
(2.4.9)
on the tangent bundle TT'M of T'M, written with respect to the holonomic co-
ordinates (<7',g',pPi) on TT'M. Due to the isomorphism a : TT'M = T'TM
(1.1.9), the form (2.4.9) is also the pull-back O'^T-TM of the canonical symplectic
formJ I TTM
on t ne
cotangent bundle T'TM of the manifold TM.
Let Hbea function on the cotangent bundle T'M. One can think of H as being
a Hamiltonian of conservative mechanics. The Hamiltonian vector field
d
n
= VUdi - diHff
for H defines the closed 2m-dimensional submanifold
q' = d"H, P, = -diH (2.4.10)
of the 4m-dimensional symplectic manifold (TT'M, fi). The restriction of the sym-
plectic form (2.4.9) to the submanifold (2.4.10) is equal to 0. Hence, this is a
Lagrangian submanifold with the generating function H(q',Pi).
2.5. PRESYMPLECTIC STRUCTURE 81
Let be a function on the tangent bundle TM. On can think of as being a
Lagrangian of conservative mechanics. The exterior differential rf of yields the
morphism
a-
1
odC : TM - T'TM - TT*M.
Its image, given by the coordinate relations
Pi = di, p, =4, (2.4.11)
is a Lagrangian submanifold of the symplectic manifold (7T*M,ft). Its generating
function is (<?',(?<).
Furthermore, let cj> bb ea nxterior r-form oo nhe eangent tundle TM, treated aa
a generalized Lagrangian [10]. Then we have the morphism
a-'o4>:TM.T'TMTT'M
and the relation
(c-
l
o4>yh = d4>.
It follows that (a-'o<j>)(TM) is a Lagrangian submanifold of the symplectic manifold
(TT'M, ft) if and only if the 1-form<j> on TM is closed.
(2.4.11)
2.5 Presymplectic structure
As in the case of Poisson structures, there are no pull-back or push-forward opera-
tions of symplectic structures by manifold maps in general. Pull-backs of symplectic
forms are the presymplectic ones.
DEFI NI TI ON 2.5.1. An exterior 2-form ftona manifold Z is said to be presymplectic
if it is closed, but not necessarily non-degenerate. A manifold equipped with a
presymplectic form is called a presymplectic manifold.
Using the formula (1.2.29), one can justify the fact that the kernel Ker ft (1.2.34)
of a presymplectic form n of constant rank is an involutive distribution, called the
characteristic distribution [116]. It defines the characteristic foliation on a presym-
plectic manifold Z. The restriction of the presymplectic form O to the leaves of this
foliation is equal to zero.
82 CHAPTER 2.
GEOMETRY OF POISSON MANIFOLDS
Remar k 2. 5. 1. In contrast with a symplectic structure, a presymplectic one is
not associated with any Poisson structure in a canonical way. Nevertheless, there
is a construction [49] which may make a Poisson bivector field w correspond to a
presymplectic form Cl on a manifold Z in accordance with the relations
w of to w = ur,
K er w' nl mn^O.

The notion of a Hamiltonian vector field on a symplectic manifold is extended
in a straightforward manner to a presymplectic manifold.
DEFI NI TI ON 2.5.2. A vector field d on a presymplectic manifold (Z, ft) is said to
be locally Hamiltonian [Hamiitonian] if the exterior form tf J Cl is closed [exact].
As in a symplectic case, we find the following.
A vector field d on a presymplectic manifold (Z, Q) is locally Hamiltonian if
and only if it is an infinitesimal automorphism of this presymplectic manifold,
i.e.,
L,,fi =d(ti\Q) =0.
A Hamiltonian vector field tij for a function / on a presymplectic manifold
obeys the relation
tij\df =0
and, consequently, is tangent to the leaves of the singular foliation of the level
surfaces of the function / at regular points, where df ^0.
Note however that, in contrast with Poisson manifolds, not each function on a
presymplectic manifold admits an associated Hamiltonian vector field.
The pull-back of a symplectic form by a manifold map is obviously a presym-
plectic form. The converse is the following.
PROPOSI TI ON 2.5.3. Any presymplectic form 0 on a manifold M can be represented
as the pull-back of some symplectic form.
2.5. PRESYMPLECTIC STRUCTURE 83
Proof. Let <f> be a presymplectic form on a manifold M. Then it is the pull-back
0 =6*(ft +7r;
M
^)
of the symplectic form ft
0
(2.4.3) on the cotangent bundle T'M of M by the global
zero section 0 of T'M >M. QED
Moreover, it is readily observed that
Orth^TO(M) = (T0)(Kertf>) C T0(M).
It follows that the imbedding 6 of the presymplectic manifold (M, <j>) into the sym-
plectic manifold (T'M, Q^) is coisotropic.
If a presymplectic form Q on a manifold M is of constant rank, the stronger
result holds [62, 72].
PROPOSI TI ON 2.5.4. Given a presymplectic manifold (M, fi) where fl is of con-
stant rank, there exists a symplectic form on a tubular neighbourhood of the zero
section of the dual bundle to the characteristic distribution Ker 0, where M can be
coisotropically imbedded.
There is another well-known pull-back construction, where a presymplectic form
is seen as a pull-back of a symplectic form by a surjective submersion.
PROPOSI TI ON 2.5.5. Let a presymplectic formQ. on manifold M be of constant rank
and its characteristic foliation be simple, i.e., its leaves are fibres of a fibre bundle
7r : M > P. Then the base P of this fibre bundle is equipped with a symplectic
form fi/>such that fl is the pull-back of Qp by n.
Proof. The symplectic form Up on P is defined by the relation
Q
p
(y)
d
!=
f
n(i7,tf)
for any v G T-n~
l
(v), vf 6 TTT"
1
^').
QED
This pull-back construction is the key point of what is called a reduction of a
symplectic structure.
84
CHAPTER 2.
GEOMETRY OF POISSON MANIFOLDS
2.6 Reduction of symplectic and Poisson structures
A submanifold N of a symplectic manifold (Z, Q) fails to be a symplectic one in
general. At the same time, if the presymplectic formQN =i%&
o n
^Y '
s
f constant
rank and the characteristic foliation of fi/v is simple, i.e., its leaves are fibres of a
fibre bundle N >P, the base P of this fibre bundle is a symplectic manifold in
accordance with Proposition 2.5.5.
This construction is called a reduction of a symplectic structure. It is general-
ized for Poisson structures. Important reduction processes appear in connection,
e.g., with constraint Hamiltonian systems and Lie group actions on symplectic and
Poisson manifolds.
DEFI NI TI ON 2.6.1. [2, 116]. A symplectic reduction of a symplectic manifold (Z, fi)
is said to be a surjective submersion 7r : N >P of a submanifold N C Z onto a
symplectic manifold (P,Q
P
), which satisfies
w'Qp = i*
N
Q.
One says that (P, fi/>) is the reduced symplectic manifold of the symplectic manifold
(Z, Q) via the submanifold N.
The above speculations show that, given a submanifold N of a symplectic ma-
nifold (Z, Q), there exists a symplectic reduction 7r : N P with connected fibres
if and only if
the presymplectic formils =?)v^
on
N '
s
of constant rank,
the characteristic foliation of Qfj is simple.
PROPOSI TI ON 2.6.2. [116]. Let (Z, f2) be a symplectic manifold and N a submani-
fold of Z.
(i) If n : N >P is a symplectic reduction, then the rank of the pull-back
presymplectic form l
N
on N is constant and equal to di mP, Kerfi/y = VN, and
the connected components of fibres of 7r are the leaves of the characteristic foliation
for fi/v. Furthermore, if n has connected fibres, the characteristic foliation of Qjv is
simple, and is exactly the fibration N >P.
(ii) Conversely, if the rank of the presymplectic form fijv on a submanifold N of
a symplectic manifold (Z, fl) is constant and the characteristic foliation for f2^is
2.6. REDUCTION OF SYMPLECTIC AND POISSON STRUCTURES 85
simple, the base P of the corresponding fibration N >P has a unique symplectic
form such that this fibration is a symplectic reduction of (Z, fi) via N. This is a
unique symplectic reduction via N with connected fibres. Furthermore, if N P'
is another symplectic reduction, there exists a surjection P > P' which is a local
symplectic isomorphism.
A coisotropic submanifold N provides the most interesting case when the presym-
plectic form fijv is of constant rank which is neither dimN nor 0.
Exa mple 2. 6. 1. Let [M,<j>) be a presymplectic manifold, where a presymplectic
form4> is of constant rank and its characteristic foliation is simple. Combining the
pull-back constructions in Propositions 2.5.3 and 2.5.5 gives the reduction of the
symplectic structure 0.$ (2.4.3) on the cotangent bundle T*M via the coisotropic
submanifold 0(M).
The reduction procedure is extended to Poisson manifolds as follows.
DEFI NI TI ON 2.6.3. [129]. By a reductive structure of a Poisson manifold (Z, w) is
meant a triple (Z, N, E) of a submanifold N of M and a vector subbundle E C TZ \N
together with a submersion 7r : N * P if the following conditions are satisfied:
(i) E fl TN is tangent to the fibres of the submersion N >P;
(ii) if df and dg, where /, g D(Z), belong to Ann E, so is d{f, g}
w
.
A reductive structure is said to be a Poisson reduction if: (i) P above is a Poisson
manifold with a Poisson bivector field W and (ii) for any local functions /, g on
P and for any local extensions f,g onto Z of the pull-backs / o ir,g o ix such that
df,dg C AnaE, the relation
{7,9~}wiN = {f,g}wn
holds good. One says that (F, W) is the reduced Poisson manifold of (Z, w) via
(N,E). a
The following assertion furnishes the necessary and sufficient conditions for a
reductive structure of a Poisson manifold to be a Poisson reduction.
PROPOSI TI ON 2.6.4. [129, 181]. Let (Z, N, E) be a reductive structure of a Poisson
manifold (Z, w). This is a Poisson reduction if and only if
io(Ann) CTN + E.
(2.6.1)
86
CHAPTER 2.
GEOMETRY OF POISSON MANIFOLDS
a
The functorality property of the Poisson reduction is given by the following.
PROPOSI TI ON 2.6.5. [129]. Let (Z,N,E) and (Z',N',E') be Poisson reductions,
and let $ : Z Z' be a Poisson map such that ${N) C N', T$(E) C E', and $
sends the leaves of JV to the leaves of N'. Then $ induces a unique Poisson map
$ : P P', called the reduction of $, such that n' o $ =$ o n.
Exa mple 2.6.2. Let Z = T'G, where G is a connected Lie group, N = Z, and
comprise the tangent space to the right G-orbits. Then (Z, E) is a Poisson reduction,
where ix(Z) = T'G/G = g' is the Lie coalgebra of G provided with the Lie-Poisson
structure (2.3.11) [129, 130].
Exa mple 2.6.3. To compare a Poisson reduction with a symplectic one, let (Z, fl)
be a symplectic manifold and w the corresponding Poisson bivector field on Z.
Let N be a submanifold of Z such that the presymplectic form QN = i%^ is of
constant rank. Put E = OithTN. Then E l~l TN = Kerft
N
is tangent to the
characteristic foliation of QN- If there exists a symplectic reduction ir : N P
of (Z, Q) to (P, ftp), one can show that (Z, N, OrthTTV) is a reductive structure of
the Poisson manifold (Z, w). The condition (i) of Definition 2.6.3 holds since the
integral manifolds of the distribution Ker fijv are connected components of fibres of
N P. Furthermore, since we are on a symplectic manifold, df G Ann (OrthTW)
if and only if the Hamiltonian vector field &; for / belongs to
iu
,
(Ann(OrthTAf))=TAf. (2.6.2)
Then, d
h
d
g
e TN implies
{#f,ti
s
}=#
{
j,
9)
eTN.
Hence, condition (ii) of Definition 2.6.3 also holds. Moreover, the inclusion (2.6.1)
obviously takes place because of (2.6.2). It follows that (Z, N, OrthTN) is a Poisson
reduction and the corresponding reduced structure W on P is associated with the
symplectic form fip. On the other hand, if we take E = Kerfi/v, the inclusion
(2.6.1) requires N to be coisotropic, since
tu^Ann E)=TN + OrthTN,
2.6.
REDUCTION OF SYMPLECTIC AND POISSON STRUCTURES 87
and one may apply the previous result.
One can extend this Example to the Poisson manifold case as follows.
LEMMA 2.6.6. [181]. Let N be a. submanifold of a Poisson manifold (Z, w) such
that
C(N)
d
^w
l
(AnnTN)C\TN (2.6.3)
is a distribution, i.e., is of constant rank. Then C(N) is involutive. Furthermore, if N
is transversal to the leaves of the symplectic foliation F for w on Z, the distribution
C(N) defines a subfoliation of F D TN which is transversally symplectic along each
leaf of the latter. D
If the foliation C(N) is simple and corresponds to a fibration n : N > P,
then F n TN projects to a symplectic foliation n(F) on P, and P has a well-
defined Poisson structure given by this foliation. This is called the leafwise reduction
of (Z,w) via N. It follows in the same way as for the symplectic case discussed
earlier that, if w
l
(AnnTN) has a constant dimension, then (Z, N, w^AnnTN)) is
a reductive structure. Because of
w^Anniw^AnnTN))) C TN
and the relation (2.6.1), the reductive triple (Z, N, w
t
(Ann TN)) is also a Poisson
reduction, while the reduced Poisson manifold P is exactly the one defined by the
leafwise reduction.
Given a Poisson manifold (Z,w), let (P,W) be a reduced Poisson manifold of
(Z, w) via a submanifold N. Then one can say that the Poisson algebra Q(P) of
smooth functions on P is the reduction of the Poisson algebra of Q(Z) of smooth
functions on Z. If N is a closed imbedded submanifold of a Poisson manifold Z,
this reduction is described in an intrinsic way, i.e., in terms of ideals of the Poisson
algebra D(Z) as follows [97].
Let N be a closed imbedded submanifold of a Poisson manifold (Z,w). Let us
consider the set / of real functions f on Z which vanish on N, i.e., i'
N
f = 0. This
set is the kernel / = Ker i'
N
of the linear morphism
t'
N
: 0(Z) - Q(N),
88
CHAPTER 2.
GEOMETRY OF POISSON MANIFOLDS
and it is an ideal of the associative commutative algebra Q(Z). Since N is a closed
and imbedded submanifold of Z, we have the isomorphism
Q{Z)/I =* >(N)
(2.6.4)
of associative commutative algebras.
Let us consider the space of all vector fields on Z which restrict to vector fields
on N. It is given by
T
N
d
^{u g T{Z) : u\df 6 /, V/ / }.
(2.6.5)
It is clear that 7]v ^T(Z) |
w
. Suppose that -d; is a Hamiltonian vector field which
restricts to a vector field on N. By the very definition (2.6.5), this means that
4f\dg = {f,g)zl, \fgl.
Hence, the functions whose Hamiltonian vector fields restrict to vector fields on N
are those functions in the normalizer of /, denoted by
7(/v)
d
J =
f
{/e0(Z): {/,<?} e/ , v
5
e/ }.
(2.6.6)
It follows from the J acobi identity that the normalizer (2.6.6) is a Poisson subalgebra
of Q(Z). Put
I'(N)
d
=I{N)nl. (2.6.7)
It is naturally a Poisson subalgebra of I{N). This subalgebra is generally non-zero
since I
2
C I'(N) due to the Leibniz rule.
The following Theorem leads us back to the reduction procedure.
THEOREM 2.6.7. [97]. Let N be a closed imbedded submanifold of a Poisson
manifold Z. Let us assume that w^AnnTN) and C (2.6.3) are distributions of
constant rank, and the foliation C is simple. Then (Z, N, u)"(Ann TN)) is a reductive
structure such that there is the associative commutative algebra isomorphism
D(F) S I(N)/I'(N).
(2.6.8)
Since the right-hand side is naturally a Poisson algebra, this isomorphism defines a
Poisson structure on P. O
Remark 2.6.4. The result is based on the fact that all sections of w
i
(AnnTN) are
restrictions to N of Hamiltonian vector fields for elements in /, while all sections of
2.7. APPENDIX. POISSON HOMOLOGY AND COHOMOLOGY 89
C are Hamiltonian vector fields for elements in I'(N) restricted to N. In particular,
if N is coisotropic, / C I(N), i.e., / =I'{N) is a Poisson subalgebra of D(Z).
Theorem 2.6.7 shows that a Poisson reduction procedure can be seen as an
algebraic one; that leads to the following purely algebraic definition [97].
DEFI NI TI ON 2.6.8. Let V be a Poisson algebra, J be an ideal of V, J{N) its
normalizer (2.6.6), and J'(N) = J{N) D J. One says that the Poisson algebra
J(N)/J'(N) is the reduction of the Poisson algebra V via the ideal J.
In particular, an ideal J of a Poisson algebra V is said to be coisotropic if J is
a Poisson subalgebra of V.
The above algebraic reduction procedure has been generalized for J acobi mani-
folds [87].
We meet the Poisson reduction in the sense of Definition 2.6.8 in connection with
the BRST scheme [97, 169].
2.7 Appendix. Poisson homology and cohomology
This Section is concerned with the Koszul-Brylinski-Poisson homology and the
Lichnerowicz-Poisson cohomology of a Poisson manifold, which find their applica-
tion in the quantization procedure.
Remar k 2. 7. 1. Let us recall briefly the basic notions of homology and cohomology
of complexes [17, 122].
A sequence
n J? n
3
i D * h_ n
a
ZL
U * >o * " 1 * ' ' ' *
a
v ' ' '
(2.7.1)
of Abelian groups B
p
and homomorphisms d
p
is said to be a chain complex if
d
p
od
p+1
=0, VpeN,
i.e., Imdp+i C Ker3
p
. The quotient
H
p
(B.)
d
^Keid
p
/lmd
p+i
is called the pth homology group of the chain complex B, (2.7.1). The chain complex
(2.7.1) is said to be exact at an element B
p
if H
P
(B.) = 0. It is called an exact
sequence if it is exact at each element.
90
CHAPTER 2.
GEOMETRY OF POISSON MANIFOLDS
A sequence
0 ^B ^ B
l
.*-,... ^B P . * . . .
(2.7.2)
of Abelian groups B
v
and homomorphisms 6
P
is said to be a cochain complex if
S" o 5*-
1
= 0, VpeN.
The pth cohomology group of the cochain complex S" (2.7.2) is the quotient
H
p
(B')
d
=Ker&'/lm6
p
-
1
.
The De Rham compiex of exterior forms on a manifold Z
... J - 0'- i ( z) -*-.0(Z) - ^D
r +1
( Z) -*-....
exemplifies a cochain complex. Its cohomology group H
P
(Z), called the De Rham
pth cohomology group, is the quotient of the space of closed p-forms by the subspace
of exact p-forms.
Given a Poisson manifold (Z, w), let us consider the operator
6
w
:Q
r
(Z)^D
r
-\Z),
6
W
= w\ o d d o w\, (2.7.3)
on the exterior algebra D*(Z), called the Poisson codifferential [19, 102, 181]. It is
given by the coordinate expression
SModfi A A df
T
) = i >l )
, + 1
{/o, h}dh A A df, A A df
T
+
i =i
(-l)*
+
Vod{/i, h) A df
x
A A df
i
A A df
i
A - A df
r
.
l<i<j<r
Henceforth, when there is no risk of confusion, we shall write 5 instead of 5
W
.
The operator 6 (2.7.3) obeys the equalities
6 o 6 =0,
do6 + 6od=0.
(2.7.4)
(2.7.5)
Due to the nilpotency property (2.7.4), we have the chain complex
... J-)P-i(z) J-D
P
{Z) ^- D
P +1
( Z) J- ,
2.7. APPENDIX. POISSON HOMOLOGY AND COHOMOLOGY 91
called thecanonical complex of the Poisson manifold (Z, w). The homology H
n
(Z)
of this complex is termed thecanonical or theKoszul-Brylinski-Poisson homology.
Furthermore, due to the property (2.7.5), the periodic double complex
E
T
,
t
(Z) = O
l
~
T
(Z),
6 : E
r
,
t
(Z) - E
r
^(Z), d:E
ri
{Z)^E
T
^{Z),
can also be defined.
Let (Z, fi) be a 2m-dimensional symplectic manifold, provided with the volume
form V (2.4.1), and IUQ the corresponding Poisson bivector field. Imitating the
well-known star isomorphism for Riemannian manifolds, one introduces the star
operation
* : Q
T
(Z) - D
2m
-
r
(Z),
where w
l
is the homomorphism (1.2.32). The (2m r)-form *0 is called the adjoint
of the r-form </> relative to the symplectic form Q. The following equalities
*{**) = <, 0 6 D'(Z),
are fulfilled [116].
LEMMA 2.7.1. [19]. In the case of a symplectic manifold, the codifferential S (2.7.3)
on >
T
{Z) is given by the formula
6 = {-iy+
l
*d*. (2.7.6)

The relation (2.7.6) establishes an isomorphism of the canonical homology group
H^{Z) with the De Rham cohomology group H
2m
~'(Z) of a symplectic manifold
as follows.
THEOREM 2.7.2. [19, 181]. If (Z, 0) is a 2m-dimensional symplectic manifold, then
Hf (Z) = H^-^Z).
(2.7.7)
*(cj>Aa) =4W)J <**) = (-l)
m
K()\(*<t>)
0=u>?,(fljy,
92
CHAPTER 2.
GEOMETRY OF POISSON MANIFOLDS
a
Given a Poisson manifold (Z, w), let us now introduce the operation
w : %{Z) -* T
r +1
(Z),
(2.7.8)
on the exterior algebra 7;(Z) of multivector fields on Z, where [.,.] is the Schouten-
Nijenhuis bracket (1.2.10). This operation satisfies the rules
w o w =0,
w{ti Av) = ui(i?) A D +(-l )
w
tf A w(v),
(2.7.9)
(2.7.10)
and is sometimes called a contravariant exterior differential [181]. Its relation to the
familiar exterior differential is
iB(w
>
(<j>)) = -w
l
(d<j>),
<t>eO'(Z),
(2.7.11)
while that to the codifferential 6
W
(2.7.3) is given by the formula
w{-d)\<t> = ti\6
m
{4>) + {-\)
w
6
w
{d\<j>),
|i ?H0] - l .
(2.7.12)
Exa mple 2.7.2. It is readily observed that
-w(f) = [w, /]=*,,
/ e To(Z),
is the Hamiltonian vector field for a function / on Z.
Due to the nilpotency rule (2.7.9), the operation w (2.7.8) provides the cochain
complex
^%.,{Z) ^%{Z) ^T
r+l
(Z) - ^,
called the Lichnerowicz-Poisson cochain complex. Cohomology groups Hl
P
(Z,w)
of this complex are said to be Lichnerowicz-Poisson or LP-cohomology groups of
the Poisson manifold {Z,w).
Exa mple 2.7.3. The LP-homology group //
p
(Z,w) is the centre of the Poisson
algebra (D(Z),w). It consists of functions / T
0
(Z) modulo constant functions
whose Hamiltonian vector fields dj =w(f) KeriD vanish.
gWg-K4 tfer.(Z),
2.7. APPENDIX. POISSON HOMOLOGY AND COHOMOLOGY 93
Exa mpl e 2. 7. 4. The first LP-homology group Hl
P
(Z,w) is the space of canonical
vector fields u for the Poisson bivector field w (i.e., h
u
w = -w(u) = 0) modulo
Hamiltonian vector fields -w(f), f e TQ{Z).
Exa mpl e 2.7.5. The second LP-homology group Hl
P
{Z,w) has the distinguished
element {w} whose representative is the Poisson bivector field w itself. We have
{w} =0 if there is a vector field u on Z such that w = w(u) = h
u
w.
It is readily observed that, due to the property (2.7.10), Hl
P
(Z,w) makes up a
graded commutative algebra with respect to the product
{i>}A{i;}
d
%At;}.
It is also provided with the bracket
[W,M]"{fM)
in accordance with the relation (1.2.13).
THEOREM 2.7.3. [181]. The relation (2.7.11) induces a homomorphism of the
graded algebras
H'(Z) - H'
h
{Z) (2.7.13)
of the De Rham cohomology and the LP-cohomology. This is an isomorphism if the
Poisson bivector field w comes from a symplectic structure on Z. D
In case of a symplectic manifold, the isomorphism (2.7.13) and the isomorphism
(2.7.7) lead to the isomorphism
HT(Z) = Hlr\Z).
In general, we have the following relationship between the LP-cohomology and the
canonical homology of a Poisson manifold.
PROPOSI TI ON 2.7.4. By virtue of the relation (2.7.12), the natural pairing (1.2.30)
induces the corresponding pairing
(,) : HUZ) x Hr(Z) - H^{Z),
(m,{<t})=mm,
94
CHAPTER 2.
GEOMETRY OF POISSON MANIFOLDS
of vector spaces of the LP-cohomology and the canonical homology of a Poisson
manifold.
The constructions of the LP-cohomology and the canonical homology for Poisson
manifolds can be extended to the J acobi ones [115].
Given a J acobi manifold (Z;w,u), let us define the operator
vm:T
r
(Z)->T
T+1
{Z),
wu
d
=-[w,#]+ruAi), dT
r
(Z),
which generalizes the contravariant exterior differential (2.7.8). This operator sat-
isfies the rules
L
u
o wu = wu o L
u
,
55
2
(i)) =- M A i o, i )eT,(Z),
wu(-d A v) = wu{ti) Av+ (-l )
1
*^A wu(v).
(2.7.14)
Its connection with the exterior differential is given by the relations
L
u
(ti/'(0)) =w>(L
u
<t>),
wu(w*(<p)) = -w
i
{d<j>) +u/^uj^) A w.
(2.7.15)
(2.7.16)
A glance at the expression (2.7.14) shows that the operator wu becomes nilpotent
on the subspaces of u-invariant multivectors
T
r
{Z) ={tf %{Z) : L
u
tf = [u, &] =0}.
This allows us to introduce the cochain complex
IUU 3 = / rj\ M1U 7r I ry\ WU ~rf- j
l
y\ VJU
The cohomology groups H[^{Z,w) of this complex are called Lichnerowicz-Jacobi
or LJ-cohomology groups of the J acobi manifold (Z; w, u). It is clear that, if u =0
and (Z, w) is a Poisson manifold, the LJ -cohomology is precisely the LP-cohomology.
Exa mple 2.7.6. The first LJ -cohomology group H[
p
(Z,w) has the distinguished
element whose representative is the vector field u, while the second LJ -cohomology
group H{jp(Z,w) has the distinguished element whose representative is w.
2.7. APPENDIX. POISSON HOMOLOGY AND COHOMOLOGY 95
In the same way that the LP-cohomology groups are connected with the De Rham
cohomology groups, the LJ -cohomology groups are connected with the cohomology
groups of the subcomplex of basic forms of the De Rham complex.
An r-form </> on a J acobi manifold (Z;w,u) is said to be basic if
uJ 0 =O, L
U
0 =O.
It is easily seen that, if <fi is a basic form, so is its exterior differential d<p. Hence,
basic forms constitute a subcomplex
^O
r
B
-
1
(Z) -^O
T
{Z)
B
J ^D
r +1
( Z)
B
- * - ..- (2.7.17)
of the De Rham complex.
On subspaces >
T
B
{Z) C D
T
(Z) of basic forms, the relation (2.7.16) takes the
form similar to (2.7.11):
wu(ui'(0)) =w
>
(d(j)),
4> e D
B
(Z). (2.7.18)
Moreover, we deduce from the relation (2.7.15) that, if </> is a basic r-form, then
w
l
((j>) T
r
(Z). In the same way as in Theorem 2.7.3, the equality (2.7.18) yields a
homomorphism
H'
B
(Z) - H'
LJ
(Z)
from the cohomology of the complex (2.7.17), called the basic De Rham cohomology,
to the LJ -cohomology.
To introduce the canonical homology of a J acobi manifold (Z; w, u), let us con-
sider the codifferential 6
W
(2.7.3) defined by the bivector field w. It was proved in
[38] that, if ^is a basic form, so is 8
W
(4>), and 8
W
o 6
W
= 0. Then the chain complex
h-O'-^Z) ^Q
P
B
{Z) &-&$
l
{Z) ^
is defined. The homology of this complex is called the canonicai homology of a
J acobi manifold. If u = 0, this is exactly the canonical homology of a Poisson
manifold.
Similarly to Proposition 2.7.4, we have the following relationship between the
LJ -cohomology and the canonical homology of a J acobi manifold.
THEOREM 2.7.5. [115]. There is the pairing
(,) : H'
LP
(Z) x HT(Z) - m(Z),
<{tf},M>=W0>}.
96
CHAPTER 2.
GEOMETRY OF POISSON MANIFOLDS
of vector spaces of the LJ -cohomology and the canonical homology of a J acobi
manifold.
The result follows from the fact that, if tf T.(Z) and 0 is a basic form, then
their pairing (d, <j>) (1.2.30) is a basic function, and from the relation
wu{-d)\4> = d\6
ul
{<j>) + {-\)
r
6
w
(d\4>), <t>eO
T
B
(Z), tfeT
r
_,(Z).
Note that, bearing additionally in mind the map (2.2.4), one can apply the
construction of the LJ -cohomology and the canonical homology of a J acobi manifold
to contact manifolds as the particular case of J acobi manifolds. We refer to [38, 115]
for more details.
2. 8 Appendix. More brackets
We will give a brief survey of several extensions of the Poisson bracket to multivectors
and differential forms and its generalization to the bracket of n > 2 functions.
In this Section, we will return to the notation [., .]
SN
for the Schouten-Nijenhuis
bracket.
Let (Z, w) be a Poisson manifold. With the nilpotent operation iD (2.7.8), one
can introduce the bracket
{d,v}
w
d
=-[w{d),v]sN
on an exterior algebra T,(Z) of multivector fields on Z. This bracket has the property
[ti,v]
w
= -(-l)W
M
[v,i9]
w
+ w([#,v}sri),
and is a graded skew-commutative Lie bracket on the quotient T,(Z)/w(T.(Z)),
where the Lie degree of a multivector field i5 is | & | 1.
Remar k 2. 8. 1. Recall that multivector fields on a manifold Z constitute a graded
Lie algebra with respect to the Schouten-Nijenhuis bracket [.,.]SN' (1-2.15), where
the Lie degree of a multivector field <d is ] "& \ 1 (see Remark 1.2.3).
The Poisson bracket can be extended to exterior forms as follows.
If the Poisson bivector w on a manifold Z is non-degenerate, i.e., (Z, w) is a
symplectic manifold, there is the isomorphism ti)" (1.2.32) of graded algebras Q'(Z)
2.8. APPENDIX. MORE BRACKETS 97
and T,(Z). This isomorphism extends the Schouten-Nijenhuis bracket to the algebra
of exterior forms [133]:
{.,.}. :)
T
(Z)xO
s
(Z)^D
r+s
-
1
(Z),
to'Ctoff},,) =[u>'0Vff]sN, 4>,a eD'(Z).
(2.8.1)
(2.8.2)
In the general case of a Poisson manifold, we have the homomorphism w" (1.2.32)
which satisfies the property (2.7.11). Therefore, in order to construct the bracket
(2.8.1), let us consider the codifferential operator 6
W
(2.7.3). Then the Schouten-
Nijenhuis bracket of exterior forms (2.8.1) is defined as
{4>, }w =( M ) A o + (-l )W^A {6
w
o) - 6
W
{4> A o)
(2.8.3)
[102, 181]. This bracket has the properties
(_l)W(W-){0, {a,6}
w
}
w
+(-l)M(W-
{ffi
{<?,0U* +
Exampl e 2.8.2. In particular, the bracket (2.8.3) of 1-forms reads
{0, cr}; = L,i
0
a - L
wP(7
0 - d(io(tf>, a)) =
(2.8.4)
It provides D
1
(Z) with a Lie algebra structure such that
w* : Q\Z) -> T{Z)
is a Lie algebra homomorphism (2.8.2). The relationship between the bracket (2.8.4)
and the Poisson bracket (2.3.1) is
{df,dg}
w
= d{f,g}.

Using the nilpotency properties (2.7.4) and (2.7.5) of the codifferential 6
W
, one
can obtain the formula
d{<j>,a}
w
= -{d<t>,o}
w
- (-l)
w
{<j>,da}
w
.
w
l
<f>\do - w'a^dcj) + d(w((/>,a)).
{*,o-} = (-l)
l
*
l
" " {^ *},
C-1)
WI
"
M)
{.{.^}.} =0-
98 CHAPTER 2.
GEOMETRY OF POISSON MANIFOLDS
Then the bracket
{4>, o}d = -{dct>, a}
w
(2.8.5)
on an algebra of exterior forms D'{Z) can be introduced [21, 133]. This is a graded
skew-commutative Lie bracket
{*,*}<= ( - i )
w
" W}. i
on the quotient D*(Z)/dD*(Z), where the Lie degree of an exterior form<f> is | <j> | 1.
Let us turn now to an n-ary generalization of Poisson and J acobi brackets.
DEFI NI TI ON 2.8.1. [84, 85). A generalized almost Jacobi bracket of order n on a
manifold Z is said to be an n-linear map
{...} : xD{Z)^0(Z) (2.8.6)
which is
skew-symmetric
| . . . , ji,..., jj,... | \..., Jj,..., Ji,... I
with respect to any pair of arguments,
and is a first order linear differential operator on Z with respect to each argu-
ment, i.e.,
{Si/i, } =9i{fx, } +/i{ffi, } - Si /i {l , -..}.

If {...} is a generalized almost J acobi bracket of order n, then there exist an
n-vector field w and an (n 1)-vector field e on a manifold Z such that
{/i, -,/} = w(df
u
...,df
n
)+ E(- l )'
+
7.e(rf /i ,- ,d/...,d/
n
)
t =i
(2.8.7)
and
{l ,/l ,- ,/n- l }=e(d/
1
,...,d/n- l ).
2.8. APPENDIX. MORE BRACKETS
99
Conversely, any pair (w,e) of an n-vector field w and an (n - l)-vector field e
defines the generalized almost J acobi bracket (2.8.6) given by the relation (2.8.7). A
manifold Z provided with the bracket (2.8.7) is called a generalized almost Jacobi
manifold. If e =0, we obtain a generalized almost Poisson manifold.
In order to reproduce a J acobi structure in the case of n =2, one should add the
integrability conditions, generalizing the J acobi identity for the generalized almost
J acobi structure. In fact, two different types of integrability conditions have been
suggested. They are
the fundamental identity
{/l. i /n- l ,{f f l ,- .- ,0n}} = {{fl, , fn-l, 9l}, 92, , 9n) +
{9\,{fl,--,fn-l,92},--,9n} +
+ {Sl , - - , Pn- l , {/ l , - - , / n- l , 5n}}
and the generalized Jacobi identity
{/l> i /n-l i {In,- )/2n-l}} +
{/2. ,/n, {fn+l, ,/bn-l i /l }} H =0.
(2.8.9)
(2.8.8)
The bracket (2.8.7) satisfying the fundamental identity (2.8.8) is called a Nambu-
Jacobi bracket and (Z;w,e) is a Nambu-Jacobi manifold [85, 128]. If e = 0, the
bracket (2.8.7) is said to be a Nambu-Poisson bracket, and (Z, w) is a Nambu-
Poisson manifold [9, 84, 175].
The bracket (2.8.7) satisfying the generalized J acobi identity (2.8.9) is called a
generalized Jacobi bracket, and (Z;w,e) is a generalized Jacobi manifold [84, 145].
If e = 0, the bracket (2.8.7) is said to be a generalized Poisson bracket, and (Z, w)
is a generalized Poisson manifold [8, 9, 84].
The relationship between Nambu-J acobi manifolds and generalized J acobi man-
ifolds can be shown directly by noticing that the generalized J acobi identity (2.8.9)
is the full antisymmetrization of the fundamental identity (2.8.8). It follows that
every Nambu-J acobi manifold is a generalized J acobi manifold [9, 85].
For n even, the integrability conditions of a generalized J acobi structure (w, e)
are given by the following assertion [145].
PROPOSI TI ON 2.8.2. A generalized almost J acobi structure (w,e) of even order
n =2p is a generalized J acobi structure if and only if the relations
[w, e]
S
N = 0, [w, W]SN =2(2p - l J mA e,
100 CHAPTER 2.
GEOMETRY OF POISSON MANIFOLDS
generalizing the conditions (2.1.2), hold good.
At the same time, the integrability conditions of a Nambu-J acobi structure {w, e)
in particular imply that
the multivector field w is a Nambu-Poisson structure of order n,
the multivector field e is a Nambu-Poisson structure of order n 1 [128],
and every Nambu-Poisson multivector w can be written as an exterior product
of vector fields [79].
Exa mple 2.8.3. Let us consider the product Z = R x T'M with the holonomic
coordinates (*,<?', p,) as in Examples 2.2.1, 2.2.3 and 2.3.7. It is equipped with both
the J acobi structure given by the pair (wi,Ei) (2.2.9) and the Poisson structure w?
(2.3.9). Then Z is also provided with the generalized J acobi structure of order 4
given by the multivector fields
W = W\ A W2, e = E Aui2
(see [85] for details).
2.9 Appendix. Multisymplectic structures
We will give a brief exposition of multisymplectic and vector-valued generalizations
of a symplectic structure.
Similarly to that a Poisson bivector field is replaced by a multivector one for a
generalized Poisson structure or a Nambu-Poisson structure, one can consider an
n-ary generalization of a symplectic form. This is a multisymplectic form.
Given an m-dimensional manifold M with coordinates (z
x
), let us consider the
fibre bundle
C.AT'M^M
whose sections are exterior /t-forms on M. This fibre bundle is equipped with the
holonomic coordinates (Z
A
,PA), where A = (Aj <... < \
k
) are multi-indices of the
2.9.
APPENDIX. MULTISYMPLECTIC STRUCTURES 101
length | A |=k. The manifold AT'M is provided with the canonical exterior fc-form
0 defined by the relation
uJ ...u
1
je(p) = TC(ti)J ...TC(u
1
)J p
1
p AT'M, u, eT
p
(AT'M).
Its coordinate expression is
e = (2.9.1)
where the sum is over all multi-indices A of the length k.
The exterior differential dQ of the form (2.9.1) is the (k + l)-symplectic form
Q
M
= de = (2.9.2)
which belongs to the class of multisymplectic forms in the sense of Martin [131]. If
k = 1, the form Q.M (2.9.2) is the familiar symplectic form on the cotangent bundle
T'M.
Exampl e 2.9.1. For instance, let Y X be a fibre bundle over a (1 < n)-
dimensional base X. Let us consider the canonical form 9 (2.9.1) on the exterior
product A T'Y. The homogeneous Legendre bundle over Y is said to be the fibre
bundle
Z
Y
=T'Y ACAT'X), (2.9.3)
equipped with the holonomic coordinates (x
x
,y
i
,p
x
y
p), where the coordinate p has
the transformation law
' -() (' -$4
The canonical bundle monomorphism over Y
i
z
:T'YA{
n
/\T'X)^AT'Y
yields the pull-back form
5 =i"
z
@ =pcj + p
x
dy
x
A uix
(2.9.4)
on the fibre bundle Z
Y
, where we use the notation (1.2.21). Its exterior differential
Q
z
= dp A u + dp
x
A dy' A u>x
(2.9.5)
A
A--- Adz
Xk
,
J^dp\
1
\
k
Adz
x
> A---Adz
x
*
A
102
CHAPTER 2.
GEOMETRY OF POISSON MANIFOLDS
is also called the multisymplectic form [63]. In particular, if Y R, the multisym-
plectic form (2.9.5) leads to the canonical symplectic form (2.4.4) on the cotangent
bundle Z
Y
=T'Y.
Let us touch on another generalization of a symplectic structure when an R
n
-
valued 2-form is considered [53, 142]. Let X be an n-dimensional manifold and
LX X the principal frame bundle coordinated by (x
x
,s'
i
a
). The frame bundle
LX is provided with the canonical Revalued 1-form6
LX
(1.5.13) which reads
OLX = s%dz" <g>t
a
,
where {t
a
} is a basis for R
n
. Let us take its exterior differential
d0
LX
= ds% A dx" <8>t
a
,
called the n-symplectic form. It is readily observed that this form is non-degenerate,
i.e., given a vector field
u =u*^+uld
on LX, the equality u\d0Lx = 0 holds if and only if u = 0. By analogy with the
symplectic case, one can write
Uf\dO
LX
= -df, (2.9.6)
where / is an R
n
-valued function on the frame bundle LX. There is the class
of functions (e.g., which are constant on the fibres of LX) such that vector fields
satisfying the equation (2.9.6) exist. Then their J acobi bracket
{f,g} = Uf\dg
is defined, and it also belongs to the above-mentioned class of functions.
Let now Y >X be an arbitrary fibre bundle. The Legendre bundle over Y is
said to be the fibre bundle
n =AT*XTX VY aVY A (Vrx),
Y Y
(2.9.7)
equipped with the holonomic coordinates (x
x
,y',p*) possessing the transition func-
tions
Pi =
(2.9.8)
<V dx'
x
dy
1
* dx
6t
{)*
2.9. APPENDIX. MULTISYMPLECTIC STRUCTURES 103
It is provided with the TX-valued form
A = dp
x
A dy
{
A u <g> d
x
, (2.9.9)
called the polysymplectic form. The Legendre bundle (2.9.7) plays the role of a
finite-dimensional phase space in the Hamiltonian formulation of classical field the-
ory [28, 56, 57, 73, 96, 158, 159].
In the case of X = K, the polysymplectic Hamiltonian formalism leads us to the
Hamiltonian formulation of time-dependent mechanics (see Chapter 5).
The main peculiarity of polysymplectic formalism is that Hamiltonian connec-
tions on the Legendre bundle U * X play a role similar to Hamiltonian vector fields
on a symplectic manifold. A connection
T = da* {d
x
+ -$& + '&%)
on the fibre bundle n X is said to be locally Hamiltonian if the exterior form
7J A is closed (see Section 3.8 for details).
Remar k 2.9.2. Note that generalization of the Poisson bracket for polysymplectic
manifolds meets the difficulty (see [91, 92] and references therein). Nevertheless, in
the particular case of an affine bundle Y >X whose base X is provided with a
non-degenerate metric g, the bracket of horizontal 1-forms on n > X,
u. i at) (94>a dap da
0
d4>
a
\
can be globally defined.
>/l7Ifa*.
<j> = <t>
x
{p)dx
x
, a = <rx{p)dx
x
,
dpf dy
{
dp? dy'
9
a0
Cha pt er 3
Ha mi l t oni a n syst ems
The theory of Hamiltonian systems is a vast subject which can be studied from many
different viewpoints. This Chapter provides a brief exposition of different types of
Hamiltonian systems on Poisson, symplectic and presymplectic manifolds in conser-
vative mechanics. We leave aside many interesting constructions of a Hamiltonian
conservative mechanics and its straightforward extension to the direct product
[time] x [autonomous phase space]
that cannot be applied to time-dependent mechanics because a Hamiltonian is not
a function under time-dependent transformations. Nevertheless, any object denned
on a phase space V of conservative mechanics or on the product R x Z has the
counterpart on a phase space
ifr-.W = Rx Z (3.0.1)
of time-dependent mechanics, where different trivializations tp (3.0.1) correspond to
different reference frames.
For instance, there is one-to-one correspondence between the connections on a
fibre bundle W * R (3.0.1) and the time-dependent vector fields
1 x Z -^TZ
z
because of the diffeomorphism
J V =J*W = J\W xZ)=RxTZ.
However, this correspondence is not canonical, but depends on a trivialization
(3.0.1).
105
106 CHAPTER 3.
HAMILTONIAN SYSTEMS
Note that Hamiltonian relativistic mechanics can be seen as an autonomous
Hamiltonian system on the hyperboloid of relativistic momenta, i.e., it exemplifies a
Dirac constraint system (see Section 6.3). In this Chapter, we pay special attention
to constraint systems, including symplectic Hamiltonian systems, Dirac Hamiltonian
systems, and Dirac constraint systems, seen as particular presymplectic Hamiltonian
systems.
3.1 Dynamic equations
There are several types of equations in conservative mechanics. These are first
and second order dynamic equations, Hamilton equations with respect to Poisson,
symplectic or presymplectic structures, and Lagrange equations. The relationship
between them is the following.
The dynamic equations, by definition, are differential equations which can be
algebraically solved for the highest order derivatives.
The second order dynamic equations on a manifold M, by definition, are
particular first order dynamic equations on the tangent bundle TM of M.
The first order dynamic equations on a manifold Z can be represented as
Lagrange equations for Lagrangians on the tangent bundle TZ of Z, but they
are not Hamilton equations on Z in general.
The Hamilton equations are not necessarily first order dynamic equations.
The Lagrange equations on a manifold M are derived as particular Hamilton
equations on the tangent bundle TM, but they are not necessarily second
order dynamic equations and differential equations in any strict sense.
The second order dynamic equations on a manifold M fail to be represented as
Lagrange equations for Lagrangians on the tangent bundle TM or as Hamilton
equations on TM in general.
This Section is devoted to the notions of first and second order dynamic equations
on a manifold, called autonomous dynamic equations.
3.1.
DYNAMIC EQUATIONS 107
DEFI NI TI ON 3.1.1. Let Z, dimZ > 1, be a manifold, coordinated by (z
x
), and u a
vector field on Z. The closed subbundle u(Z) of the tangent bundle TZ, given by
the coordinate relations
i
A
= u
x
(z), (3.1.1)
is said to be an (autonomous) .first order dynamic equation on a manifold Z. This
is a system of first order differential equations on the fibre bundle K x Z >R in
accordance with Definition 1.3.3 (see Example 5.2.1).
By solutions of the first order dynamic equation (3.1.1) are meant integral curves
of the vector field u.
Remar k 3. 1. 1. Generalizing Definition 3.1.1, we will say that an (autonomous)
first order differentia] equation on a manifold Z is a closed submanifold < of the
tangent bundle TZ of Z. A solution of this differential equation is an integral curve
of a vector field on a submanifold J V cZ, which takes its values into TN n <.
In this Chapter, we will deal with the first order dynamic equations (3.1.1) when
vector fields u are Hamiltonian vector fields for Poisson, symplectic or presymplectic
structure on a manifold Z, treated as a momentum phase space of conservative
mechanics. However, they are not equivalent to the Hamilton equations in general.
The Hamilton equations with respect to a Poisson structure, by definition, are first
order dynamic equations, and so are the Hamilton equations with respect to a
symplectic structure. This is not the case of Hamilton equations with respect to a
presymplectic structure.
Let us bear in mind the well-known inverse problem in conservative Hamiltonian
mechanics, which consists in trying to represent a first order dynamic equation on a
manifold Z as Hamilton equations with respect to some Poisson structure on Z [32]
(see the next Section). The important motivation of the inverse problem is that,
with a Poisson structure, one may find integrals of motion of a dynamic equation
and also quantize it.
Rema r k 3. 1. 2. As in the particular case of the general result [76, 81, 156], let us
point out that every first order dynamic equation on a manifold Z can be represented
as Lagrange equations for a Lagrangian on the tangent bundle TZ, but not in an
explicit form.
108 CHAPTER 3.
HAMILTONIAN SYSTEMS
Let M be a manifold, coordinated by (q'). If it is a configuration space of
conservative mechanics, the corresponding velocity phase space is the tangent bundle
TM of M, while the cotangent bundle T'M plays the role of the momentum phase
space.
DEFI NI TI ON 3.1.2. An autonomous second order dynamic equation on a manifold
M is denned as a first order dynamic equation (3.1.1) on the tangent bundle TM,
which is associated with a hoionomic vector field
E = q% + E
i
(q^^)d
l
(3.1.2)
on TM. This vector field, by definition, obeys the condition
/(=) = UTM,
where J is the endomorphism (1.2.38) and UTM is the Liouville vector field (1.2.6)
on TM (see the notation (1.2.4)).
The vector field (3.1.2) is called a dynamic vector Geld. In the literature, it is
often termed a second order dynamic equation.
Let the double tangent bundle TTM be provided with the coordinates
(q
l
,?,$,?)
(3.1.3)
With respect to these coordinates, the second order dynamic equation defined by
the hoionomic vector field 5 (3.1.2) reads
q' = g', ? = =?(?,?). (3.1.4)
By solutions of the second order dynamic equation (3.1.4) are meant the curves
c : () M in a manifold M whose tangent prolongations c : () TM are
integral curves of the hoionomic vector field H or, equivalently, whose second order
tangent prolongations c live in the subbundle (3.1.4). They satisfy the second order
differential equations
c' (t) = S(c' (t),c' (t)).
A particular case of second order dynamic equations on a manifold M is that of
geodesic equations on the tangent bundle TM.
3.1.
DYNAMIC EQUATIONS 109
Given a connection
K = dq> (dj + K)di)
on the tangent bundle TM >M, let
K :TMxTM - TTM
M
(3.1.5)
be the corresponding linear bundle morphism over TM which splits the exact se-
quence
0 V
M
TM >-+ TTM >TM xTM * 0.
DEFI NI TI ON 3.1.3. A geodesic equation on TM with respect to the connection tf
is defined as the image
q' = <f,
? = K)q>
(3.1.6)
of the morphism (3.1.5) restricted to the diagonal TM C TM x TM. D
By a solution of a geodesic equation on TM is meant a geodesic curve c in M,
whose tangent prolongation c is an integral section (a geodesic vector Held) over
c C M for the connection K. The geodesic equation (3.1.6) can be written in the
form
fdtf = Ktf,
where q'{qi) is a geodesic vector field (which exists at least on a geodesic curve),
while q'di is the formal operator of differentiation (along a curve).
It is readily observed that the morphism K \TM is a holonomic vector field
on TM. It follows that any geodesic equation (3.1.5) on TM is a second order
equation on M. The converse is not true in general. Nevertheless, we have the
following theorem.
THEOREM 3.1.4. [136]. Every second order dynamic equation (3.1.4) on a manifold
M defines a connection K-= on the tangent bundle TM > M whose components are
K) = \60.
(3.1.7)
D
no
CHAPTER 3.
HAMILTONIAN SYSTEMS
However, the second order dynamic equation (3.1.4) fails to be a geodesic equa-
tion with respect to the connection (3.1.7) in general. In particular, the geodesic
equation (3.1.6) with respect to a connection K determines the connection (3.1.7)
on TM M which does not necessarily coincide with K. A second order equation
E on M is a geodesic equation for the connection (3.1.7) if and only if E is a spray,
i.e.,
[u
TM
, E] = =.,
where UTM is the Liouville vector field (2.4.9) on TM, i.e.,
H* =ayrywy.
In Section 4.3, we will improve Theorem 3.1.4 (see Proposition 4.3.3 below).
An extensive literature is devoted to the inverse problem for second order dy-
namic equations in Lagrangian mechanics. In Section 4.9, we will touch on the
following two aspects of this problem in time-dependent mechanics:
the condition for an Euler-Lagrange type operator to be an Euler-Lagrange
one as the particular inverse problem in field theory,
and the corresponding relation between Newtonian and Lagrangian systems.
We refer the reader to [136] for a survey on the inverse problem in conservative
Lagrangian mechanics. Recall that, in accordance with Remark 3.1.2, a second
order dynamic equation on a manifold M, being a first order dynamic equation on
TM, can always be thought of as Lagrange equations for a Lagrangian on the double
tangent bundle TTM of M.
In this Chapter, we will give two examples of the inverse problem for second
order dynamic equations in conservative Hamiltonian mechanics.
As is well known, the tangent bundle TM fails to possess any canonical Pois-
son, symplectic, or presymplectic structures. Nevertheless, there are proce-
dures to study a second order dynamic equation as the Hamilton ones with
respect to some Poisson structure on TM (see Remark 3.2.3 below).
Lagrange equations for a Lagrangian on the tangent bundle TM can be also
seen as Hamilton equations with respect to the presymplectic structure (or
the symplectic structure, if is regular), defined by this Lagrangian on TM
(see Examples 3.3.2 and 3.4.2 below).
3.2. POISSON HAMILTONIAN SYSTEMS 111
3.2 Poisson Hamiltonian systems
Let (Z, w) be a A:-dimensional Poisson manifold and H a real function on Z.
DEFI NI TI ON 3.2.1. A Poisson Hamiltonian system (w,H) on a manifold Z for an
(autonomous) Hamiltonian H with respect to the Poisson structure w is the set
SH= \J{VT
Z
Z:V- w\dH){z) =0}.
26Z
(3.2.1)
a
A solution of the Hamiltonian system (3.2.1) is a vector field tfona submanifold
N C Z, which takes its values into TN n Sa-
lt is readily observed that the Poisson Hamiltonian system S
w
(3.2.1) has a
unique solution which is the Hamiltonian vector field
&H = w
l
{dH) (3.2.2)
for the Hamiltonian H, that passes through any point of Sn- Hence, 5 is a first
order equation in accordance with Definition 3.1.1. It is called the Hamilton equation
for the Hamiltonian H with respect to the Poisson structure w.
With respect to the local canonical coordinates {q',Pi,z
A
) (2.3.7) for the Poisson
structure w, the Hamilton equation (3.2.1) reads
while the vector field $n (3.2.2) takes the form
&-H = PHdi - diHdK
Its integral curves r : () Z satisfy the equations
r* = d*n o r, U = diH o r, f
A
= 0. (3.2.3)
Recall that, given a dynamic equation (3.1.1) associated with a vector field u on
a manifold Z, the Lie derivative L
u
of a function f on Z along u can be treated as
the evolution of / along solutions of this dynamic equation. If
Uf = 0,
(3.2.4)
the function / is constant on any solution of the above-mentioned dynamic equation
or equivalently on integral curves of u. A function / ^const, is called a first integral
g' =dm, pj = -djH, z
A
=0,
112 CHAPTER 3.
HAMILTONIAN SYSTEMS
of motion if it obeys the equation (3.2.4). Prom the geometric viewpoint, this means
that the vector field u is tangent to the level surfaces of the function / at its regular
points, where df = 0.
If a dynamic equation is a Hamilton equation with respect to a Poisson structure,
one can find its first integrals of motion as functions in involution with a Hamiltonian
as follows.
Let H be a Hamiltonian of a Poisson Hamiltonian system (w,H). The Lie
derivative of an arbitrary real function / on Z along the Hamiltonian vector field
tin for H reads
UJ = $H\d}={HJ}.
(3.2.5)
The equality (3.2.5) is called the evolution equation. Substituting solutions r of the
Hamilton equations (3.2.3) in (3.2.5), we obtain the evolution
d
t
(for) = {H,f}or
of a function / along the integral curves of the Hamiltonian vector field d-^. In
particular, if
{HJ} = 0,
the function / is a first integral of motion of the Hamiltonian system (w,H).
Exampl e 3.2.1. A Hamiltonian "H itself is obviously a first integral of motion of
the Hamiltonian system (tu,W).
It is easily seen that, if / and / ' are first integrals of motion of a Poisson Ha-
miltonian system, so is their Poisson bracket {/,/'}. It follows that first integrals
of motions constitute a Lie algebra.
A vector field v on Z is called an infinitesimal symmetry of a Hamiltonian H if
KH =0.
In particular, if / is a first integral of motion of a Poisson Hamiltonian system
(w,H), the Hamiltonian vector field $/ for / is an infinitesimal symmetry of the
Hamiltonian H, i.e.,
U,n = 4
f
\dH = -{H,f}=0.
3.2.
POISSON HAMILTONIAN SYSTEMS 113
Let us turn now to the inverse problem in conservative Hamiltonian mechanics,
mentioned in Remark 2.3.5. It has the following solution.
PROPOSI TI ON 3.2.2. For any first order dynamic equation (3.1.1) on a manifold Z,
there exists locally a generating Poisson Hamiltonian system (w, H) on Z such that
(3.1.1) is locally a Hamilton equation.
Proof. If u = 0, the statement is obvious. Let u{z) / 0 at a point z Z. There
exists a coordinate system (q
1
,. .. , q
k
) on an open neighbourhood of z such that
U =
W
and [u
'
v] = 0
'
v =
w
Then u A v is a local Poisson bivector field of rank 2 (see Example 2.3.2). It is
readily observed that u is locally a Hamiltonian vector field for the function q
2
{z
x
)
with respect to the Poisson structure uA. QED
Proposition 3.2.2 leads us to the conditions for the inverse problem of a first
order dynamic equation (3.1.1) to have a global solution.
PROPOSI TI ON 3.2.3. [59, 82]. Let u be a nowhere vanishing vector field on a
manifold Z. If there exist
a nowhere vanishing vector field v on Z such that [u, v] = 0 everywhere on Z,
and a function f on Z such that L
u
/ =0 and L / =1,
then u is the Hamiltonian vector field for the function / with respect to the Poisson
structure uA.
Exampl e 3.2.2. Every first order dynamic equation (3.1.1) on a manifold Z is a
Hamilton equation on the product RxZ with coordinates (t, z
x
) for the Hamiltonian
Ti = t and with respect to the Poisson structure w = u A d
t
. We refer the reader to
[59] for the physically interesting examples.
Remark 3.2.3. The method based on Propositions 3.2.2 and 3.2.3 can be applied
to study the existence of a generating Poisson structure for a second order dynamic
equation (3.1.4) on a manifold M defined by a holonomic vector field E (3.1.2)
on the tangent bundle TM. In particular, for any second order dynamic equation
114 CHAPTER 3.
HAMILTONIAN SYSTEMS
(3.1.4) on M, there exists locally a generating Poisson Hamiltonian system {w,H)
on TM such that the associated holonomic vector field S is the Hamiltonian vector
field for the Hamiltonian H with respect to the Poisson structure w. The necessary
condition for a Poisson structure w on the tangent bundle TM to be a generating
Poisson structure for a holonomic vector field 5 on TM is
L=(w) = [E,tu] =0.
We refer the reader to [182] for a detailed analysis of this condition.
3.3 Sympl ecti c Hami l toni an systems
Let (Z, fi) be a symplectic manifold. The notion of a symplectic Hamiltonian system
is a repetition of Definition 3.2.1 (see [127]).
DEFI NI TI ON 3.3.1. A symplectic Hamiltonian system (ft.H) on a manifold Z for a
Hamiltonian H with respect to the symplectic structure fi is the set
Sn = \J{veT
z
Z: v\n + dH(z) = 0}.
2Z
(3.3.1)

As in the general case of Poisson systems, the symplectic Hamiltonian system
(fi, H) has a unique solution which is the Hamiltonian vector field tf
K
such that
V | fi =-dH
for the Hamiltonian H, that passes through any point of Sn- Hence, S-H is a first
order dynamic equation, called the Hamilton equation.
With respect to the local canonical coordinates {q\pi) for the symplectic struc-
ture fi, the Hamilton equation (3.3.1) reads
q* = d
x
H, pj = -djTi.
(3.3.2)
In the symplectic case, there is one-to-one correspondence between the Hamil-
tonian systems and the Hamiltonian vector fields on a symplectic manifold (Z, f2).
The first integrals of motion of a symplectic Hamiltonian system and infinitesimal
symmetries of a Hamiltonian are defined as a repetition of those for a Poisson
Hamiltonian system.
3.3. SYMPLECTIC HAMILTONIAN SYSTEMS 115
Recall that a Hamiltonian system on a 2m-dimensional symplectic manifold is
called completely integrable if there exist m first integrals of motion which are
pairwise in involution and whose differentials are linearly independent on a dense
open subset of that manifold.
Exampl e 3.3.1. Let Z = T*M be a symplectic manifold provided with the canon-
ical symplectic form fi (2.4.2). In accordance with Example 2.4.7, any Hamiltonian
system (ft, H) on the symplectic manifold (T'M, fi) is the Lagrangian submanifold
Sn =$K{T'M) (2.4.10) of the symplectic manifold TT'M equipped with the sym-
plectic form Q (2.4.9). The generating function of this submanifold, given by the
coordinate relations
q' = &n, pi = -aoi,
(3.3.3)
is the Hamiltonian 7i(q
x
,pi). Any Hamiltonian H on the symplectic manifold T'M
defines the fibred morphism over M
H
d
=TT.TM o a o $
H
: T'M -^TT'M ->T'TM - TM,
q
i
o
;
R=d
,
n,
(3.3.4)
called the Hamiltonian map.
Exampl e 3.3.2. Any non-degenerate autonomous Lagrangian system can be seen
as a symplectic Hamiltonian system as follows.
An autonomous Lagrangian is defined as a real function C on the tangent bundle
TM of an event space M. Throughout this Chapter, we will use the compact
notation
7Ti = d
{
C, Kji djdiC.
Using the tangent-valued form 4>j (1.2.39) on the tangent bundle TM, let us intro-
duce the Poj'ncare-Cartan 1-form
He =4>j\d = n
t
dq\
Its differential
tie =dn
t
A dq
i
= K^dc? A dq' + djitidq
1
A dq'
(3.3.5)
is a symplectic form on TM if and only if the Lagrangian C is regular, i.e., its
Hessian det (TT^) is nowhere vanishing.
116 CHAPTER 3.
HAMILTONIAN SYSTEMS
If a Lagrangian C is regular, one can consider the symplectic Hamiltonian systen
(Qci Ec) with respect to the symplectic formQc for the Hamiltonian
Ec = v-TMJdC C = q'lTi ,
(3.3.6)
called the energy function, where we make use of the Liouville vector field UTM
(1.2.6) on TM. Then we come to the Hamilton equation for the Hamiltonian vector
field dc such that
ticltoL = ~dE
c
.
(3.3.7)
With respect to the coordinates (3.1.3) on the double tangent bundle TTM, the
equation (3.3.7), called the Cartan equation for the Lagrangian , reads
T(q* - ?') = o,
(3.3.8a)
(3.3.8b)
Since the Lagrangian C is regular, the equation (3.3.8a) reduces to the equality
q' = <?',
(3. 3. 9)
while the equation (3.3.8b) comes to the Lagrange equations
d,C (pifji qidj-Ki = 0 (3.3.10)
for the Lagrangian .
The condition (3.3.9) means that, in the case of a regular Lagrangian , the
Hamiltonian vector field dc is holonomic. It defines the second order dynamic
equation on TM which is equivalent to the Lagrange equations (3.3.10).
Let us consider the fibre bundles TT'M and T'TM provided with the coor-
dinates (q',Pi,q',Pi) and (q',q', (?,, <ji), respectively, and recall the isomorphism a
(1.1.9):
a : TT'M S T'TM,
R
. ft, ft ^ q,.
The symplectic form fi

(3.3.5) yields the fibred morphism (1.2.33) over TM:


fij. : 7TM ->TTM,
q\ o n
b

= q%
y
,
diC - q'lTji - q
J
9j7r, + (q
J
- q')d
t
TX
]
= 0.
<7i o n
b

= -q>TT
:i
+ q
J
(3j 7rj - ^T T , ) ,
3.3.
SYMPLECTIC HAMILTONIAN SYSTEMS 117
and the corresponding fibred morphism
Q"
1
o Cl
L
: TTM ~> T'TM TT'M,
P, O a
_1
f^. = -q>TC
jx
+ q'idiTTj - djTT
t
), p, O Q
_ 1
O l
c
= q*7Ty.
(3.3.11)
We have the relation
ci
c
=-nn
T
- = -{a~
l
o n).yn,
where fi is the tangent lift (1.2.25) of Q,c onto TTM and by CIT-TM is meant the
canonical symplectic form on the cotangent bundle T'TM of TM.
It is readily observed that the system of Lagrange equations (3.3.10) is a Lagran-
gian submanifold of the symplectic manifold (TTM, &c)- Accordingly, the image of
the Lagrange equations (3.3.10) into TT'M by the morphism a
- 1
o l
c
(3.3.11) is
the Lagrangian submanifold
Pi = ^TTjn
pi = -d,C +tfdi-Kj
of the symplectic manifold (TT'M, fi). The generating function of this Lagrangian
submanifold is the energy function Ec (3.3.6).
Every Lagrangian on the configuration space TM yields the fibred morphism
over M

d
=ir
T
.
M
o Q"
1
o dC : TM T'TM ->TT'M -> T'M;
p, = diC,
(3.3.12)
(3.3.13)
called the Legendre map. A Lagrangian is regular if and only if the corresponding
Legendre map is a local diffeomorphism. A Lagrangian is called hyperregular
if the Legendre map is a diffeomorphism.
The tangent map TC to the Legendre map C is the fibred morphism
TC : TTM -> TT'M,
PioT = TTi, p, = ix
%J
q' + djiTiq>,
over M. Then the image of the Lagrange equations (3.3.10) into TT'M by TC is
another Lagrangian submanifold
Px = TTt,
Pi = di
(3.3.14)
fi

=- [d(- f Try +q*(flbi - djTT,)) A <tf +rf(q


J
'%) A <#)],
118 CHAPTER 3.
HAMILTONIAN SYSTEMS
of the symplectic manifold (TT'M, ft). This is exactly the Lagrangian submanifold
(2.4.11), defined by the differential dC (see Example 2.4.7). Its generating function
is C.
One can find the conditions of a hypen-egular Lagrangian on the tangent
bundle TM and a Hamiltonian H on the cotangent bundle T'M to define the same
Lagrangian submanifolds (3.3.14) and (3.3.3) of the symplectic manifold (TT'M, Q).
This takes place if and only if
u
T
-
M
\dH-H = CoH,
p
i
d
l
H-H = C(q\d
i
H),
(3.3.15)
where U
T
-M is the Liouville vector field (1.2.5) on T'M M, or equivalently
u
TM
\dC - = H o ,
q
i
d
i
C-C = H(q
i
,d
i
C),
where UTM is the Liouville vector field (1.2.6) on TM M. Taking the differential
of the equation (3.3.15), we find that
H = C~\
(diC) oH= -d
t
H.
Then we have the commutative diagram
TM - ^ TTM
C \ \ TC
T'M >TT' M

The constructions in Examples 3.3.1, 3.3.2 lead to description of Hamiltonian and
Lagrangian dynamics of autonomous systems in terms of Lagrangian submanifolds
and to constructing unified Lagrangian-Hamiltonian formalism on the phase space
TT'M [111, 176, 177] (see Example 3.4.2 below).
3.4 Presymplectic Hamiltonian systems
The notion of a Hamiltonian system is naturally extended to presymplectic manifolds
(see [12, 60, 138, 172]).
3.4. PRESYMPLECTIC HAMILTONIAN SYSTEMS 119
DEFI NI TI ON 3.4.1. Let Z be a fc-dimensional manifold equipped with a presym-
plectic form fi, and let H be a real function on Z. A presymplectic Ha.miltonia.ii
system for a Hamiltonian H, like (3.3.1), is the set
Sn
d
= \J{ve T
Z
Z : vjfi +dH(z) = 0}.
zez
(3.4.1)

A solution of the presymplectic Hamiltonian system (3.4.1) is a Hamiltonian
vector field d-u for H which lives in Sn- It satisfies the equality
tfwjfi = -<m. (3.4.2)
In comparison with the symplectic case, the form 1 is degenerate, the set (3.4.1)
fails to be a submanifold in general, and a solution of the equation (3.4.2) does not
necessarily exist everywhere on the manifold Z.
Note that, for any point z e Z, the fibre
{vT
z
Z : v\n + dH(z) =0} (3.4.3)
of the set S-H over z is an affine space modelled over the vector space
Ker
2
Q = {v G T
Z
Z : u\v\Q = 0, Vu 6 T
Z
Z}
which is the fibre over z of Kerfl (1.2.34) of the presymplectic form fi. The fibre
(3.4.3) may be empty.
PROPOSI TI ON 3.4.2. The equation
v\il + dH{z) = 0, v G T
Z
Z, (3.4.4)
has a solution only at the points of the subset
N
2
= {z G Z : u\<m{z) =0, Vu G Ker
2
Q} C Z,
(3.4.5)
i.e., where Ker
z
ft C Ker
z
dH [60, 138].
Proof. Let a vector v G T
Z
Z, satisfying the equation (3.4.4), exist. Then, the
contraction of the right-hand side of the equation (3.4.4) with an arbitrary element
u G Ker
z
Q leads to the equality u\dH(z) = 0. In order to prove the converse
assertion, it suffices to show that
dH(z) G Imn
b
,
120 CHAPTER 3.
HAMILTONIAN SYSTEMS
that results from the inclusions
dft(z) Ann(Ker dH{z)) C Ann(Ker
2
fi) = Imfi*.
QED
It follows that, if Z ^ N
2
a. presymplectic Hamiltonian system is not a differential
equation on Z. If N
2
=Z, it is a differential equation, but not a dynamic equation.
From now on, let us suppose that a presymplectic form f2 is of constant rank, and
that N
2
(3.4.5) is a submanifold of Z, but not necessarily connected. Then, by virtue
of Proposition 1.1.4, the kernel Kerfi is a closed vector subbundle (moreover, an
involutive distribution) of the tangent bundle TZ. Accordingly, Sn \N
7
is an affine
bundle over N
2
. It admits a global section, but this section does not live necessarily
in TN
2
C TZ, i.e., is not a vector field on the submanifold N
2
in general.
To obtain a solution of the presymplectic Hamiltonian system (3.4.1), let us
consider a submanifold N C N
2
C Z such that
Sn\
N
nT
x
N^9,
Vze N,
or equivalently
dH(z) 6 Q\TN), Vze N.
If the morphism 0* \TN is of constant rank, then S-H H TN >N is an affine bundle
modelled over the vector bundle Ker Q n TN TV which may be 0. Sections of
the affine bundle S-H fl TN TV are Hamiltonian vector fields $x (3.4.2) for the
Hamiltonian Wona submanifold N. If such a submanifold exists, it may be found
by the following well-known algorithm [12].
Let us take the intersection
S
H
\
N2
nTN
2
and project it to Z. We obtain the subset
N
3
= K
Z
{SH\
N
,OTN
2
)CZ.
If N
3
is a submanifold, let us take the intersection
S
H
\
N3
nTN
3
.
Its projection to Z gives a subset N
4
C Z, and so on. Since a manifold Z is finite-
dimensional, the procedure is stopped after a finite number of steps by one of the
following results.
3.4. PRESYMPLECTIC HAMILTONIAN SYSTEMS 121
There is i > 2 such that a set TVj is empty. This means that a presymplectic
Hamiltonian system has no solution.
A set N,, i > 2, fails to be a submanifold. It follows that a solution may exist,
but not everywhere on A/j.
If N
i+
i = N
t
for some i > 2, this is the desired submanifold N. If the morphism
\TN is of constant rank, a solution of the presymplectic Hamiltonian system
(3.4.1) exists everywhere on N.
Remar k 3. 4. 1. Since N
2
^ Z in general, one can consider another variant of how
to solve a presymplectic Hamiltonian system. Let N
2
(3.4.5) be a submanifold of
Z. It can be provided with the pull-back presymplectic form Q
2
=i)v
2
^- Then the
pull-back presymplectic Hamiltonian system
SH, = U i
v
e T
Z
N
2
:v\Q
2
+ dH
2
(z) = 0}
zew
2
for the pull-back Hamiltonian H
2
= i'^Ti on ./V2 can be examined. As for the initial
systemS-H, we find that the equation
v\Q
2
+ dH
2
(z)=0, veT
z
N
2
,
has a solution on the subset
N*
i
= {zeN
i
: u\i'
N2
dn{z) =0, Vu Ker
2
ft
2
C TN
2
} C N
2
,
and so on. We obtain the chain of submanifolds
N
2
D N'
3
D N'
4

which is finite, and is stopped by one of the following results.
There is i > 2 such that a set N- is empty.
A set AT/, i > 2, fails to be a submanifold.
If N[
+l
= N? for some t > 2, this is a desired submanifold N' C Z. If the
presymplectic formi'
N
,Q, on N' is of constant rank, then there exists a solution
of the presymplectic Hamiltonian system (i*
N
,l, i^H) everywhere on N'.
122 CHAPTER 3. HAMILTONIAN SYSTEMS
It is readily observed that any solution $n C TN of the equation (3.4.2) on a
submanifold N C Z also obeys the equation
ti
H
\i'
N
{n + dH) = 0
and, consequently, is a solution of the pull-back presymplectic system on N. It
follows that N C N'.
Solutions of the presymplectic Hamiltonian system 5 (3.4.1) on a submanifold
N constitute an affine space modelled over the linear space of sections of the vector
bundle KerCl n TN N. If this vector bundle is not 0, different solutions of the
Hamilton equation (3.4.1) correspond to different first order dynamic equations on
./V. Therefore, a Hamilton equation with respect to a presymplectic structure on a
manifold Z is not equivalent to a first order dynamic equation on Z in general.
Sections of the vector bundle Kerfi Z are sometimes called gauge fields in
order to emphasize that, being solutions of the presymplectic Hamiltonian system
(fi,0) for the zero Hamiltonian, they do not contribute to a physical state, and are
responsible for the gauge freedom [12, 172]. At the same time, there are physically
interesting presymplectic Hamiltonian systems, e.g., in relativistic mechanics when
a Hamiltonian is equal to zero (see also Remark 4.8.9 below). In this case, Ker dH =
TZ and the Hamilton equation (3.4.4) has a solution everywhere on a manifold Z.
Another pull-back construction is connected with the above-mentioned gauge
freedom. Let a presymplectic form fi on manifold Z be of constant rank and let its
characteristic foliation be simple, i.e., its leaves are fibres of a fibre bundle n : Z P
over a symplectic manifold (P, $7/>) in accordance with Proposition 2.6.2. Then ft is
the pull-back ir'Vlp of a symplectic form ftp on the base P by this fibration. Let a
Hamiltonian H be the pull-back n'Hp of a function Tip on P. Then we have
Kerfi = VN C Ker dH,
and the presymplectic Hamiltonian system (f2, Ti) has solutions everywhere on the
manifold Z. Any such solution tin is projected onto a unique solution of the symplec-
tic Hamiltonian system (dp, Tip) on the manifold P, while gauge fields are vertical
vector fields on the fibre bundle Z P. This is the case of a gauge invariant
Hamiltonian system, where P is called a physical phase space.
In the case of a presymplectic Hamiltonian system on a manifold Z, this manifold
fails to be provided with a Poisson structure in general (see Remark 2.5.1), that is an
essential problem for quantization. Nevertheless, by virtue of Propositions 2.5.3 and
3.5.
DIRAC HAMILTONIAN SYSTEMS 123
2.5.4, every presymplectic form on a manifold Z can be represented as a pull-back of
a symplectic form by the coisotropic imbedding. It follows that each presymplectic
Hamiltonian system can be seen as a Dirac constraint system [26] that we will
investigate in Section 3.6.
Exa mple 3. 4. 2. Autonomous Lagrangian systems with constraints are often treated
as presymplectic Hamiltonian systems (see [30, 61, 112, 113, 138] and references
therein). Every Lagrangian on the tangent bundle TM of a configuration mani-
fold M provides TM with the presymplectic form Qc (3.3.5). Let us consider the
presymplectic Hamiltonian system(Qc, Ec) on TM for the Hamiltonian E
c
(3.3.6).
This is exactly the Cartan equation (3.3.8a) - (3.3.8b) for the Lagrangian . Its
restriction to the submanifold q' = q' (3.3.9) of the double tangent bundle TTM
of M leads to the Lagrange equations (3.3.10). This means that solutions of the
Lagrange equations are solutions of the Cartan equation which are dynamic vector
fields on TTM.
Conversely, a symplectic Hamiltonian system (fi, H) on the cotangent bundle
T'M can be seen as a degenerate Lagrangian system on the velocity phase space
TT'M of the momentum phase space T'M as follows. Put the Lagrangian
n = Piq' -U{q\pi)
(3.4.6)
on this velocity phase space. It is readily observed that the pre-image in TTT'M of
the Hamilton equation (3.3.2) on TT'M for the Hamiltonian H by the projection
TTT'M ->TT'M is exactly the Lagrange equations (3.3.10) on TTT'M for the
Lagrangian (3.4.6). It follows that they have the the same solutions r : () T'M.
It should be emphasized, that, in general, neither Cartan equations nor Lagrange
equations are differential equations in a strict sense. They are differential equations,
e.g., when the Hessian matrix 7iy, of is of constant rank.
3.5 Dirac Hamiltonian systems
There is an extensive literature devoted to autonomous mechanical systems with
constraints (see, e.g., [121, 127, 137, 166, 172] and references therein). In this and
the next Sections, we will deal only with autonomous Hamiltonian systems with con-
straints. We leave the constraints in time-dependent Lagrangian and Hamiltonian
mechanics for Chapters 4 and 5.
124 CHAPTER 3.
HAMILTONIAN SYSTEMS
Let (Z, fi) be a 2m-dimensional symplectic manifold and H a Hamiltonian on
Z. Let N be a (2m - n)-dimensional closed imbedded submanifold of Z, called the
primary constraint space or simply a constraint space. We aim to investigate the
following two problems:
(A) solutions of the Hamiltonian system (Q, H) on a manifold Z, which live in
the constraint space N,
(B) and the restriction of the Hamiltonian system (fi, H) to the constraint space
N.
Remark 3.5.1. The following local relations will be useful in the sequel. Let the
constraint space N be given locally by the equations
f
a
{z) = 0, o=l , . ..,n, (3.5.1)
where /
a
(z) are local functions on Z, called the primary constraints. Let us consider
the set
I
N
= Ken'
N
cQ{Z) (3.5.2)
of real functions f on Z which vanish everywhere on N, i.e., i'
N
f = 0. This is an
ideal of the associative commutative algebra 0(Z) such that
D(Z)/I
N
= 0(N)
(cf. (2.6.4)). Its elements / are written locally in the form
/ =
(3.5.3)
where g
a
are functions on Z. This means that the ideal //v is locally generated by
the constraints /. We will call {/
a
} a local basis for the ideal 1^.
Let dI
N
be the submodule of the D(Z)-module O'(Z), which is generated locally
by the exterior differentials df of functions f I
N
. Its elements are the finite sums
a = /, I
N
, 9' 0(Z).
By virtue of (3.5.3), they are given by local expressions
a=J2(9
a
dfa + fa4>
a
),
a =l
(3.5.4)
T.9'df
i
a =l
3.5. DIRAC HAMILTONIAN SYSTEMS 125
where g" are functions and 4>
a
are 1-forms on Z.
It should be emphasized that the expressions (3.5.3) and (3.5.4) are fulfilled
locally on a domain where the constraint space N is given by the equations (3.5.1).
The solution of problem (A) is obvious. It exists if a Hamiltonian vector field
d-H, restricted to the constraint space N, is tangent to N, i.e.,
SH\N = \J{veT
z
Z: v\Q + dH{z) =0} C TN. (3.5.5)
Then integral curves of the Hamiltonian vector field tin do not leave N.
The condition (3.5.5) is satisfied if and only if
{H, IN} C IN, (3.5.6)
i.e., if and only if the Hamiltonian H belongs to the normalizer I(N) (2.6.6) of the
ideal IN- With respect to the local basis {fa}, the condition (3.5.6) reads
#n\df
a
= (3.5.7)
where g
c
are functions on Z.
If the relation (3.5.6) does not hold, let us introduce the secondary constraints
{HJ
a
} = 0.
If the collection of primary and secondary constraints is not closed with respect to
the relation (3.5.7), let us add the tertiary constraints
{H,{H,f
a
}} = 0,
and so on. If a solution exists anywhere on N, the procedure is stopped after a finite
number of steps by constructing the complete system of constraints. The complete
system of constraints defines the final constraint space, which do not include the
points of Z where the Hamiltonian vector field i?
M
is transversal to the primary
constraint space N.
From the algebraic viewpoint, we find an ideal 7
nn
of D(Z) which is the minimal
extension of the ideal IN such that
{W, /fin} C /fin-
{ ,/}
n

126 CHAPTER 3. HAMILTONIAN SYSTEMS
In the framework of the algebraic theory of constraints, problem (A) can be
reformulated as follows.
Let AT be a closed imbedded submanifold of a symplectic manifold (Z, ft) and I
N
(3.5.2) the ideal of functions which vanish everywhere on N. All its elements are said
to be constraints. We aim to find a Hamiltonian, called an admissible HamiltoniaD,
on Z such that the symplectic Hamiltonian system (ft, H) has a solution everywhere
on the constraint space N. One can treat N as a final constraint space. Moreover,
we may take an ideal IN whether it is an ideal of functions vanishing on some
submanifold N C Z or not.
In accordance with the condition (3.5.6), any element of the normalizer I(N)
(2.6.6) is an admissible Hamiltonian.
Let us consider the intersection I'{N) = I(N) n I
N
(2.6.7). Its elements are
called the first-class constraints, while the remaining elements of IN axe the second-
class constraints. The first-class constraints make up a Poisson subalgebra of the
normalizer I(N) which, in turn, is a Poisson subalgebra of the Poisson algebra D{Z)
on the symplectic manifold (Z, ft) (see Section 2.6). Recall that I
N
C I'(N), i.e.,
the products of second-class constraints are also first-class constraints.
Exampl e 3.5.2. If N is a coisotropic submanifold of Z, then IN C I{N) and
I'{N) = I
N
- It follows that we have only the first-class constraints.
The admissible Hamiltonians which are not first-class constraints are the rep-
resentatives of the non-zero elements of the quotient I(N)/I'(N), which is the re-
duction of the Poisson algebra D(Z) via the ideal I
N
(see Definition 2.6.8). In
particular, let the presymplectic form i'
N
fi on N be of constant rank and its char-
acteristic foliation be simple, i.e., it defines a fibration 7r : N P over a symplectic
manifold P. In view of the isomorphism (2.6.8), one can think of elements of the
quotient I(N)/I'(N) as being the Hamiltonians on the physical phase space P. It
follows that the restriction of an admissible Hamiltonian H to the constraint space
N coincides with the pull-back onto N of some function / on P, i.e.,
i-
N
H = */, f e o(P).
Let us turn now to problem (B). Given a Hamiltonian Wonasymplectic manifold
(Z, Q) and a constraint space N C Z, a Dirac Hamiltonian system (ft, H, N) on the
constraint space N is defined as the subset
S.H = \J{ve T
Z
Z : i'
N
(v\n + dri(z)) = 0}
(3.5.8)
3.5. DIRAC HAMILTONIAN SYSTEMS 127
[127, 137]. Note that
SH\N C S.
n
.
Therefore, the set (3.5.8) is not empty.
We will say that a Dirac Hamiltonian system on N has a solution if there is a
vector field on a neighbourhood of each point z 6 /V, which lives into the subset S.
n
(3.5.8). For instance, every solution of problem (A) is also a solution of problem
(B).
Exampl e 3.5.3. Problem (B) has a complete solution in the germ terms in the
following case. Let Z = T'M be a symplectic manifold equipped with the canonical
symplectic form fi, and N its coisotropic submanifold. Then TN is a coisotropic
submanifold of the symplectic manifold TT'M equipped with the symplectic form
Q (2.4.9). Indeed, let utTTN and vTTT'M\TTN. There exists a vector field
T
U
on JV such that its canonical lift T
U
onto TN contains tt. There is also a vector
field T on T'M such that its canonical lift T
V
onto TT'M contains v. Let u and v
be projected over the same element q TN. By the virtue of the relation (1.2.23),
one can then write
v\u\a =n&}J f
tt
()J fi = Kvz)K(z)J ft), z = x
z
{q).
In the case of an arbitrary u 6 T
q
TN, this expression differs from 0 because there
exists a vector field T on a neighbourhood of z, which does not belong to Orth
n
TN
since N is coisotropic. It follows that Orth-TTN does not include elements of
TTT'M\TTN.
Let 7{ be a Hamiltonian on T'M. Then the symplectic Hamiltonian system Sn
(3.3.1) is a Lagrangian submanifold of the symplectic manifold (TT'M,fy (see Ex-
ample 2.4.7). A Dirac Hamiltonian system {Sl,H,N) on a (2m - n)-dimensional
coisotropic submanifold TV has a solution if the coisotropic submanifold TN C
TT'M and the Lagrangian submanifold S
w
are not transversal. The germ of any
such pair (TN,S\.) is symplectically equivalent to the pair
({,' =. =,* =0}, {q
a
= 8PS{p)) a = 1,.... 2m}} 0)) (q,p) eR
4m
,
where S is the germ of a function such that the rank of the tangent map to the
morphism
P- (d^S,.,.,d
2n
S)
128 CHAPTER 3.
HAMILTONIAN SYSTEMS
at 0 is not maximal. We refer the reader to [48] for a detailed classification of the
germs of such generating functions S.
A Dirac Hamiltonian system (Q, H, N) is called compieteiy integrable if there
exists its solution through each point of the set S.n- The obvious necessary condition
for a Dirac Hamiltonian system to be completely integrable is the inclusion S.n C
TN. In this case, a Dirac Hamiltonian system {Q, H, N) reduces to the symplectic
Hamiltonian system (i'
N
Q, J/yW) on the constraint space N.
Let us inspect the above-mentioned necessary condition. Note that, for each
point 2 G N, the fibre
{v e T
Z
Z : i'
N
(v\Sl + dH{z)) =0}
of the set S.n over z N is a non-empty affine space modelled over the vector space
{v e T
Z
Z : u\v\Q. =0, Vu T
Z
N} (3.5.9)
which is the fibre of the Q-orthogonal space OrthnTW (2.4.6) of TN. It follows that
OrthnT/V C TN. Therefore, a Dirac Hamiltonian system on the constraint space
N is completely integrable only if N is a coisotropic submanifold of the symplectic
manifold (Z, Q). Then we have
u\dH =0, Vu e OrthnN.
In order to formulate a sufficient condition for a Dirac Hamiltonian system to be
completely integrable, let us assume that the vector spaces (3.5.9) for all z G N are
of the same dimension. Then the fi-orthogonal OrthnTTV of TN is a vector bundle
over N, while the Dirac Hamiltonian system S.n ,/V is an affine bundle over ./V.
PROPOSI TI ON 3.5.1. [127], Let a submanifold i Vofa symplectic manifold (Z,f2)
be coisotropic. Then OrthnTW is an involutive distribution on N. If the exterior
differential dH is constant on the leaves of the corresponding foliation, then the
Dirac Hamiltonian system (3.5.8) is completely integrable.
Pr oof. Since the set S.n is not empty, it suffices to show that this set belongs to
TN. Let v S.
n
and z = TT
Z
(V). Whenever u Orth
n
TAf C TN over z, the
equality
v\u\Sl = u\dH = 0
3.6.
DIRAC CONSTRAINT SYSTEMS 129
holds. It follows that v T
Z
N. Since S. C TN, there is a vector field d on N
through any point of the Dirac Hamiltonian system S.. Such a solution is not
unique. If & is the above-mentioned vector field, then any vector field d + v where v
lives in Oxia
n
TN C TN is also a solution of the Dirac Hamiltonian system. QED
Thus, one should consider completely integrable Dirac Hamiltonian systems on
coisotropic submanifolds N. In this case, a Dirac Hamiltonian system (fl, H, N)
reduces to the presymplectic Hamiltonian system on the constraint space N for the
pull-back Hamiltonian i'
N
H with respect to the pull-back presymplectic form i'
N
Q.
This is the case of a Dirac constraint system.
3.6 Dirac constraint systems
Let (Z, fi) be a 2m-dimensional symplectic manifold, and let Nbea closed imbedded
submanifold of Z. Let H be a Hamiltonian on Z.
DEFI NI TI ON 3.6.1. A Dirac constraint system on the constraint space N is the set
SN-H \J{v e T
Z
N : v\i%(fl + dH{z)) =0}. (3.6.1)
D
In particular, we are dealing with Dirac constraint systems when a Hamiltonian
description of degenerate autonomous Lagrangian systems is investigated (see, e.g.,
[58, 172] and references therein). They are called generalized Hamiltonian systems.
In this case, a momentum phase space Z is the cotangent bundle T'M of a config-
uration space M. This phase space is provided with the canonical symplectic form.
A primary constraint space iV C Z is an image of the Legendre map L (3.3.12)
defined by a degenerate Lagrangian on the velocity phase space TM. In fact, one
introduces some initial Hamiltonian H on T'M such that the energy function Ec
(3.3.6) for C is the pull-back C"H of this Hamiltonian H by the Legendre map C.
Thus, we come to a Dirac constraint system on the Lagrangian constraint space N.
The goal consists in constructing a Hamiltonian H' on the momentum phase space
T'M (but not uniquely in general) such that the solution dw of the symplectic Ha-
miltonian system (fi, H') on T'M provides a solution of the above-mentioned Dirac
constraint system on a final constraint space.
130 CHAPTER 3.
HAMILTONIAN SYSTEMS
In fact a Dirac constraint system S
N
-n is the pull-back presymplectic Hamil-
tonian system (i'
N
Q, i'
N
H) on a submanifold N C Z for the Hamiltonian i'
N
H and
with respect to the presymplectic formi'
N
l on N. For instance, Proposition 3.4.2
shows that the equation
i'
N
(v\Q. + dH{z)) =0, v g T
Z
N,
has a solution only at the points of the subset
N
2
= {z e N : u\i'
N
dH(z) =0, Vu Ker
z
fi
N
}
which is assumed to be a manifold. Such a solution, however, fails to be tangent to
N
2
. Then one should repeat the algorithm for presymplectic Hamiltonian systems
in Remark 3.4.1. Nevertheless, one can say more since the presymplectic system
S
N
-H (3.6.1) on TV is the pull-back of the symplectic system (Q, H) on Z.
Let a (2m n)-dimensional closed imbedded submanifold TV of Z be a finai
constraint space of the Dirac constraint system (3.6.1). This means that the equation
v\Q
N
+ dH
N
(z) =0,
has a solution at each point z N. By virtue of Proposition 3.4.2, this is equivalent
to the condition
u\dH
N
(z) = 0, Vu G Ker
2
Q
w
, VzG N,
or
Ker n
N
= TNC\ Orth
n
TN C Ker H. (3.6.2)
We will reformulate these condition in algebraic terms. Let Is be the ideal of
functions on Z which vanish everywhere on the final constraint space N. Let I(N)
be the normalizer (2.6.6) of 1^ and I'{N) = I{N) H I
N
. It is readily observed that,
restricted to TV, Hamiltonian vector fields dj for elements / of I'{N) with respect
to the symplectic formQ on Z take their values into TN nOrth
n
TJ V [97]. Then the
condition (3.6.2) can be written in the form
{H,I'{N)}CI
N
, (3.6.3)
whereas
{H,I
N
}I
N
(3.6.4)
v e T
Z
N,
O/v ^TV^J
Hpj = i*
N
H,
3.6.
DIRAC CONSTRAINT SYSTEMS 131
in general. A glance at the relations (3.6.3) and (3.6.4) shows that, though the
Dirac constraint system on N has a solution, the Hamiltonian vector field dn for
the Hamiltonian H with respect to the symplectic form fi on Z is not tangent to
N, and its restriction to N is not such a solution in general.
Exa mple 3. 6. 1. Let the final constraint space N be a coisotropic submanifold of
the symplectic manifold (Z, fi). Then I
N
= I'{N), i.e., there are only first-class
constraints. In this case, the Hamiltonian vector fields of the Hamiltonian H and
of other Hamiltonians H +/, / IN, provide solutions of the Dirac constraint
system on N. Note that, if a primary constraint space is coisotropic, we arrive at
a Dirac Hamiltonian system, where the relation (3.6.3) is exactly the condition of
Proposition 3.5.1.
The goal of constructing a generalized Hamiltonian system is to find a constraint
/ IN such that the modified Hamiltonian H + f would satisfy both the conditions
{H + f,I'(N)}cI
N
,
{H + /, I
N
} C I
N
-
(3.6.5)
(3.6.6)
The condition (3.6.5) is fulfilled for any / e IN, while (3.6.6) is an equation for a
second-class constraint / .
Remar k 3. 6. 2. Since 1^ is the ideal of functions vanishing on a manifold N,
the normalizer of IN coincides with the normalizer of J '(J V) [97]. Therefore, the
conditions (3.6.5) and (3.6.6) may be combined into
{H + f,I'(N)}cI'(N).

The equation (3.6.6) is solved in many concrete models, that implies separating
first- and second-class constraints. The general problem lies in the fact that the
collection of elements which generate If, C I'(N) is necessarily infinite reducible
[97]. At the same time, the Hamiltonian vector fields for elements of I
N
vanish on
the constraint space N. Therefore, one can apply the following procedure [137].
Since N is a (2mn)-dimensional closed imbedded submanifold of Z, the ideal IN
is locally generated by a finite basis {/}, a = 1,..., n, while N is defined locally by
the equations (3.5.1). Let the presymplectic formQ
n
be of constant rank 2m-n k,
i.e., Ker
z
fl<2 has the dimension k < n at all points z Q. It defines a fc-dimensional
132 CHAPTER 3. HAMILTONIAN SYSTEMS
characteristic foliation on N. Since N C Z is closed, there exist locally k linearly
independent vector fields u
b
on Z which, restricted to N, are tangent to the leaves
of this foliation. They read
n
6 =E &"*/.. 6=1, . . . , * ,
o=l
where g% are functions on Z and #/
0
are Hamiltonian vector fields for constraints
f
a
. Then one can choose the collection of constraints fa, b = 1,... ,n, where the
first k functions take the form
n
4>b=Y^9bfa< 6=l ,...,f c.
Let d^ be the Hamiltonian vector fields for these functions. One can easily justify
the fact that
0fc|jv =u|w, b-l,...,k.
It follows that the constraints fa belong to I'(N) \ I%- Therefore, they are called
the first-class constraints, while the remaining fa.+\, , fa are the second-class con-
straints. We have the relations
n
{fa, fa} = J2
C
bc<t>a,
b=l,...,k, c= l ,...,n,
where C. are local functions on Z.
It should be emphasized that the constraints fa, , fa do not constitute any
local basis for I'{N), though fa,- ,fa make up a local basis for the ideal IN-
Now let us consider a local Hamiltonian on Z
n
%' = Ti + 2_, ^
a
0a, (3.6.7)
where A" are functions on Z. Since Ti obeys the condition (3.6.5), we find
{H,fa)=Y
J
B
a
b
fa, 6=1. . . , A,
a=l
where B% are functions on Z. Then, the equation (3.6.6) takes the form
{H,fa}+ Y. A"{^,0c}= Y,D%, c=k+l,...n,
o=t +l 6=1
(3.6.8)
3.7.
HAMILTONIAN SYSTEMS WITH SYMMETRIES 133
where D
b
c
are functions on Z. This is the system of linear algebraic equations for
the coefficients A, a = k + 1,... n, of second-class constraints. These coefficients
are defined uniquely by the equations (3.6.8), while the coefficients \
a
, a = 1,..., k,
of first-class constraints in the Hamiltonian (3.6.7) remain arbitrary.
Then, being restricted to the constraint space N, the Hamiltonian vector field
of the Hamiltonian H' (3.6.7) on Z provides a local solution of the Dirac constraint
system on N.
We refer the reader to [137] for a global variant of the above procedure.
Exa mpl e 3. 6. 3. If AT is a symplectic submanifold of Z, then I'(N) = 1%. Therefore,
we have only second-class constraints, and the Hamiltonian (3.6.7) of a generalized
Hamiltonian system is defined uniquely.
Remark 3. 6. 4. In concrete models, the final constraint space N fails to be given
in advance. Therefore, a different procedure of separating first- and second-class
constraints is usually applied (see, e.g., [39, 58, 172]), but its global treatment is
under discussion.
3.7 Hamiltonian systems with symmetries
In this Section, we summarize the well-known results connected with a group action
on symplectic and Poisson manifolds [2, 116, 126, 130, 181]. By G throughout is
meant a real connected Lie group.
We start from a symplectic manifold case. Let (Z, Q) be a symplectic manifold
on which a real connected Lie group G acts on the left so that, whenever g G, the
mapping z *>gz is a symplectic isomorphism of Z. Such an action of a group G is
called a symplectic action.
Remar k 3. 7. 1. The classification of transitive symplectic actions of connected Lie
groups has been done (see [69, 72, 103]).
Since G is connected, its action on a manifold Z is symplectic if and only if, for
any element e of the right Lie algebra g
r
of a the group G, the corresponding vector
field
c
on Z is locally Hamiltonian, i.e.,
L
Ct
fi =d(&jn)=0.
134 CHAPTER 3. HAMILTONIAN SYSTEMS
Let us suppose that it is a Hamiltonian vector field, i.e.,

J fi=-d/

,
where J
c
is a real function on Z.
DEFI NI TI ON 3.7.1. A symplectic action of a connected Lie group G on a symplectic
manifold Z is called a ffamiitonian action if, whenever e 6 Q
T
, there exists a function
J
c
on Z such that f

is the Hamiltonian vector field for J


c
.
PROPOSI TI ON 3.7.2. An action of a connected Lie group G on a symplectic manifold
Z is Hamiltonian if and only if there exists a mapping, called a momentum mapping,
from Z to the Lie coalgebra
J Z -
0
*
such that, whenever e 0
r
, the function
J
t
(z) = (J{z),e), Veefl
r
,
(3.7.1)
is that in Definition 3.7.1.
Pr oof. If J exists, we obviously have a Hamiltonian action. Conversely, if
c
for
any e g
r
is the Hamiltonian vector field for a function J
e
on a manifold Z, then
the momentum mapping J is defined by the relation (3.7.1) applied to the basis
{e
m
} of the Lie algebra g
T
. QED
QED
It is readily observed that, if j and J' are different momentum mapping for the
same action of G on Z, then
d{J(z)-J'(z),e)=0,
Exa mple 3.7.2. Let a symplectic form on Z be exact, i.e., ft = dO, and let 6 be
G-invariant, i.e.,
Then the map J : Z 0*, given by the relation
<J (z),

) = (&j0)(z),
Le,0 =d(&J *) +&jn =o,
Ve e a,..
Ve 6 fl
r
,
i.e., J J' = const, on Z.
3.7. HAMILTONIAN SYSTEMS WITH SYMMETRIES 135
is a momentum mapping.
PROPOSI TI ON 3.7.3. If a Hamiltonian H ona symplectic manifold Z is invariant
under a group G acting on Z, a momentum mapping J defines a family of first
integrals of motion of the Hamiltonian system for H.
Pr oof. We have
LfcW =Z
e
\dM = 0,
Ve fl
r
,
and then
{H, J

) =0, Ve
flr
.
QED
Thus, the functions J
e
{z) (3.7.1) are first integrals of motion. However, they an
not pairwise in involution in general. Let us find their Poisson brackets.
DEFI NI TI ON 3.7.4. A momentum mapping J is called equivariant if
J(gz)= Ad'g(J(z)), Vp e G,
where Ad'g is the coadjoint representation (1.5.3).
Exampl e 3.7.3. The momentum mapping in Example 3.7.2 is equivariant. In
accordance with the relation (1.5.3), it suffices to show that
J
e
{gz) = J
Mg
-n
e
)(z),
i.e.,
(^\8)(gz) = (Ad,-i()J *)(*).
The latter results follow from the relation (1.5.2).
Exa mpl e 3. 7. 4. Let T'M be a symplectic manifold equipped with the canonical
symplectic form fi (2.4.2). Let a group G act on M on the left by the generators
U = 4,(9)*-
136 CHAPTER 3. HAMILTONIAN SYSTEMS
The canonical lift of this action onto T'M has the generators
lm = e'
m
di - PAe
3
m
d\
(3.7.2)
and preserves the canonical Liouville form6 on T'M. In accordance with Example
2.4.4, the vector fields (3.7.2) are Hamiltonian vector fields for the functions
J
m = e'
m
(q)pi
(3.7.3)
and, as follows from Example 3.7.3, there exists the equivariant momentum mapping
J = e\
n
{q)p
i
e
m
.
In the case of a G-invariant Hamiltonian on T'M, the functions (3.7.3) are first
integrals of motion.
In the case of non-equivariant momentum mappings, let us consider the difference
a(g) =J(gz) - Ad'g(J(z)). (3.7.4)
We refer the reader to [2] for proof that this difference is constant on a symplectic
manifold Z, and fulfills the equality
o(gg') = o(g) +Ad'g(o(g')). (3.7.5)
DEFI NI TI ON 3.7.5. A map a : G > g* is called the cocycle on a group G if the
condition (3.7.5) holds. A cocycle a is said to be a coboundary on a group G, if
there exists an element p, g* such that
a(g) = p- Ad'g(p,). (3.7.6)
The equivalence classes of cocycles modulo coboundaries make up the cohomology
group of the group G.
By this definition, the difference (3.7.4) is the cocycle of the group G associated
with the action of G on Z.
PROPOSI TI ON 3.7.6. Each symplectic action of a group G on a symplectic manifold
Z defines a cohomological class [a] of G.
3.7. HAMILTONIAN SYSTEMS WITH SYMMETRIES 137
Proof. Let J and J' be different momentum mappings associated with a symplectic
action of a group G on a symplectic manifold Z. Since the difference J J' is constant
on Z, then the difference of the corresponding cocycles a a' is the coboundary
(3.7.6) where n= J- J'. QED
In particular, an equivariant momentum mapping defines the zero cohomological
class of the group G.
THEOREM 3.7.7. Given a momentum mapping J associated with a symplectic
action of a group G on a symplectic manifold Z, the following relation
{J
e
,J
c
,} = J
[c
,
c>]
-{T
e
a{e'),e)
(3.7.7)
takes place (we refer the reader again to [2], where, however, the left Lie algebra is
utilized and Hamiltonian vector fields differ in the minus sign from those we use).

In the case of an equivariant momentum mapping as in Example 3.7.4, the
relation (3.7.7) leads to the homomorphism
{Je ,Je'} = J\t,e'\
(3.7.8)
of the Lie algebra g
r
to the Lie algebra of functions on a symplectic manifold Z
with respect to the Poisson bracket. These are the desired Poisson brackets of first
integrals of motion for a G-invariant Hamiltonian system.
We will refer to this result in connection with the conservation laws in time-
dependent mechanics (see Section 5.8).
Let now (Z, w) be a Poisson manifold, on which a connected Lie group G acts on
the left so that, whenever g e G, the mapping z > gz is a Poisson automorphism
of Z. Such an action of a group G is called a Poisson action. Since G is connected,
its action on a manifold Z is a Poisson action if and only if, for any element e of
the right Lie algebra g
r
of a group G, the corresponding vector field
t
on Z is an
infinitesimal Poisson automorphism, i.e., the condition (2.3.3) holds. The equivalent
conditions are
&({/,}) = {&(/).$} + {/.&($)}.
&({/,$}) = [ &,*/] ()-f c.W).
[ &.*/] = *.</).
138 CHAPTER 3. HAMILTONIAN SYSTEMS
where / , j D(Z), and by d
t
is meant the Hamiltonian vector field (2.3.4) for a
function /. Note that, by very definition of a Hamiltonian action, generators of a
Hamiltonian action of a group G on a Poisson manifold (Z, w) are tangent to the
leaves of the symplectic foliation on Z, and there is a Hamiltonian action of G on
every symplectic leaf in the sense of symplectic geometry.
Definition 3.7.1, Propositions 3.7.2, 3.7.3, and Definition 3.7.4 are naturally ex-
tended to a Poisson action of a connected Lie group G on a Poisson manifold. Let
us consider the case of an equivariant momentum mapping.
PROPOSI TI ON 3.7.8. [181] An equivariant momentum mapping J : Z >0* is a
Poisson morphism if the coalgebra a* is provided with the Lie-Poisson structure
(2.3.11). a
There is the following reduction theorem.
THEOREM 3.7.9. [181]. Let (Z,w) be a Poisson manifold endowed with a Hamilto-
nian action of a connected Lie group G and the equivariant momentum map J. Let
q be a point of g* such that
q is a non-critical value for all the restrictions of J to the symplectic leaves
F
L
of Z, i.e., the level set Z
q
= J~
l
(q) is a submanifold of Z and, for any
symplectic leaf F
t
, F
uj
= Z
q
V\ F
L
is a submanifold of F
L
;
intersections F^ of the submanifold Z
q
with symplectic leaves of Z are clean,
i.e., TFu, =TF
L
f~l TZ
q
, and so are its intersections with the orbits of G in Z.
Let a subgroup G
q
C G be the stabilizer of the point q. Then intersections F^ make
up a regular foliation on Z,, and its leaves are the orbits of the connected component
of the unit in G
q
for the action of G
q
on Z
q
. Furthermore, if this foliation is defined
by a submersion Z
q
>P, the base P has a well-defined Poisson structure whose
characteristic distribution is the projection of TC\TZ
q
, where T is the characteristic
distribution on (Z, w). This is exactly the reduced Poisson manifold of {Z,w) via
(Z E = T(orbits G)) (see Definition 2.6.3). a
Note that if the momentum mapping J of a Hamiltonian action on a Poisson
manifold is not equivariant, we have the relation
J(gz)=Ad'g(J(z))+<j(g,z),
(3.7.9)
3.8. APPENDIX. HAMILTONIAN FIELD THEORY
139
similar to (3.7.4) in the symplectic case, where the function a(g, z) is constant on
the leaves of the symplectic foliation on Z. Since now a(g, z) depends on z, we have
different stabilizers of q for the action given by (3.7.9) on g* for every symplectic
leaf. Therefore, the foliation on Z
q
is no longer regular.
We end this Section with the theorem which shows the possibility of reducing a
Poisson Hamiltonian system possessing symmetries.
THEOREM 3.7.10. [129, 181]. Let (Z, N, E) be a Poisson reduction and P a reduced
manifold of Z via (N, E). Let H be a Hamiltonian on Z such that the flow $
t
of
its Hamiltonian vector field d
H
preserves the submanifold N and the bundle E, and
dH is an annihilator of E. Then $, induces a flow $

of Poisson automorphisms of
P (see Proposition 2.6.5) along the Hamiltonian vector field dq for the Hamiltonian
H on P, defined uniquely by the relation H \
N
= n'H. Moreover, we have
dq o 7r = TIT O T3
H
.
a
Thus, one can say that, under the hypotheses of Proposition 3.7.10, the Poisson
Hamiltonian system on a Poisson manifold Z reduces to a Poisson system on the
reduced Poisson manifold P.
Exa mpl e 3. 7. 5. Let (Z, Z
q
, E =T(orbits G)) be a Poisson reduction as in Theorem
3.7.9. If H is a G-invariant Hamiltonian on the Poisson manifold Z, the Poisson
Hamiltonian system reduces to a Poisson system on the reduced Poisson manifold
of (Z, w) via (Z E =T(orbits G)).
3.8 Appendix. Hamiltonian field theory
As is well-known, applied to a field theory on a fibre bundle Y -+ X, the famil-
iar symplectic techniques take the form of instantaneous Hamiltonian formalism
on an infinite-dimensional momentum phase space [64], in contrast with the finite-
dimensional velocity phase space J
l
Y. The Hamiltonian counterpart of the La-
grangian formulation of field theory is polysymplectic Hamiltonian formalism on
the Legendre bundle n (2.9.7) provided with the polysymplectic form A (2.9.9) (see
[57, 159] for a survey). Here we aim to give a brief exposition of this formalism
140 CHAPTER 3. HAMILTONIAN SYSTEMS
because its restriction to fibre bundles Q ->K leads in a straightforward manner to
the Hamiltonian formulation of time-dependent mechanics [57, 161).
Let Y >X be a fibre bundle over an n-dimensional manifold X, and
n = VY ACA'T'X)
(3.8.1)
the Legendre bundle (2.9.7). It admits the composite fibration
TnX = 7T O 1TUY fl V > X.
(3.8.2)
For the sake of convenience, we will call n >Y the Legendre vector bundle, while
the term Legendre bundle stands for the fibration 17 X.
Given fibred coordinates (x\ y') on the fibre bundle Y X, the Legendre bundle
(3.8.1) is equipped with the holonomic coordinates (x
x
,y',p*) together with the
transition functions (2.9.8). These coordinates are compatible with the composite
fibration (3.8.2) and are linear bundle coordinates on the vector bundle U > Y. We
will call them the canonical coordinates.
There is the canonical bundle monomorphism over Y
G: IT -^
n
/\T'YTX,
Y Y
9 : (x\ y\p$) ~ -pfdy' Acjd
x
,
(3.8.3)
which defines the tangent-valued Liouville form on the Legendre bundle IT Then
the polysymplectic form
A = dp,
A
A dy
{
A u <g>d
x
(3.8.4)
on II is defined as a unique TX-valued (n +2)-form on II such that the relation
A\cf> =-d(G]<j>)
holds for any exterior 1-form4> on X.
Given the vector Legendre bundle II over a fibre bundle Y X, we have the
exact sequence
o >UX A T " X ^-> z
Y
>n >o,
x
(3.8.5)
where Zy is the homogeneous Legendre bundle (2.9.3) and
Tzn : Z
Y
II (3.8.6)
3.8. APPENDIX. HAMILTONIAN FIELD THEORY 141
is an affine bundle over II with 1-dimensional fibres.
DEFI NI TI ON 3.8.1. Let h be a section of the fibre bundle (3.8.6). Then the pull-back
H = h*Z : II-> AT*Y,
H =p
x
dy* AUJ\- Hu,
(3.8.7)
of the canonical form B (2.9.4) on Zy by h is called the polysymplectic Hamiltonian
form (or simply the Hamiltonian form).
The exterior differential of the Hamiltonian form (3.8.7) is the pull-back
dH = h'Q
z
on II of the multisymplectic form (2.9.5).
Exampl e 3. 8. 1. Let
r =dx
x
<8>(d
x
+ T\di)
be a connection on the fibre bundle Y > X. Hence, we have the splitting
T : V'Y - T'Y,
T.dtf^ dy
l
- T\dx
x
,
of the exact sequence (1.1.8a). Then T also yields the splitting
h
r
-. n - Z
Y
,
h
r
: pfdy' w
A
<-> p
X
dy
{
A w
x
- p
X
T\u),
of the exact sequence (3.8.5). It follows that every connection T on the fibre bundle
Y X defines the Hamiltonian form
H
r
= h'
r
E,
H
r
= p
x
dy
{
ALJ
X
- p,
A
I >,
(3.8.8)
on the Legendre bundle n.
PROPOSI TI ON 3.8.2. [57, 159]. Hamiltonian forms on n constitute an affine space
modelled over the linear space of horizontal densities
H = Hw:Yl^ AT'X
(3.8.9)
142 CHAPTER 3. HAMILTONIAN SYSTEMS
on the Legendre bundle II X. They axe called Hamiltonian densities.
This means that, if H is a Hamiltonian form and H is a horizontal density (3.8.9),
then H - H is also a Hamiltonian form. Conversely, if H and H' are Hamiltonian
forms, their difference H - H' is a Hamiltonian density (3.8.9).
Example 3.8.1 and Proposition 3.8.2 combine into the following.
COROLLARY 3.8.3. Every Hamiltonian form on the Legendre bundle n admits the
decomposition
H = H
r
- H
r
= p
x
dy
i
A w
A
- P
X
T\UJ - H
r
u>,
(3.8.10)
where T is a connection on Y X. O
Furthermore, every Hamiltonian form admits a canonical decomposition as fol-
lows. We mean by a Hamiltonian map any fibred morphism
$ : II -> J
l
Y,
Y
(3.8.11)
over Y. Its composition with the canonical morphism A (1.3.4) yields the bundle
morphism
A o$: n ^T'XTY
Y Y
represented by the TY-valued 1-form
* =dx
x
9 (d
x
+ $\(q)d
t
) (3.8.12)
on the vector Legendre bundle n > Y.
Exampl e 3.8.2. Let T be a connection on Y X. Then, the composition
f =ro7r
ny
: n y ->J
1
Y,
f =dx
x
(d
x
+ r
A
6ts),
is a Hamiltonian map. Conversely, every Hamiltonian map $ : n J
l
Y yields the
associated connection
T* =$ o 0
y[ o $ = &
x
{q), <?en,
3.8. APPENDIX. HAMILTONIAN FIELD THEORY 143
on Y X, where 6 is the global zero section of the vector Legendre bundle II Y.
In particular, we have
r~ = r
1
r
1


PROPOSI TI ON 3.8.4. [57, 159] Every Hamiltonian form H on the Legendre bundle
II defines the associated Hamiltonian map
H : U - J
1
^
(3.8.13)
D
COROLLARY 3.8.5. Every Hamiltonian form H on the Legendre bundle n deter-
mines the associated connection
T
H
= HoO
on Y -> X. a
In particular, we have
T//J. r,
where HY is the Hamiltonian form (3.8.8) associated with the connection T on
Y -> X.
COROLLARY 3.8.6. Every Hamiltonian form (3.8.10) admits the canonical splitting
H = HY
H
H.
a
The following assertion generalizes Example 3.8.1.
PROPOSI TI ON 3.8.7. Every Hamiltonian map (3.8.11) represented by the form
(3.8.12) on n defines the associated Hamiltonian form
H* = $j e =pfdy* Aw-p*$'
A
w,
y\oH = d{H.
144 CHAPTER 3. HAMILTONIAN SYSTEMS
where 0 is the tangent-valued Liouville form (3.8.3).
In particular, if
H
S
= H,
then H = Hr for some connection r on Y.
Hamilton equations in conservative mechanics are the equations of integral curves
of Hamiltonian vector fields. Hamilton equations in polysymplectic Hamiltonian for-
malism are the equations of integral sections for Hamiltonian connections as follows.
Let J ' n be the first order jet manifold of the Legendre bundle II >X. It is
equipped with the adapted fibred coordinates
(x
A
,
3
/
,
,p,
A
,^,pV).
We have the commutative diagram
I I
n 52. Y
y]t J^ar = vl-
DEFI NI TI ON 3.8.8. A connection
y = dx*(d
x + y
\d, + rt
i
dl)
on the Legendre bundle n X is said to be a locally Hamiltonian connection if
the exterior form
7
J A =drf A dy
l
Au
x
- (
7
|dp
A
- y^dy') A u
(3.8.14)
is closed. D
Exa mple 3.8.3. Every connection T on a fibre bundle Y > X gives rise to the
connection
f = dx
x
[dx + r
x
(y)di +
(3.8.15)
(-d
}
r
x
tf + K
X
\
P
] - K
x
\f
}
)^\
j^J'j^Xjly
3.8. APPENDIX. HAMILTONIAN FIELD THEORY
145
on the Legendre bundle II X, where A" is a symmetric linear connection (1.4.9)
on TX. The connection f (3.8.15) obeys the relation
fjA = d(rj 6).
It follows that T is a locally Hamiltonian connection.
Thus, locally Hamiltonian connections always exist on the Legendre bundle n *
X, and every connection r on Y >X gives rise to a locally Hamiltonian connection
on n -> X.
It is easily observed that a connection 7 on the fibre bundle n >X is a locally
Hamiltonian connection if and only if 7 fulfills the conditions
dW, ~ dill = 0,
d.7^- %"!& = 0,
d/7i + = 0 .
(3.8.16)
(3.8.17)
(3.8.18)
Using the relation (3.8.18), we find that the second term in the right-hand side
of the expression (3.8.14) is a closed form. Then, in accordance with the relative
Poincare lemma, this expression is brought locally into the form
7J A =d(p
x
dy* A uj - Hyu) = dH
7
, (3.8.19)
where H-, is a local function on n such that
l l =W , 7A
A
, = -dtUr
Given a connection T on the fibre bundle Y > X, the local formH^ in the expression
(3.8.19) can be written as
H
1
= H
r
- H
V
UJ,
where Hpui is a local horizontal density on n X. In accordance with Proposition
3.8.2, it follows that //
7
defines the local section
h
1
:(x
x
,y\p?)~(x
x
,y\plp=-HJ
of the fibre bundle Z
Y
- n, i.e., H^ is a local Hamiltonian form. Thus, we have
proved the following.
146 CHAPTER 3. HAMILTONIAN SYSTEMS
PROPOSI TI ON 3.8.9. For every locally Hamiltonian connection 7 on the Legendre
bundle n X, there exists a local Hamiltonian formH in a neighbourhood of each
point q fl such that
7J A = dH.
a
Let us formulate the converse assertion.
DEFI NI TI ON 3.8.10. The Hamilton operator
H
for a Hamiltonian form H on the
Legendre bundle n X is denned to be the first order differential operator
E
H
:J
1
n^
n
A
1
T'U,

H
= dH-A = [(y\ - d\H)dp* - (pk +djTi)^] A u, (3.8.20)
where
A =dp
x
x
A dy' Au
A
+p
x
M
dy' Au - y\dp* Au
is the pull-back of the polysymplectic form A (3.8.4) on J 'n. D
A glance at the expression (3.8.20) shows that the Hamilton operator EH is an
afEne morphism over n of constant rank. Thereby, its kernel is an affine subbundle

H
=Ker
H
-* n (3.8.21)
of the affine jet bundle J ' n > U which is given by the coordinate relations
v\ = d\n,
This affine bundle is modelled over the vector subbundle of the vector bundle
r*xvn->n
n
(3.8.22)
which is defined by the coordinate relations
yl = o, PA, = 0
with respect to the fibre coordinates (FA.PAI)
on
(3-8.22).
(3.8.23)
Pi = -diH.
3.8. APPENDIX. HAMILTONIAN FIELD THEORY 147
Since < (3.8.21) is an affine subbundle, it is a closed imbedded submanifold
of the jet bundle J
l
U >X and, therefore, is a system of first order differential
equations on II X in accordance with Definition 1.3.3.
DEFI NI TI ON 3.8.11. The first order differential equations
w
(3.8.21) are called the
Hamilton equations for the Hamiltonian formH on the Legendre bundle II.
Since the subbundle <BH (3.8.21) is affine, it always admits a global section 7.
Any such section is a connection on II X which meets the condition

H
o 7 = 0.
This condition takes the form
7J A = dH. (3.8.24)
It follows that every connection on II X which takes its values into the Hamilton
equations # is a locally Hamiltonian connection.
DEFI NI TI ON 3.8.12. A locally Hamiltonian connection 7 on the Legendre bundle
n >X is said to be a Hamiltonian connection associated with a Hamiltonian form
H if 7 obeys the relation (3.8.24).
Thus, we have proved the following.
PROPOSI TI ON 3.8.13. Every Hamiltonian form on the Legendre bundle n has an
associated Hamiltonian connection. D
We have the equations of a Hamiltonian connection associated with a given
Hamiltonian form:
7i = d\H,
7A
A
i = - m
(3.8.25)
(3.8.26)
By the equation (3.8.25), every Hamiltonian connection 7 for a Hamiltonian form
H satisfies the relation
J
l
ir
n
y 07 = //,
7 = dx
x
(d
x
+ diHdi + T&S*),
(3.8.27)
148 CHAPTER 3. HAMILTONIAN SYSTEMS
where H is the Hamiltonian map (3.8.13). It projects over the connection YH on
Y >X associated with the Hamiltonian form H. We have the following commuta-
tive diagram
j ' n
n -> J
1
Y
o\/* r
Y
A glance at the equations (3.8.26) shows that there is a set of Hamiltonian
connections associated with the same Hamiltonian form H. They differ from each
other in soldering forms a on n X which obey the equations
5JQ =0,
31 =0,
and take their values into the constraint subbundle (3.8.23) of the vector bundle
(3.8.22).
A classical solution of the Hamilton equations (3.8.21), by definition, is a section
r of the Legendre bundle n X such that its jet prolongation J
1
r takes its values
into the kernel of the Hamilton operator ZH (3.8.20). Then, r satisfies the differential
equations
8^ = diU or, (3.8.28a)
(3.8.28b)
Every integral section J
l
r = 7 o r of a Hamiltonian connection 7 associated
with a Hamiltonian form H is a classical solution of the corresponding Hamilton
equations. Conversely, if r is a global solution of the Hamilton equations (3.8.28a)
- (3.8.28b) for a Hamiltonian form H, there exists an extension of this solution
Jh-.riX) j ' n
to a Hamiltonian connection which has r as an integral section.
Substituting J
l
r in (3.8.27), we obtain the identity
J '(7Tny or) = H or
J'*nr
si = o,
a
A
r,
A
= -diH o r.
3.8. APPENDIX. HAMILTONIAN FIELD THEORY 149
for every classical solution r of the Hamilton equations (3.8.28a - (3.8.28b).
Rema r k 3. 8. 4. Note that the Hamilton equations (3.8.28a) - (3.8.28b) can be
introduced without appealing to the Hamilton operator. They are equivalent to the
relation
r'{u\dH) = Q
which is assumed to hold for any vertical vector field u on the Legendre bundle
n-> X.
Cha pt er 4
Lagr angian t i me-dependent
mecha ni cs
This Chapter is devoted to the formulation of time-dependent mechanics on a phase
space of coordinates and velocities. This phase space is the first order jet manifold
J
l
Q of a configuration bundle Q R. The familiar case of the direct product
Q = R x M corresponds to the choice of a certain reference frame. For the sake
of brevity, the above-mentioned formulation is called Lagrangian time-dependent
mechanics although equations of motion are not necessarily Lagrange equations.
Connections provide the main ingredients in the Lagrangian formulation of time-
dependent mechanics. These are the reference frames seen as connections on a
configuration bundle Q R, the dynamic equations represented by connections
on the jet bundle J
l
Q R, the dynamic connections on the affine jet bundle
J
l
Q Q, the connections on the tangent bundles TQ >Q and TJ
l
Q J
l
Q,
and the Lagrangian connections. For instance, we show that every non-relativistic
dynamic equation on a configuration space Q can be seen as a geodesic equation
with respect to some connection on the tangent bundle TQ >Q.
Reference frames are necessarily involved when we deal with dynamic equations,
forces, accelerations, and conservation laws in time-dependent mechanics.
The mass metric is another interesting object of Lagrangian and Newtonian
systems, of the inverse problem, and of systems with holonomic and non-holonomic
constraints.
Throughout this Chapter,
TT:Q R
(4.0.1)
151
152 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
is a fibre bundle whose typical fibre M is an m-dimensional manifold, while its base
R is parameterized by the Cartesian coordinates t possessing the transition functions
t' = t+ const. In the universal unit system, the physical dimension of t is equal to
[length]. A fibre bundle Q R is endowed with bundle coordinates {t,q'}. For the
sake of convenience, we also use the compact notation q
x
where q = t.
4.1 Fibre bundles over R
In this Section, we point out the most important peculiarities of fibre bundles over
R, which we refer to in the sequel.
1.
The base R of a fibre bundle (4.0.1) is provided with the standard vector field 3,
and the standard 1-formdt which are invariant under the coordinate transformations
f = t+ const. The same symbol dt also stands for any pull-back of the standard 1-
formdt onto a fibre bundle over R. Note on one-to-one correspondence between the
vector fields fd
t
and the real functions / on R. The similar correspondence between
the horizontal densities 4>dt and the real functions 4> on a, fibre bundle Q ^ R
takes place. Roughly speaking, we may neglect the contribution of TTR and T*R to
several expressions. At the same time, one should be careful with such simplification
in the framework of the universal unit system. For instance, the coefficient 0 of a
horizontal density <j>dt has the physical dimension [length]
-1
, whereas a function (j>
is physically dimensionless.
2.
Since R is contractible, any fibre bundle over R is obviously trivial. Different
trivializations
rj) :Q =R x M
(4.1.1)
differ from each other in the fibrations Q >M, while the fibration Q * R is once
for all.
Let J 'Q be the first order jet manifold of a fibre bundle Q R (4.0.1). It is
provided with the adapted coordinates (t,q',q'
t
). Given a direct product
Q = R x M - > 1
4.1. FIBRE BUNDLES OVER K 153
coordinated by (f.,if), there is the canonical isomorphism
J
l
{RxM) =RxTM,
i
(4.1.2)
that one can justify by inspection of the transition functions of the coordinates q[ and
q , when the transition functions of q
x
are independent of t. Due to the isomorphism
(4.1.2), every trivialization (4.1.1) yields the corresponding trivialization of the jet
manifold
J
l
Q = 1 x TM. (4.1.3)
3.
Given the jet manifold J
l
Q of a fibre bundle Q > R, the canonical imbedding
(1.3.4) takes the form
A: J'Q^TQ,
X:(t,q\q'
t
)^(t,q',i = l,q' = ql).
(4.1.4)
For brevity, we will write
A =ck =d
t
+ q\di, (4.1.5)
where by dt is meant the total derivative. From now on, we will identify the jet
manifold J
l
Q with its image in TQ.
Rema r k 4. 1. 1. Following precisely the expression (1.3.4), we should write the
morphism A (4.1.5) in the form
A =dt (d
t
+ qldi). (4.1.6)
With respect to the universal unit system, the physical dimension of A (4.1.5) is
[length]"
1
, while A (4.1.6) is dimensionless.
Using the morphism (4.1.4), one can define the contraction
J
l
QxT'Q ->QxR,
Q Q
(|; t, 9.) - ^J fat + q,dq') = i + q\q
(4.1.7)
where (i, g', t, q
{
) are the coordinates on the cotangent bundle T'Q.
154 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
4.
A glance at the expression (4.1.4) shows that the affine jet bundle J 'Q >Q is
modelled over the vertical tangent bundle VQ of the fibre bundle of Q >R. As a
consequence, we have the following canonical splitting (1.1.6) of the vertical tangent
bundle V
Q
J 'Q of the affine jet bundle J
X
Q - Q:
a : V
Q
J
l
Q Si J
l
Q x VQ,
"(flf) =ft,
(4.1.8)
together with the corresponding splitting of the vertical cotangent bundle VqJ
l
Q
of J
l
Q - Q:
a* : V<V'Q S J 'Q x VQ,
a*(a
7
j)=3
9
i
!
(4.1.9)
where dq\ and dq' are the holonomic bases for VQ J
l
Q and V'Q, respectively. Then
the exact sequence (1.6.9a) of vertical bundles over the composite fibre bundle
J
X
Q -* Q - R (4.1.10)
reads
I ""' 1
0 *V
Q
J
l
Q <^VJ
l
Q ^*J
l
QxVQ *0.
Hence, we obtain the following linear endomorphism over J
1
Q of the vertical tangent
bundle V J
l
Q of the jet bundle J
X
Q R:
v^ioa^ony : VJ'Q^ VJ
l
Q,
v(dj) = dl jjjgj =0-
This endomorphism obeys the nilpotency rule
v o v =0.
(4.1.11)
(4.1.12)
Combining the horizontal splitting (1.3.11), the corresponding projection
pr
2
: J
X
Q x T Q ^ J
l
Q xVQ^ VQJ'Q,
Q Q
d
t
~ -q\d\, d
x
-> d\,
4.1. FIBRE BUNDLES OVER R
155
and the inclusion VJ
l
Q - TJ
l
Q, one can extend the endomorphism (4.1.11) to
the tangent bundle TJ
l
Q:
v:TJ
l
Q^TJ
l
Q,
v(d
t
) = -qldj vjdi) g, g(g) =0.
(4.1.13)
This is called the verticaJ endomorphism. It inherits the nilpotency property (4.1.12).
The transpose of the vertical endomorphismv (4.1.13) is
v' : T'J
l
Q ->TV 'Q,
v'(dt) = 0, v'(dq') = 0, v*(dql) = 6\
(4.1.14)
where
9
l
= dq
i
- q\dt
are the contact forms (1.3.6). The nilpotency rule
v" o v* =0
is also fulfilled. The homomorphisms v and v' are associated with the tangent-valued
1-form
v =^ a?
in accordance with the relations (1.2.36) - (1.2.37).
With the endomorphism {?*, one can introduce the vertical exterior differential
dy = v* o d,
acting on the algebra D'iJ^) of exterior forms on the jet manifold J
l
Q. For
example, if / is a function on J
l
Q, we have
d.f = %f#.
5.
In view of the morphism A (4.1.4), any connection
r =dt (d
t
+ra<)
(4.1.15)
156 CHAPTER 4.
LAGRANGIAN TIME-DEPENDENT MECHANICS
on a fibre bundle Q - R can be identified with a nowhere vanishing horizontal
vector field
T = d
t
+ r<9,
(4.1.16)
on Q which is the horizontal lift rd
t
(1.4.7) of the standard vector field d, on R
by means of the connection (4.1.15). Conversely, any vector field T on Q such that
dt\T = 1 defines a connection on Q R. Of course, the integral curves
c : R D ( ) ^Q,
i(r) = l, c
i
(T) = roc(r)
<
r e( ) ,
of the vector field (4.1.16) coincide with the integral sections
d
t
c\t) = (roc)'(t)
of the connection (4.1.15).
Connections on a fibre bundle Q * R constitute an affine space modelled over
the vector space of vertical vector fields on Q R. Accordingly, the covariant
differential (1.4.5) associated with a connection r on Q >R takes its values into
the vertical tangent bundle VQ of Q >R:
D
r
: J
l
Q -+VQ,
q'oD
r
= q\ - P.
(4.1.17)
A connection T on a fibre bundle Q > R is obviously flat. It yields a horizontal
distribution on Q. The integral manifolds of this distribution are integral curves
of the vector field (4.1.16) which are transversal to the fibres of the fibre bundle
Q~*R.
COROLLARY 4.1.1. By virtue of Proposition 1.4.1, each connection T on a fibre
bundle Q R defines an atlas of local constant trivializations of Q R such
that the associated bundle coordinates (t,q') on Q possess the transition function
q' >q"(q
3
) independent of (, and
r = d
t
(4.1.18)
with respect to these coordinates. Conversely, every atlas of local constant trivial-
izations of the fibre bundle Q >R determines a connection on Q > R which is
equal to (4.1.18) relative to this atlas. D
4.1. FIBRE BUNDLES OVER R
157
A connection T on a fibre bundle Q R is said to be complete if the horizontal
vector field (4.1.16) is complete.
PROPOSI TI ON 4.1.2. Every trivialization of a fibre bundle Q- >R yields a complete
connection on this fibre bundle. Conversely, every complete connection r on Q R
defines its trivialization (4.1.1) such that the vector field (4.1.16) equals d
t
relative
to the bundle coordinates associated with this trivialization.
Proof. Every trivialization of Q R defines the horizontal lift of the standard
vector field d
t
on R to the vector field r = d
t
on Q which is obviously complete.
Then the corresponding connection on Q R is also complete. Conversely, let T
be a complete connection on a fibre bundle Q > R. The horizontal vector field T
(4.1.16) is the generator of a 1-parameter group Gr acting freely on Q. The orbits
of this action are integral curves of the vector field F. Hence, we obtain a projection
Q - Q/G
r
= M (4.1.19)
of Q along the above-mentioned integral curves onto some fibre of Q, e.g., Q
t
=o
M. Combining the projection (4.1.19) and the projection Q > R gives a desired
trivialization (4.1.1) of Q. QED QED
6.
Let J
1
J
1
Q be the repeated jet manifold of a fibre bundle Q >R, provided with
the adapted coordinates
(t,q\qlq\
t
),qlt)-
For a fibre bundle Q * R, we have the canonical isomorphism k between the affine
fibrations 7r
n
(1.3.13) and J
l
n\ (1.3.15) of J
X
J
[
Q over J
l
Q, i.e.,
7r
n
ok = J
0
7T
0
1,
kok^ld^J'Q,
where
q't
0
k = q\
t)
, q\t) k = q\, qtt k = q'u-
(4.1.20)
In particular, the affine bundle 7r
u
(1.3.13) is modelled over the vertical tangent
bundle V J
l
Q of J
l
Q - R which is canonically isomorphic to the underlying vector
bundle J
l
VQ J
l
Q of the affine bundle J
1
^ (1.3.15).
158 CHAPTER 4.
LAGRANGIAN TIME-DEPENDENT MECHANICS
By JQJ
1
Q throughout is meant the first order jet manifold of the affine jet bundle
J
l
Q Q. The adapted coordinates on J
Q
J
l
Q are (g\ q\, q\
t
).
For a fibre bundle Q -> 1, the sesquiholonomic jet manifold PQ coincides with
the second order jet manifold J
2
Q, coordinated by
(<.*. {,)
The affine bundle J
2
Q J 'Q is modelled over the vertical tangent bundle
V
Q
J 'Q ^Q x VQ - J 'Q
(4.1.21)
of the affine jet bundle J
X
Q Q. There are the imbeddings
J
2
Q ^T V ' Q " VQTQ * T
2
Q C TTQ,
>2 : (t, ?', {, qtt) >- (*, ?', $> * = 1,9* =<?!> ?l =?),
TA o A
2
: (t, q\ q\, g'
t
) w (t, <?\ t =t = 1, q* = q<= g*, t =0, <j* = q'
tt
),
(4.1.22)
(4.1.23)
where
(i ,
g
\i ,
g
' ,t,d/,i ' ,<?)
are the coordinates on the double tangent bundle TTQ, by VQTQ is meant the
vertical tangent bundle of TQ Q, and T
2
Q C TTQ is a second order tangent
space, given by the coordinate relation i = i.
Due to the morphism (4.1.22), any connection on the jet bundle J
l
Q >R is
represented by a horizontal vector field on J
l
Q such that \dt = 1.
Exa mple 4.1.2. A connection on the jet bundle J*Q R is defined as a section of
the affine bundle ir
n
(1-3.13). A connection T (4.1.16) on a fibre bundle Q R has
the jet prolongation to the section J
1
F of the affine bundle J 'TQ. By virtue of the
isomorphism k (4.1.20), every connection r on Q R gives rise to the connection
JT
d
=ko J T : J
l
Q^ J 'J 'Q,
JT = dt + Pd, + d
t
rdl (4.1.24)
on the jet bundle J^Q -* R.
A connection on the jet bundle J
1
Q R is said to be holonomic if it is a section
= 4(A + flfo+fflf)
4.2. DYNAMIC EQUATIONS
159
of the holonomic subbundle J
2
Q >J
l
Q of J
1
J
l
Q -* J
l
Q. In view of the morphism
(4.1.22), a holonomic connection is represented by a horizontal vector field
(, =a, +q\d
{
+ gg (4.1.25)
on J 'Q. Conversely, every vector field f on J
l
Q which fulfills the conditions
dt\t. = 1, v(0 =0,
where v is the vertical endomorphism (4.1.13), is a holonomic connection on the jet
bundle J
l
Q ->R.
Holonomic connections (4.1.25) make up an affine space modelled over the linear
space of vertical vector fields on the affine jet bundle J
l
Q Q, i.e., which live in
VQJ'Q.
A holonomic connection defines the corresponding covariant differential (4.1.17)
on the jet manifold ^Q:
(4.1.25)
Ds : J ' J ' Q ^V
Q
J
l
Q C VJ
l
Q,
q'oD
i =
0, $ o D
(
= q\
t
- f,
which takes its values into the vertical tangent bundle VQ.J
1
Q of the jet bundle
J
l
Q y Q. Then, by virtue of Proposition 1.3.1, any integral section c : () J
l
Q
for a holonomic connection is holonomic, i.e., c =c where c is a curve in Q.
4.2 Dynamic equations
We start our exposition of Lagrangian time-dependent mechanics from the notion
of a second order dynamic equation on a configuration bundle Q R.
Recall that a dynamic equation, by definition, is a differential equation which
can be algebraically solved for the highest order derivatives.
DEFI NI TI ON 4.2.1. Let
r =d
t
+ F'di
be a connection on a fibre bundle Q > M. The corresponding covariant differential
D
T
(4.1.17) is a first order differential operator on Q. Its kernel, given by the
coordinate relations
fl* = !*(*,').
(4.2.1)
160 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
is a closed subbundle of the jet bundle J
l
Q - R. Therefore, in accordance with
Definition 1.3.3, this is a first order differential equation on the fibre bundle Q > R,
called the first order dynamic equation on Q R.
Solutions of the first order dynamic equation (4.2.1) are integral sections (geodesic
curves) for the connection I \
DEFI NI TI ON 4.2.2. Let us consider the first order dynamic equation (4.2.1) on the
jet bundle J
1
Q R, which is associated with a holonomic connection f (4.1.25)
on J 'Q R. This is a closed subbundle of the second order jet bundle J
2
Q R,
given by the coordinate relations
& = ?,<!>,<d).
(4.2.2)
Consequently, it is a second order differential equation on the fibre bundle Q R
in accordance with Definition 1.3.3. This equation is called a second order dynamic
equation, or simply a dynamic equation, if there is no danger of confusion. The
corresponding horizontal vector field f (4.1.25) is also termed a dynamic equation.
a
A solution of the dynamic equation (4.2.2), called a motion, is a curve c in Q
whose second order jet prolongation c lives in (4.2.2). It is clear that any integral
section c for the holonomic connection is the jet prolongation c of a solution c of
the dynamic equation (4.2.2), i.e.,
c' = e c, (4.2.3)
and vice versa.
Remar k 4. 2. 1. By very definition, the second order dynamic equation (4.2.2) on a
fibre bundle Q > R is equivalent to the system of first order differential equations
9(0 = 9t.
Qlt = C(t,q
j
,ql),
(4.2.4)
on the jet bundle J
l
Q > R. Any solution c of these equations takes its values into
J
2
Q and, by virtue of Proposition 1.3.1, is holonomic, i.e., c = c. Therefore, the
equations (4.2.2) and (4.2.4) are equivalent. The equation (4.2.4) is said to be the
first order reduction of the second order dynamic equation (4.2.2) .
4.2. DYNAMIC EQUATIONS
161
One can easily find the transformation law of a second order dynamic equation
under bundle coordinate transformations q
l
>q'^t^g
3
). This transformation law
reads
<& = ?,
e = (?dj +gjgfod
k
+ 24djdt +9?W% ?)
(4.2.5)
Exa mpl e 4. 2. 2. From the physical viewpoint, the most interesting dynamic equa-
tions are the quadratic dynamic equations, i.e., those which are polynomials in the
coordinates q\ of degree <2. This property is global due to the transformation law
(4.2.5). The dynamic equation of a relativistic particle exemplifies a non-polynomial
dynamic equation.
A dynamic equation ( ona fibre bundle Q K is said to be conservative if
there exists a trivialization (4.1.1) of Q whose corresponding trivialization (4.1.3)
of J*Q is such that the vector field f (4.1.25) on J
l
Q is projectable over M. Then
this projection
H
f
= <j
,
'3 + e(<rW)<
is a second order dynamic equation on the typical fibre M of Q * K in accordance
with Definition 3.1.2. Conversely, every autonomous second order dynamic equation
E on a manifold M can be seen as a conservative dynamic equation
is = d
t
+ 4% + u'di
(4.2.6)
on the fibre bundle K x M >K in accordance with the isomorphism (4.1.3).
We conclude this Section with the following theorem.
THEOREM 4.2.3. Any dynamic equation on a fibre bundle Q K is equivalent to
an autonomous second order dynamic equation on a manifold Q.
Pr oof. Given a dynamic equation on a fibre bundle Q >R, let us consider the
diagram
J*Q -^T
2
Q
' I I
s
JIQ J u TQ
(4.2.7)
where 5 is a holonomic vector field on the tangent bundle TQ, and we use the
morphism (4.1.23). A glance at the expression (4.1.23) shows that the diagram
162
CHAPTER 4.
LAGRANGIAN TIME-DEPENDENT MECHANICS
(4.2.7) can be commutative only if the component 5 of a vector field S vanishes.
Since the transition functions t > t' are independent of g\ such a vector field
may exist on TQ. Now, the diagram (4.2.7) becomes commutative if the dynamic
equation f and a vector field E fulfill the relation
C = S%q>,i = l,q>=q>).
(4.2.8)
It is easily seen that this relation holds globally because the substitution of q
l
= q\
in the transformation law of a vector field H restates the transformation law (4.2.5)
of the holonomic connection . In accordance with the relation (4.2.8), the desired
vector field S is an extension of the section TXo\
2
o of the fibre bundle T*Q >TQ
over the closed submanifold J
l
Q C TQ to a global section. Such an extension always
exists by virtue of Theorem 1.1.2, but is not unique. Then the dynamic equation
(4.2.2) can be written in the form
Itt - - I (
= 1 I
^
= ?
J
(4.2.9)
This is equivalent to the autonomous second order dynamic equation
t = Q, f = l, ? =S*', (4.2.10)
on Q. Being a solution of the equation (4.2.10), a curve c in Q also satisfies the
equation (4.2.9), and vice versa. QED
It should be emphasized that, written in the bundle coordinates (, q'), the second
order equation (4.2.10) is well defined with respect to any coordinates on Q.
QED
4.3 Dynamic connections
In order to say more than was said in Theorem 4.2.3, we will consider the relationship
between the holonomic connections on the jet bundle J
l
Q > R and the connections
on the affine jet bundle J
l
Q >Q (see Propositions 4.3.1 and 4.3.2 below).
Let
7 : J'Q - J
l
Q
J
l
Q
be a connection on the affine jet bundle J*Q Q. It takes the coordinate form
7 =<V(dA +7ldf),
(4.3.1)
4.3. DYNAMIC CONNECTIONS
163
with the transformation law
Tj = (VTi + *tf)f.
(4.3.2)
Remar k 4.3.1. In view of the canonical splitting (4.1.8), the curvature (1.4.8) of
a connection 7 (4.3.1) reads
R: ^Q-^AT'Q VQ,
J
l
Q
R = 2^dq
X
A dq 5, = {h?
ki
dt}
k
A dcf + R^dt A dq>) d
u
R\ =9A7" - d^\ + 7i a
j 7
; - yid^. (4.3.3)
Using the contraction (4.1.7), we obtain the soldering form
AJ.R =[(*,? +R^)dql - R^tfdt] a,
on the affine jet bundle J
l
Q Q. Its image by the canonical projection T'Q
V'Q is the tensor field
R:J
1
Q-*V*QVQ,
Q
R=(Ry^ + Rl
J
W^d
i
, (4.3.4)
and then we come to the scalar field
R:J
l
Q^ R,
R = R-liit +Ron
(4.3.5)
on the jet manifold J
l
Q.
PROPOSITION 4.3.1. Any connection 7 (4.3.1) on the affine jet bundle J
l
Q -* Q
defines the holonomic connection

7
=po
7
: J
l
Q^ J^J
X
Q^J
2
Q,
(4.3.6)
on the jet bundle J
l
Q ->R. D
e, =^+q^+(7i +qf7j)af.
164 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
Proof. Let us consider the composite fibre bundle (4.1.10) and the morphism p
(1.6.6) which reads
p : J^Q 3 (q\ q\, q\
t
) - . (q\ q\, q\
t)
= q\, =^ +4<?jt) ^ J'Q- (4-3-7)
A connection 7 (4.3.1) and the morphismp (4.3.7) combine into a desired holonomic
connection f
7
(4.3.6) on the jet bundle J
l
Q ->R. QED
It follows that every connection 7 (4.3.1) on the affine jet bundle J
1
Q Q
yields the dynamic equation
on the configuration bundle Q > R. This is precisely the restriction to J
2
Q of
the kernel KerD^of the vertical covariant differential D
7
(1.6.11) defined by the
connection 7:
D
y
: J'J'Q -> VQJ'Q,
q'tD-, = q'tt - 7o - <Al)-
(4.3.9)
(4.3.8)
(4.3.7)
Qu = 7o + <dYj
p : J'
Q
J
l
Q 3 (q\q
l
v
q\
t
) ~ (q\ql,q\
t)
= q\,qlt = 4 +4<?jt) ^ J*Q-
Therefore, connections on the jet bundle J
l
Q ><5 are called the dynamic connec-
tions. The corresponding equation (4.2.3) can be written in the form
C
1
= p o 7 o c,
where p is the morphism (4.3.7).
Of course, different dynamic connections can lead to the same dynamic equation
(4.3.8).
PROPOSI TI ON 4.3.2. Any holonomic connection (4.1.25) on the jet bundle J 'Q *
R defines the dynamic connection
7j =dt \d
t
+ (g - y
t
^)dj] + dq> [d, + jgjgg] (4.3.10)
on the affine jet bundle J
l
Q * Q.
Proof. Let be a holonomic vector field (4.1.25). Given an arbitrary vertical vector
field
u = a% +Vd',
4.3. DYNAMIC CONNECTIONS 165
on the jet bundle J
1
Q >R, let us put
h(u)
d
=[SMu)]-mM),
where v is the vertical endomorphism (4.1.13). Then, there is the endomorphism
h 4% + 9$ - . -q% + (q\ - cfd]V)d\.
It is readily observed that this endomorphism obeys the condition
hh = h-
Then, one can construct the projection
J< ?% + &! "(il-l?%?)%
Recall that a holonomic connection on J
l
Q >R defines the projection (1.4.3)
which reads
: TJ
l
Q 3 td
t
+ q% + q\d\ ->(q
l
- %)% + (41 ~ i?)$ VJ
l
Q.
Let us consider the composition of morphisms
id
t
+ q% + gjd? ~
This corresponds to the connection 7
5
(4.3.10) on the affine jet bundle J
l
Q > Q.
QED
It is readily observed that the dynamic connection 7^(4.3.10), denned by a
dynamic equation, possesses the property
7* = grj + gjgj
(4.3.11)
which implies the relation
atf = af-tf.
It(u) =-aa
t
+(b
{
- aiq&di
h-VJ
l
Q ^VJ'Q,
J ^ ^ + ldVJ'Q) :VJ
l
Q^
Q
V
Q
J
l
Q,
[Q\-m~\4^ )-\^ d]m-
/
?
o : TJ
l
Q - VJ'Q - V
Q
J
l
Q,
166 CHAPTER 4.
LAGRANGIAN TIME-DEPENDENT MECHANICS
Therefore, a dynamic connection 7, obeying the condition (4.3.11), is said to be
symmetric. Thetorsion of a dynamic connection 7 is defined as the tensor field
T:J
l
Q^V'QVQ,
Q
T =J *3j* dk,
(4.3.12)
It follows at once that a dynamic connection is symmetric if and only if its torsion
vanishes.
Let 7 be a dynamic connection (4.3.1) and
7
the corresponding dynamic equa-
tion (4.3.6). Then the dynamic connection (4.3.10) associated with the dynamic
equation
7
takes the form
^! = 5(7? +ahJ +&#), 7,S =" - ftSc?-
It is readily observed that 7 =7* if and only if the torsion T (4.3.12) of the dynamic
connection 7 vanishes.
Exampl e 4.3.2. Since the jet bundle J
1
Q >Q is affine, it admits an affine
connection
7 =dq
x
[9, +(
7
y O +7yOtf )di ]- (4-3.13)
This connection is symmetric if and only if 7^ =f'
x
. One can easily justify that an
affine dynamic connection generates a quadratic dynamic equation, and vice versa.
Nevertheless, a non-affine dynamic connection, whose symmetric part is affine, also
define a quadratic dynamic equation.
An affine connection (4.3.13) on the affine jet bundle J
1
Q * Q yields the linear
connection
7 = dq
X
[d
X
+lU<f)q(d
i
}
on the vertical tangent bundle VQ Q.
Using the notion of a dynamic connection, we may modify Theorem 3.1.4 as
follows. Let H be an autonomous second order dynamic equation on a manifold
M, and f=(4.2.6) the corresponding conservative dynamic equation on the bundle
R x M >R. The latter yields the dynamic connection 7 (4.3.10) on the fibre bundle
R x TM - R x M. R x TM ~*RxM.
7 = dq
X
[d
X
+lU<f)q(d
i
}
7=dq
x
[d, + (
7
yo+7yo<?n^].
(4.3.13)
Tt =7* - aho* - d"#tf
4.3. DYNAMIC CONNECTIONS 167
Its components 7! are exactly those of the connection (3.1.7) on the tangent bundle
TM >M in Theorem 3.1.4, while 7J make up a vertical vector field
e =T $ =(H
1
- i ^S^j a, (4.3.14)
on TM M. Thus, we have shown the following.
PROPOSI TI ON 4.3.3. Every autonomous second order dynamic equation 5 (3.1.4)
on a manifold M admits the decomposition
H
1
=Kjq* + e
{
where K is the connection (3.1.7) on the tangent bundle TM M, and e is the
vertical vector field (4.3.14) on TM - M.
Remar k 4. 3. 3. With a dynamic connection 75 (4.3.10), we also restate the well-
known linear connection on the tangent bundle TJ
X
Q >J 'Q, associated with a
dynamic equation on Q [132], and can write it with respect to holonomic bases.
Given a holonomic connection f (4.1.25), we have the corresponding horizontal
splitting (1.4.1) of the tangent bundle TJ
l
Q of J
l
Q:
TJ
l
Q = Hc VJ
l
Q,
J
l
Q
tit + q% + $% =tf +(<?' - t&dt + (4$ - f )flf,
(4.3.15)
(4.3.16)
and the corresponding horizontal splitting (1.4.2) of the cotangent bundle T'J
l
Q of
T"J
1
Q = R W' Q ,
tdt +qM + #M =(t +9t9i +{'$< +<?/ +9i(<*2f " Cdt),
(4.3.17)
Using these spUttings, one can introduce the non-holonomic bases (, d d\) for the
tangent bundle TJ
X
Q, and the non-holonomic bases
(dt,6',dql-Cdt)
for the cotangent bundle T'^Q of J
l
Q.
168 CHAPTER 4.
LAGRANGIAN TIME-DEPENDENT MECHANICS
Furthermore, the dynamic connection 7
{
(4.3.10), associated with the dynamic
equation , provides the corresponding splitting (1.6.10a) of the vertical tangent
bundle VJ
l
Q of J
l
Q:
VJ
l
Q = V
Q
J
l
Q //-,.,
q% + q\d\ =(q\ - i^fljf
1
)^+?'(a, + \&d[).
(4.3.18)
The splittings (4.3.16) and (4.3.18) combine into the splitting
TJ'Q = H,@ H
lt
V
Q
J
X
Q
(4.3.19)
together with the non-holonomic bases
(, h, =d
}
+ \d)Cdl a?)
(4.3.20)
for the tangent bundle TJ
l
Q [55]. Note that the fibre bundle H^ in the splitting
(4.3.19), like VQJ 'Q, is isomorphic to the pull-back
H
lt
= J
l
Q xVQ^ J 'Q, h, <>d
r
(4.3.21)
Accordingly, the dynamic connection 7
{
(4.3.10) determines the corresponding
splitting (1.6.10b) of the vertical cotangent bundle V'J
l
Q of J
l
Q. Combining this
splitting with the decomposition (4.3.17) gives the splitting
T'J
l
Q = R V'Q VAJ
l
Q
J
l
Q J
l
Q
together with the non-holonomic bases
{dt,e\dq\-^dq
x
)
for the cotangent bundle T'J
l
Q of J
l
Q.
Every connection 7 (4.3.1) on the affine jet bundle J
l
Q * Q gives rise to the
connection V7 (1.6.12) on the composite vertical tangent bundle
V
Q
J
l
Q - J
l
Q - Q.
This connection reads
Vy = dq
x
(dx +7iaf +^7iJ 'af).
4.3. DYNAMIC CONNECTIONS 169
Since J
l
Q Q is an affine bundle, the connection Vj, in turn, can be seen as the
composite connection (1.6.8) generated by the connection 7 on the affine jet bundle
J
1
Q >Q, and the linear connection (1.6.14):
7 =dq
x
(d
x
+ d)i\ct
t
d\) + dq
k
dl (4.3.22)
on the vertical tangent bundle VQJ
1
Q > J
l
Q. Due to the splitting (4.1.8), the
connection 7 (4.3.22) in fact is a linear connection
7 =dq
x
(d
x
+ d'-yitfdi) +dq
k
t
6
k
(4.3.23)
on the pull-back (4.3.21). In particular, if 7 = 75 is the connection (4.3.10), the
connection 7^(4.3.23) yields a connection on the pull-back bundle H-,
(
>J
l
Q in
the splitting (4.3.19).
Due to the splitting (4.3.19), the tangent bundle TJ
l
Q * J
l
Q can be provided
with a linear connection A [132] which is the direct sum of
the trivial connection on the linear bundle H^ >J
1
Q,
the connection 7^(4.3.23) on the pull-back bundle H
1(
> J 'Q,
and the similar connection (4.3.22) on the vertical tangent bundle VQJ
1
Q >
J
l
Q.
With respect to the non-holonomic basis (4.3.20), where
id
t
+cfdi + q\d\ =it, + u% + v*dl (4.3.24)
the above-mentioned linear connection A reads
A = dq
x
(d
x
+ d;7>
J
^ +
d
f^J^\
+
^
a
'
( 4 3 2 5 )
Recall the expression (4.3.10) for the components 7^of the dynamic connection 7
?
.
Substituting
q
i
= iq\+u
i
, q\ = ie + u
k
Y
k
+v
i
from (4.3.24) in (4.3.25), we obtain the linear connection A written with respect to
the holonomic bases (d
x
,df) for TJ
l
Q:
A = dq
x
(d
x
+ Ax) + dq\ (dl + A\),
Ax =i[-4dn\d, + (dxC - <drf
k
- 4l\^
k
x
+ <&~&drf
x
- ^DSf ] +
fld^A + (d
xl
) +
7
&7* - I'daDdW + fariM,
A\ = ifa + (d*? - q?dH)dt] + q>dh
k
di.
7 =dq* (d
x
+ dftqid!) + dq
k
dl (4.3.22)
(4.3.23)
7 =dc? (d
x
+ d'jxtfdi) +dq
k
d
k
(4.3.24)
(4.3.25)
q
i
= iq\+u
i
, q\ =tf +U
k
f
k
+ v*
A = dq
x
(d
x
+ d^W-^ +W ^ ) +
d
4 dl
A = dq
x
(d
x
+ A
x
) + dq\ (dl + A\),
Ax =i[-4dn\d, + (dxC - rf^j " <7?7^7
A
fc
+H ^ \ ~ S*af7i)4f] +
fldrfA + (d
xl
) +
7
&7* - 7^*7l )9|] + #9, 7^,
170 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
The covariant derivatives V relative to the connection A fulfill the conditions
VA(0 =VKO =o,
V,(5j) = -dtfA,
V
A
(^) = -CSTJ ^,
(4.3.26)
In particular, by virtue of the equality (4.3.26), every solution c of the dynamic
equation is also a geodesic of the linear connection A.
4.4 Non-r elat ivist ic geodesic equat ions
In this Section, we aim to show that every dynamic equation on a configuration
bundle Q > M is equivalent to a geodesic equation on the tangent bundle TQ Q.
We start from the relation between the dynamic connections 7 on the affine jet
bundle J
l
Q * Q and the connections
K = dq
x
(d
x
+ Kfa) (4.4.1)
on the tangent bundle TQ Q of the configuration space Q. We will use the
notation (1.2.4).
Let us consider the diagram
J
l
Q
J
l
Q
J
-J
l
Q
TQ
, I I .
J
l
Q --> TQ
(4.4.2)
where JQTQ is the first order jet manifold of the tangent bundle TQ Q, coordi-
nated by
(t,q\i,q\(i)^ (q%).
The jet prolongation over Q of the canonical imbedding A (4.1.4) reads
/'A : (t,q\ql,qU) - (t,q',i = l,f =g|, (t) =0,(q% = <,).
Then we have
J
J
A o
7
; (t, g', gj) ,_ (<, g\ t =l,g =q', (*) =0, (g% =-),
v|(5j) =V|(A,) =o, v*(a
3
) = -flftflj.
X o A : (t, q\ ql) (t, e?\ =!,? =flj, (t) =Kj, (% =/if*).
4.4. NON-RELATIVISTIC GEODESIC EQUATIONS 171
It follows that the diagram (4.4.2) can be commutative only if the components K
of the connection K (4.4.1) on the tangent bundle TQ >Q vanish.
Since the transition functions t * t* are independent of q', a connection
K = dq
x
(3
A
+iq3i) (4.4.3)
with A" =0 may exist on the tangent bundle TQ Q in accordance with the
transformation law
K'i-idjfKi + drfl^-y (4.4.4)
Now the diagram (4.4.2) becomes commutative if the connections 7 and K fulfill
the relation
^ = K^o X = Kl(t,q',t = l,q' = q^).\ (4.4.5)
It is easily seen that this relation holds globally because the substitution of q
l
= q
l
t
in (4.4.4) restates the transformation law (4.3.2) of a connection on the affine jet
bundle J
l
Q Q. In accordance with the relation (4.4.5), the desired connection K
is an extension of the section J 'A o 7 of the affine jet bundle JQTQ * TQ over the
closed submanifold J
l
Q C TQ to a global section. Such an extension always exists
by virtue of Theorem 1.1.2, but is not unique. Thus, we have proved the following.
PROPOSI TI ON 4.4.1. In accordance with the relation (4.4.5), every dynamic equa-
tion on a configuration bundle Q >R can be written in the form
ql
t
= K'
0
\ + q{K>o\,\ (4.4.6)
where 7f is a connection (4.4.3) on the tangent bundle TQ Q. Conversely, each
connection K (4.4.3) on TQ >Q defines the dynamic connection 7 (4.4.5) on the
affine jet bundle J
X
Q > Q and the dynamic equation (4.4.6) on aconfiguration
bundle Q - R.
Then we come to the following theorem.
THEOREM 4.4.2. Every dynamic equation (4.2.2) on a configuration bundle Q - R
is equivalent to the geodesic equation
g
0
=0, q = 1,
i? = K\{<f,q)q\ (4.4.7)
K"x = (d
J
qKi
+
dq*)^-
x
.
K = dq
x
(d
x
+ K\di)
(4.4.4)
(4.4.5)
(4.4.3)
(4.4.7)
if =0, q = 1,
? = Ki{(f,q)e,
172 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
on the tangent bundle TQ relative to a connection K with the components K =0
and K\ (4.4.5). Its solution is a geodesic curve in Q which also obeys the dynamic
equation (4.4.6), and vice versa.
In accordance with this theorem, the second order equation (4.2.10) in Theorem
4.2.3 can be chosen as a geodesic equation. It should be emphasized that, written
in the bundle coordinates (,<?'), the geodesic equation (4.4.7) and the connection
K (4.4.5) are well defined with respect to any coordinates on Q.
As was mentioned, from the physical viewpoint, the most interesting dynamic
equations are the quadratic ones
? = ^(Qniti + b)(qnii + f'(q")-
(4.4.8)
This property is global due to the transformation law (4.2.5). Then one can use the
following two facts.
PROPOSI TI ON 4.4.3. There is one-to-one correspondence between the affine con-
nections 7 on the affine jet bundle J 'Q Q and the linear connections K (4.4.3)
on the tangent bundle TQ Q. D
Proof. This correspondence is given by the relation (4.4.5), written in the form
< =
7Jo +lU<& = Vo(?")t +KJiWW]^^* =
K;
0
{q") + K*i{<f)4
t
i.e.,
7^>
KJ\.
QED
In particular, if an affine dynamic connection 7 is symmetric, so is the corre-
sponding linear connection K.
COROLLARY 4.4.4. Every quadratic dynamic equation (4.4.8) on a configuration
bundle Q R of non-relativistic mechanics gives rise to the geodesic equation
if = 0, 7 = 1,
9" = J
jk
(q
t
'W<i
k
+ l>)(q''W<i
0
+
/V)V
(4.4.9)
4.4. NON-RELATIVISTIC GEODESIC EQUATIONS 173
on the tangent bundle TQ with respect to the symmetric linear connection
Kx\ =0, ^ o' o = / ' i
KOJ = -b),
K
kj = a'kj
(4.4.10)
on the tangent bundle TQ Q.
The geodesic equation (4.4.9), however, is not unique for the dynamic equation
(4.4.8).
PROPOSI TI ON 4.4.5. Any quadratic dynamic equation (4.4.8), being equivalent to
the geodesic equation with respect to the symmetric linear connection K (4.4.10),
is also equivalent to the geodesic equation with respect to an affine connection K'
on TQ >Q which differs from K (4.4.10) in a soldering form a on TQ >Q with
the components
"S*o,
4 =A* +(s - l)h'
k
q, a'
0
= -shlq* - htf + hi
where s and h\ are local functions on Q. D
Proof. The proof follows from direct computation.
Proposition 4.4.5 can also be deduced from the following lemma.
LEMMA 4.4.6. Every affine vertical vector field
QED
a =
irm+bWWttf
(4.4.11)
on the affine jet bundle J
l
Q Q is extended to the soldering form
a = {f'dt + b\dq
k
) di
(4.4.12)
on the tangent bundle TQ > Q.
Proof. Similarly to Proposition 4.4.3, one can show that there is one-to-one corre-
spondence between the VQ J 'Q-valued affine vector fields (4.4.11) on the jet manifold
J '<3 and the VgTQ-valued linear vertical vector fields
a =
[{?)# + ftf)?\di
on the tangent bundle TQ. This linear vertical vector field determines the desired
soldering form (4.4.12). QED
174 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
In Section 6.4, we will use Theorem 4.4.2, Corollary 4.4.4 and Proposition 4.4.5
in order to study the relationship between non-relativistic and relativistic equations
of motion.
Now let us extend our inspection of dynamic equations to connections on the
tangent bundle TM M of the typical fibre M of a configuration bundle Q >R.
In this case, the relationship fails to be canonical, but depends on a trivialization
(4.1.1) of Q-*R.
Given such a trivialization, let (t.g*) be the associated coordinates on Q, where if
are coordinates on M with transition functions independent of t. The corresponding
trivialization (4.1.3) of J
X
Q R takes place in the coordinates (t,q',q ), where q
are coordinates on TM. With respect to these coordinates, the transformation law
(4.3.2) of a dynamic connection 7 on the affine jet bundle J
i
Q Q reads
-v"
1 0 dq'
% H
(
dq'
ln
+
at*-
dq"
It follows that, given a trivialization of Q >R, a connection 7 on J 'Q >Q defines
the time-dependent vertical vector field
dq
R x TM -* VTM
and the time-dependent connection
dq
k

(W
+
*
(t,<?,if)
) dq'
d
RxTM J
X
TM C TTM (4.4.13)
on the tangent bundle TM M.
Conversely, let us consider a connection
K dq
h

(-
+
Ktf,$)
d
d?
)
on the tangent bundle TM >M. Given the above-mentioned trivialization of the
configuration bundle Q >R, the connection K defines the connection K (4.4.3),
with the components
K
0. KI
K
k
,
on the tangent bundle TQ >Q. The corresponding dynamic connection 7 on the
affine jet bundle J
1
Q > Q reads
%
0,
7i
r
fc
.
(4.4.14)
4.5. REFERENCE FRAMES
175
Using the transformation law (4.3.2), one can extend the expression (4.4.14) to
arbitrary bundle coordinates (t, q') on the configuration space Q as follows:
it
^K&(q
r
),it(Q
r
,q
r
t
))
+
dq^dq^
+
an
dkT,
%
d
t
r
9^4- 7ir*,
where
P
$*(*,?)
(4.4.15)
is the connection on Q - R, corresponding to a given trivialization of Q, i.e., P =0
relative to (t,(f). The dynamic equation on Q defined by the dynamic connection
(4.4.15) takes the form
<? =
d
t
r +qid^r
+ iM - r
A
). (4.4.16)
By construction, it is a conservative dynamic equation. Thus, we have proved the
following.
PROPOSI TI ON 4.4.7. Any connection K on the typical fibre M of a configuration
bundle Q K yields a conservative dynamic equation (4.4.16) on Q. O
4.5 Reference frames
From the physical viewpoint, a reference frame in non-relativistic mechanics deter-
mines a tangent vector at each point of a configuration space Q, which characterizes
the velocity of an "observer" at this point. This speculation leads us to suggest the
following mathematical definition of a reference frame in non-relativistic mechanics
[57, 132, 161].
DEFI NI TI ON 4.5.1. In non-relativistic mechanics, a reference frame is a connection
T on the configuration bundle Q * K.
In accordance with this definition, one can think of the horizontal vector field
(4.1.16), associated with a connection r on Q > R, as being a family of "observers",
while the corresponding covariant differential
q^D
r
(qt) = ql-r
176 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
determines the relative velocities with respect to the reference frame I \
In particular, given a motion c : R Q, its covariant derivative V
r
c (1.4.6)
with respect to a connection T is the velocity of this motion relative to the reference
frame T. For instance, if c is an integral section for the connection T, the velocity
of the motion c relative to the reference frame T is equal to 0. Conversely, every
motion c : R > Q, defines a reference frame T
c
such that the velocity of c relative
to T
c
vanishes. This reference frame T
c
is an extension of the section c(R) > J
l
Q
of the affine jet bundle J
l
Q > Q over the closed submanifold c(R) G Q to a global
section in accordance with Theorem 1.1.2.
Remark 4. 5. 1. It should be emphasized that the vertical tangent bundle VQ of a
configuration bundle Q, but not the jet manifold J
:
Q plays the role of the "space
of coordinates and velocities", while elements of J
1
Q may be termed the absolute
velocities. In the universal unit system, elements of VQ, however, have the same
physical dimension [q] as elements of Q, whereas absolute velocities are of physical
dimension [g] 1.
By virtue of Corollary 4.1.1, any reference frame T on a configuration bundle
Q >R is associated with an atlas of local constant trivializations, and vice versa.
The connection T reduces to
r = d
t
(4.5.1)
with respect to the corresponding coordinates (t,?
1
), whose transition functions
q
1
> q" are independent of time. One can think of these coordinates as being
also the reference frame, corresponding to the connection (4.5.1). They are called
the adapted coordinates to the reference frame I \ Thus, we come to the following
definition, equivalent to Definition 4.5.1.
DEFI NI TI ON 4.5.2. In non-relativistic mechanics, a reference frame is an atlas of
local constant trivializations of a configuration bundle Q > R.
In particular, with respect to the coordinates if adapted to a reference frame T,
the velocities relative to this reference frame are equal to the absolute ones
D
r(q't) =fr =5'c
A reference frame is said to be compJ ete if the associated connection F is com-
plete. By virtue of Proposition 4.1.2, every complete reference frame defines a
trivialization of a bundle Q >R, and vice versa.
Remark 4. 5. 2. Given a reference frame T, one should solve the equations
r'M(
t
,f)) = ,
(4.5.2a)
a^a^v,
+
56
(4.5.2b)
in order to find the coordinates {t,q
a
) adapted to T.
Let (i, g") and ((, g
2
) be the adapted coordinates for reference frames Tj and r
2
,
respectively. In accordance with the equality (4.5.2b), the components r\ of the
connection 1^with respect to the coordinates (t, g
2
) and the components r
2
of the
connection T
2
with respect to the coordinates (t, g") fulfill the relation
Mr- +r
a
- 0

Using the relations (4.5.2a) - (4.5.2b), one can rewrite the coordinate transfor-
mation law (4.2.5) of dynamic equations as follows. Let
?u=T
(4.5.3)
be a dynamic equation on a configuration space Q, written with respect to a reference
frame (t,if
1
). Then, relative to arbitrary bundle coordinates (t,q
l
) on Q >R, the
dynamic equation (4.5.3) takes the form
ql
t
= d
t
r
i
+ djr{qi-P)
-g^ -
r
' -
r ,
>+f ^
(4.5.4)
where T is the connection corresponding to the reference frame (t,q"). The dynamic
equation (4.5.4) can be expressed in the relative velocities <j{. =q\ T' with respect
to the initial reference frame (t,q
a
). We have
*t-V4-gij?**+&M-*>-
(4.5.5)
Accordingly, any dynamic equation (4.2.2) can be expressed in the relative velocities
<jf, =q\ P with respect to an arbitrary reference frame T as follows:
Mi =( - Jry
t
= e- ckr,
(4.5.6)
4.5. REFERENCE FRAMES 177
178 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
where jr is the prolongation (4.1.24) of the connection T onto the jet bundle J
l
Q *
R.
For instance, let us consider the following particular reference frame T for a
dynamic equation f. The covariant derivative of a reference frame T with respect
to the corresponding dynamic connection 7^(4.3.10) reads
V
7
r =Q - T'Q x V
Q
J'Q,
wr = 'v}r
k
dq
x
d
k
,
vjr* =a
A
r*-7$or.
(4.5.7)
A connection T is called ageodesic reference frame for the dynamic equation f if
rjv^r
r
A
(d
A
r
k
- 7* r) (diC - c o r)di
o.
(4.5.8)
PROPOSI TI ON 4.5.3. Integral sections c of a reference frame T are solutions of a
dynamic equation f if and only if T is a geodesic reference frame for f.
Proof. The proof follows at once from substitution of the equality (4.5.8) in the
dynamic equation (4.5.6). QED
Remark 4.5.3. The left- and right-hand sides of the equation (4.5.6) separately are
not well-behaved objects. This equation will be brought below into the covariant
form (4.7.6).
Reference frames play a prominent role in many constructions of time-dependent
mechanics. In particular, we obtain a converse of Theorem 4.4.2.
THEOREM 4.5.4. Given a reference frame T, any connection K (4.4.1) on the
tangent bundle TQ >Q defines a dynamic equation
e =(K\ - rK)q* \.
0=l
.,
=q
, .
D
This theorem is a corollary of Proposition 4.4.1 and the following lemma.
LEMMA 4.5.5. Given a connection T on the fibre bundle Q- >R and a connection
K on the tangent bundle TQ -* Q, there is the connection K on TQ -* Q with the
components
*A =0,
K\ = K\ - VKl
4.6. FREE MOTION EQUATIONS 179

Proof. The proof follows from the inspection of transition functions. QED
4.6 Free motion equations
Let us point out the following interesting class of dynamic equations which we agree
to call the free motion equations.
DEFI NI TI ON 4.6.1. We say that the dynamic equation (4.2.2) is a free motion
equation if there exists a reference frame ((,?*) on the configuration space Q such
that this equation reads
Tu = 0-
(4.6.1)

With respect to arbitrary bundle coordinates (t, <?'), a free motion equation takes
the form
* = 4P +
dj
r(<d - V) - $^-
k
(4 - *)<<* - n
(4.6.2)
where P = d
t
g
t
(t,'^) is the connection associated with the initial frame (t,q') (cf.
(4.5.4)). One can think of the right-hand side of the equation (4.6.2) as being the
general coordinate expression for an inertial force in non-relativistic mechanics. The
corresponding dynamic connection 7^on the affine jet bundle J 'Q Q reads
7i =d
k
r -
dq
m
dq3dq
k{ql
>'
75 =^ + ^- 7 * 1 * .
(4.6.3)
It is affine. By virtue of Proposition 4.4.3, this dynamic connection defines a linear
connection K on the tangent bundle TQ >Q, whose curvature necessarily vanishes.
Thus, we come to the following criterion of a dynamic equation to be a free motion
equation.
PROPOSI TI ON 4.6.2. If is a free motion equation on a configuration space Q,
it is quadratic, and the corresponding symmetric linear connection (4.4.10) on the
tangent bundle TQ Q is a curvature-free connection.
180 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
This criterion is not a sufficient condition because it may happen that the com-
ponents of a curvature-free symmetric linear connection on TQ Q vanish with
respect to the coordinates on Q which are not compatible with the fibration Q > R.
The similar criterion involves the curvature of a dynamic connection (4.6.3) of a
free motion equation.
PROPOSI TI ON 4.6.3. If f is a free motion equation, then the curvature R (4.3.3) of
the corresponding dynamic connection 7^is equal to 0, and so are the tensor field
R (4.3.4) and the scalar field R (4.3.5). D
Proposition 4.6.3 also fails to be a sufficient condition. If the curvature R (4.3.3)
of a dynamic connection 7^vanishes, it may happen that components of 7^are
equal to 0 with respect to non-holonomic bundle coordinates on the affine jet bundle
J
l
Q-^Q.
Nevertheless, we can formulate the necessary and sufficient condition of the ex-
istence of a free motion equation on a configuration space Q.
PROPOSI TI ON 4.6.4. A free motion equation on a fibre bundle Q > R exists if and
only if the typical fibre M of Q admits a curvature-free symmetric linear connection.
D
Proof. Let a free motion equation take the form (4.6.1) with respect to some atlas
of local constant trivializations of a fibre bundle Q > R. By virtue of Proposition
4.3.2, there exists an affine dynamic connection 7 on the affine jet bundle J
l
Q Q
whose components relative to this atlas are equal to 0. Given a trivialization chart
of this atlas, the connection 7 defines the curvature-free symmetric linear connection
(4.4.13) on M. The converse statement follows at once from Proposition 4.4.7. QED
The free motion equation (4.6.2) is simplified if the coordinate transition func-
tions q
1
* q
1
are affine in the coordinates <p. Then we have
q
x
tt
= d
t
P - PdjP + 2q
1
t
d
]
r.
(4.6.4)
Exampl e 4.6.1. Let us consider a free motion on a plane R
2
. The corresponding
configuration bundle is R
3
- R, coordinated by (t,r). The dynamic equation of
this motion is
f = 0.
(4.6.5)
4.6. FREE MOTION EQUATIONS 181
Let us choose the rotatory reference frame with the adapted coordinates
r = AT,
A _ /coswt- si nwA
\ sin art cos art )
(4.6.6)
Relative to these coordinates, the connection T corresponding to the initial reference
frame reads
r =d
t
r = d
t
A A-
X
T.
Then the free motion equation (4.6.5) with respect to the rotatory reference frame
(4.6.6) takes the familiar form
r
M
=
M" )*
(4.6.7)
The first term in the right-hand side of the equation (4.6.7) is the centrifugal force
(PojjT'), while the second one is the Coriolis force (2q\djY
x
).
The following lemma shows that the free motion equation (4.6.4) is affine in the
coordinates q' and q\.
LEMMA 4.6.5. Let (,<f) be a reference frame on a configuration bundle Q R
and T the corresponding connection. Components P of this connection with respect
to another coordinate system (i, q
%
) are affine functions in the coordinates q' if and
only if the transition functions between the coordinates q and q' are affine.
Proof. If
q* = o(t)g +b'(t),
then we have
rHO^a^- f rO + d^.
Conversely, let
r = c)(t)q> + s'(t)
such that the homogeneous linear system of differential equations
a
J
<?
i
() =c'(ty() +s'W
(4.6.8)
182 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
has a solution on R. Let a)(t) be the mxm matrix whose columns are m fundamental
solutions qi{t),--,q
m
(t) of the equations (4.6.8) (see, e.g., [22]). This matrix is
invertible for all t 6 R. Then the connection T is associated with the coordinates
((,5") such that
9
i
=oi (*)r+6' (o,
where functions b
l
(t) fulfill the equations
d
t
V =( 0W" +*'
QED
One can easily find the geodesic reference frames for the free motion equation
lit =o.
(4.6.9)
They are F* = V* = const. By virtue of Lemma 4.6.5, these reference frames define
the adapted coordinates
5* =V
jq
> - v't - a\ k'j = const., v
x
= const., a' =const. (4.6.10)
The equation (4.6.9) obviously keeps its free motion form under the transformations
(4.6.10) between the geodesic reference frames. It is readily observed that these
transformations are precisely the elements of the Galilei group.
4.7 Relative acceleration
In comparison with the notion of a relative velocity, that of a relative acceleration
is more intricate.
To consider a relative acceleration with respect to a reference frame T, one
should prolong the connection T on the configuration bundle Q * R to a holonomic
connection
r
on the jet bundle J
l
Q ->R. Note that the jet prolongation JV (4.1.24)
of r onto J
l
Q R is not holonomic. We can construct the desired prolongation
by means of a dynamic connection 7 on the affine jet bundle J
l
Q >Q.
LEMMA 4.7.1. Let us consider the composite bundle (4.1.10). Given a frame T
on Q > R and a dynamic connections 7 on J 'Q - Q, there exists a dynamic
connection 7 on J 'Q Q with the components
7i = 7i,
7; =tkr - Y
k
r
k
.
(4.7.1)
4.7. RELATIVE ACCELERATION
183
D
Proof. Combining the connection r on Q ->R and the connection 7 on J
l
Q -> Q
gives the composite connection (1.6.8) on J
l
Q ->R which reads
B =dt (ft +Pft +(
7
jp +7*)ft).
Let j r be the jet prolongation (4.1.24) of the connection T on J
l
Q -* R. Then the
difference
JT-B = dt {dtV -
7
'r* - 7J )9f
is a VQJ 'Q-val ued soldering form on the jet bundle J
1
Q - R, which is also a
soldering form on the affine jet bundle J
l
Q -+Q. The desired connection (4.7.1) is
7 =
7
+j r - B =dt (ft +(d
(
r - 7*r*)ft) +A?* (ft +
7
j$).
QED
Now, we construct a certain soldering form on the affine jet bundle J
l
Q ^ Q
and add it to this connection. Let us apply the canonical projection T'Q >V'Q
and then the imbedding T : V'Q -> T'Q to the covariant derivative (4.5.7) of
the reference frame T with respect to the dynamic connection 7. We obtain the
VQJ
1
Q- valued 1-form
a = [-P(ftP -
7
* o r)dt + (ftp -
7
* o r)d
q
'} ft
on Q whose pull-back onto J
l
Q is the desired soldering form. The sum
def~ .
7r =7 +,
called the frame connection, reads
7r{, =*r* -
7
J r* - p(ftp -
7
j o r),
7ri =7l +d
k
r - 7* o r.
(4.7.2)
This connection yields the desired holonomic connection
& =ckv +(ftp +y
t
-
7
' o r)(q* - r")
on the jet bundle J
X
Q - R.
184 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
Let be a dynamic equation and 7 =7^the connection (4.3.10) associated with
Then one can think of the vertical vector field
ar=g-fr =(f-f
r
)ff
(4.7.3)
on the affine jet bundle J
l
Q Q as being a relative acceleration with respect to
the reference frame T in comparison with theabsolute acceleration .
For instance, let us consider a reference frame which is geodesic for the dynamic
equation , i.e., the relation (4.5.8) holds. Then the relative acceleration of a motion
c with respect to the reference frame T is
tt - fr) r = 0.
Let now be an arbitrary dynamic equation, written with respect to coordinates
(t, q
l
) adapted to the reference frame T, i.e., P =0. In these coordinates, the relative
acceleration with respect to the reference frame T is
4 = eW. <A) - \<}t(d
k
c - d
k
c i^
=0
).
(4.7.4)
Given another bundle coordinates {t,q") on Q * R, this dynamic equation takes
the form (4.5.5), while the relative acceleration (4.7.4) with respect to the reference
frame T reads
4 = S,g''4.
Then we can write a dynamic equation (4.2.2) in the form which is covariant under
coordinate transformations:
P-rrgl =^91 ~r =
a
r,
(4.7.5)
where D
7r
is the vertical covariant differential (4.3.9) with respect to the frame
connection 7r (4.7.2) on the affine jet bundle J 'Q Q.
In particular, if is a free motion equation which takes the form (4.6.1) with
respect to a reference frame V, then
D^q\ =0
relative to arbitrary bundle coordinates on the configuration bundle Q >R.
4.7. RELATIVE ACCELERATION 185
The left-hand side of the dynamic equation (4.7.5) can also be expressed in the
relative velocities such that this dynamic equation takes the form
^9r - yr'kQr = a
r
(4.7.6)
which is the covariant form of the equation (4.5.6).
The concept of a relative acceleration is understood better when we deal with
the quadratic dynamic equation , and the corresponding dynamic connection 7 is
affine.
LEMMA 4.7.2. If a dynamic connection 7 is affine, i.e.,
7A 7AO +7Ak9t!
so is a frame connection 7
r
for any frame I \
Proof. The proof follows from direct computation. We have
1
rl = d
t
r + (d
J
r
i
-
1
'
k]
r
k
)(<ti-n,
7ri =a
t
r +7i
J
(^-P)
or
7r}k 7j*>
TTofc
&r-7-
f c
F,
7rio
= dkT*-T^P,
(4.7.7)
-rr
l
oo = d
t
r-Pd
J
r + Y
}k
rr
k
.
QED
In particular, we obtain
7r}/t = Y
]k
, 7rok =7r
l
M) =7rw =0
relative to the coordinates adapted to a reference frame T.
A glance at the expression (4.7.7) shows that, if a dynamic connection 7 is
symmetric, so is a frame connection jp.
COROLLARY 4.7.3. If a dynamic equation is quadratic, the relative acceleration
a
r
(4.7.3) is always affine, and it admits the decomposition
a
'
r
=-(r
A
vir +2^vir),
(4.7.8)
186 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
where 7 =7c is the dynamic connection (4.3.10), and
9r =9t - T\ 9? =1,
r =i,
is the relative velocity with respect to the reference frame T.
Note that the splitting (4.7.8) gives a generalized Coriolis theorem. In particu-
lar, the well-known analogy between inertia! and electromagnetic forces is restated.
Corollary 4.7.3 shows that this analogy can be extended to arbitrary quadratic dy-
namic equation.
4.8 Lagrangian systems
A velocity phase space J
l
Q of time-dependent mechanics does not admit any canon-
ical structure mentioned in the previous Chapter. Since J
1
Q is an odd-dimensional
manifold, it has no symplectic structure, whereas presymplectic and Poisson struc-
tures are defined only for Lagrangian systems, and depend on a Lagrangian. By
a Lagrangian system is meant a mechanical system whose motions are solutions of
Lagrange equations for some Lagrangian on a velocity phase space J 'Q. Obviously,
a mechanical system is not necessarily Lagrangian, whereas Lagrange equations are
not necessarily dynamic equations. Moreover, it may happen that Lagrange equa-
tions fail to be differential equations in a strict sense (see Definition 1.3.3). Note that,
in the framework of Lagrangian formalism, one also meets Cartan and Hamilton-De
Donder equations, besides the Lagrange one. These equations are equivalent to each
other and to Lagrange equations in the case of a regular Lagrangian.
A Lagrangian of a mechanical system is defined as a horizontal density
L = dt,
C : J
l
Q ->R,
(4.8.1)
on the velocity phase space J
l
Q, where is called a Lagrangian function or simply
a Lagrangian if there is no danger of confusion. With respect to the universal unit
system, a Lagrangian (4.8.1) is physically dimensionless.
Here, we do not study the calculus of variations in depth, but apply in a straight-
forward manner the first variational formula [57].
Let us consider a projectable vector field
u =u
l
d
t
+ u'di, u =0,1, (4.8.2)
4.8. LAGRANGIAN SYSTEMS 187
on a configuration bundle Q >M and calculate the Lie derivative of a Lagrangian
(4.8.1) along the jet prolongation
u = tfdt + u
{
d, + dtu%
dt = dt + q\di + q\
t
dl
(1.3.7) of u. We obtain
l^L = (u\dC)dt = (u'dt + u% + dtu'dDCdt. (4.8.3)
The first variational formula provides the following canonical decomposition of
the Lie derivative (4.8.3) in accordance with the variational problem:
u\dC = (u' - u'qpS, + d
t
{u\H
L
), (4.8.4)
where
H
L
= v'idL) + L = n
t
dq
{
- {ir
iq
\ - C)dt (4.8.5)
is the Poincare-Cartan form (see (4.1.14)), and
L : J
2
Q - VQ,

L
= S.dq' = (d, - dtdPCdtf (4.8.6)
is the Euler-La.gra.nge operator for a Lagrangian L. The latter can be seen as a
2-form
Si = (d< - dtdl)Cdq
%
A dt.
Its coefficients Si are called variational derivatives. We will use the notation
iTi =d\C, 7r
Jt
= d\d\C
The kernel Ker
L
C J
2
Q of the Euler-Lagrange operator (4.8.6) defines the
system of second order differential equations on Q
(ft - dtdj)C =0,
(4.8.7)
caUed the Lagrange equations. Its solutions are (local) sections c of the fibre bundle
Q R whose second order jet prolongations c five in (4.8.7). They obey the
equations
d
t
C o c -y{^i o c) =0.
dt
(4.8.8)
188
CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
Remark 4. 8. 1. The kernel of an Euler-Lagrange operator Kerj, fails to be a
subbundle of the fibre bundle J
2
Q R in general. Therefore, it may happen that, as
was mentioned above, the Lagrange equations (4.8.7) are not differential equations
in a strict sense. In particular, by virtue of Proposition 1.3.4, it is a differential
equation when the morphism (4.8.6) is of constant rank, e.g., if a Lagrangian L is
regular. Recall that a Lagrangian L is called regular (non-degenerate), if
det 7r^^0
everywhere on the velocity phase space J
1
Q. If a Lagrangian L is non-degenerate,
the Lagrange equations can be solved algebraically for second order derivatives, and
are equivalent to a dynamic equation.
Example 4.8.2. Let Q = R
2
K be a configuration space, coordinated by
{t,q). The corresponding velocity phase space J
l
Q is equipped with the adapted
coordinates (t,q,q
t
). The Lagrangian
L = -q
2
q
2
dt
on J
l
Q leads to the Euler-Lagrange operator
L = \qq
2
t ~ d
t
(q
2
q
t
)]dq
whose kernel is not a submanifold at the point q =0.
Every Lagrangian L on the jet manifold J
l
Q yields the Legendre map
L:J
l
Q^ V'Q,
P, O L = 7T,,
(4.8.9)
where (t,g',Pi) are coordinates on the vertical cotangent bundle V'Q. Indeed, due
to the vertical splitting (4.1.8), the vertical tangent morphism VL to L yields the
linear morphism
VL: J
l
Qx VQ-+R
and, consequently, the morphism (4.8.9).
In time-dependent mechanics on a configuration space Q R, the vertical
cotangent bundle V'Q plays the role of a momentum phase space (see Chapter 5).
4.8. LAGRANGIAN SYSTEMS
189
Remark 4. 8. 3. The Legendre map (4.8.9) is a local diffeomorphism, i.e., it is of
maximal constant rank if and only if a Lagrangian L is regular. A Lagrangian L is
called hypenegular if the Legendre map L is a diffeomorphism.
In Section 5.1, we will show that the vertical cotangent bundle V'Q is provided
with the canonical 3-form
ft
d
=d
Pi
A dq' A dt,
(4.8.10)
derived from the polysymplectic form (2.9.9). Let us consider the pull-back
il
L
= L'U = d-Ki A dq
i
A dt
(4.8.11)
on J
l
Q of the canonical form (4.8.10) by means of the Legendre map L (4.8.9).
The form fi^ provides Lagrangian formalism with the construction similar to Ha-
miltonian mechanics, but which depends on the choice of a Lagrangian L. In the
framework of this construction, the Lagrangian counterpart of a Hamiltonian form
is the Poincare-Cartan form (4.8.5).
If a Lagrangian L is regular, the formQ,
L
(4.8.11) defines a Poisson structure on
the jet manifold J
i
Q as follows. By means of f2
L
, every vertical vector field
d = tf% + &%
on the jet bundle J
l
Q >R corresponds to the 2-form
tfjftz, = {[<5%, +^{djin - BkKjffl ~ $%dqi} A dt.
This is one-to-one correspondence if a Lagrangian L is regular. Indeed, given an
arbitrary 2-form
<t> = [4>idq
i
+ fadql) A dt
on J
l
Q, the algebraic equations
&nj
i
+#
>
(d
i
n
i
- dtirj) = <k,
-#Kfi =<Pj
have a unique solution
& = -{*-ifL,
$
=
(
T
-i)*[^+( j r
1
) *"^** - ftr*)].
190
CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
In particular, every function / on J
l
Q determines a vertical vector field
#
f
= fr-yfye, - (n-
l
y%f + (Tr-^&jid^ - di^d)
(4.8.12)
on J
1
Q > K in accordance with the relation
d,\U
L
= -df Adt.
Then the Poisson bracket
{ /,$} x.*=*
f
J*/j ni . f,
9
eO(J
l
Q),
{f,9}i = (*-
l
Y
3
(difd
J
g-d
t
i
gd
]
f) +
(ft.r
i k
-Sl
k
ir
n
)(ir-
1
)(ir-
1
)^/ajff,
(4.8.13)
can be defined on the space D(J
l
Q) of functions on the jet manifold J
l
Q. In
particular, the vertical vector field d
}
(4.8.12) is the Hamiltonian vector field for
the function / with respect to the Poisson structure (4.8.13). The Poisson structure
(4.8.13) defines the corresponding symplectic foliation on the jet manifold J
1
Q,
which coincides with the fibration J
l
Q >R. The symplectic form on the leaf J]Q
of this foliation is
f2
t
= dir
x
A dq'
[57, 182].
Example 4.8.4. If a Lagrangian L is hyperregular, the Poisson structure (4.8.13) is
isomorphic to the canonical Poisson structure on the momentum phase space V'Q
(see Section 5.1). Indeed, the Poisson bracket (4.8.13) reads
{Wj} = {q',q
i
} = 0, {7r,,g
,
} =ff.

Remark 4.8.5. If a Lagrangian is degenerate, one may try to introduce a Poisson
structure on J
l
Q corresponding to the presymplectic form dH
L
(see (4.8.23) below)
in accordance with the procedure suggested in [49] (see Remark 2.5.1).
DEFI NI TI ON 4.8.1. A connection
ZL = dt + f%+edl
4.8. LAGRANGIAN SYSTEMS 191
on the jet bundle J 'Q H is said to be a Lagrangian connection for the Lagrangian
L if it obeys the equation
qj n
L
=d//
L
(4.8.14)
which takes the coordinate form
(/' - #" * = o,
a, - a
(
7r, - fd^i - e*
it
+ {p - 4)d^ = o.
(4.8.15a)
(4.8.15b)

Exampl e 4. 8. 6. A glance at the equations (4.8.15a) - (4.8.15b) shows that, for a
regular Lagrangian L, a Lagrangian connection is necessarily holonomic and unique.
In order to clarify the meaning of the equation (4.8.14), let us consider the
Lagrangian
C
d
^C(t,q',qi) + (ql
t)
-qi)n
l
(t,
q
\ql)
on the repeated jet manifold J
1
J
i
Q. The corresponding Euler-Lagrange operator,
called the Euler-Lagrange-Cartan operator, reads

z
: J
l
J'Q -* T'J'Q,

T
= [(diC + diTTjiq!^ -qi)- dtn,)dq' + 7r
0
(^
()
- q})dq\} A dt, (4.8.16)
dt = d
t
+ q\
t)
di + q\
t
d\.
Then the equation (4.8.14) is equivalent to the condition
UiJ'Q) C K er
r
which is the first order differential equation on the jet manifold J 'Q, called the
Cartan equations. They read
*v(4t) ~ d) = -
(4.8.17a)
diC - dtiTi + {q
j
{t)
- qi)diir
}
=0.
Remark 4. 8. 7. In Section 5.2, we will show that the Euler-Lagrange-Cartan
operator (4.8.16) is the Lagrangian counterpart of a Hamilton operator.
(4.8.17b)
192 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
Solutions of the Cartan equations (4.8.17a) - (4.8.17b) are (local) integral sec-
tions :(} J
l
Q of Lagrangian connections &, for a Lagrangian L, i.e., J
l
c = LC.
Furthermore, the equation (4.8.14) for these integral sections is equivalent to the
relation
?{d\dH
L
) =0
(4.8.18)
which is assumed to hold for any vertical vector field d on the jet bundle J
l
Q R.
It is easily seen that the restriction
Ker
L
=J
2
Q nK er
r
(4.8.19)
of the Euler-Lagrange-Cartan operator j (4.8.16) to the second order jet manifold
J
2
Q C J
l
J
l
Q recovers the Euler-Lagrange operator
L
(4.8.6) for the Lagrangian
L. Therefore, the Lagrange equations (4.8.7) are equivalent to the Cartan equations
(4.8.17a) - (4.8.17b) on the holonomic sections c = J
l
c of the jet bundle J
l
Q ->R.
It follows that solutions of the Lagrange equations (4.8.7) are integral sections of
holonomic Lagrangian connections for the Lagrangian L, which take their values into
the kernel of the Euler-Lagrange operator (4.8.19). Different holonomic Lagrangian
connections lead to different dynamic equations associated with the same system of
Lagrange equations.
In the case of a regular Lagrangian L, the Cartan equations (4.8.17a) - (4.8.17b)
are equivalent to the Lagrange equations (4.8.7), and lead to the unique dynamic
equation
q
3
t
t = (n-
1
Y
3
[-diC + d
t
n
l
+ qtd
k
ir
l
} (4.8.20)
on the configuration space Q.
By very definition, the Poincare-Cartan form Hi (4.8.5) defines the fibred mor-
phism
HL-.J'Q -^T'Q, (4.8.21)
{p
t
,p)oH
L
= (TT,, C-n
t
qi),
from the affine jet bundle J
X
Q >Q to the cotangent bundle T'Q Q of Q,
which plays the role of the homogeneous Legendre bundle (2.9.3), equipped with the
coordinates (t,q',p,pi). The morphism (4.8.21) is termed the Legendre morphism
4.8. LAGRANGIAN SYSTEMS 193
associated with Hi. There is the following relation between the Legendre map L
(4.8.9) and the Legendre morphism H
L
(4.8.21):
L = C o H
L
, (4.8.22)
where C is the canonical projection T'Q >V'Q (1.1.7).
Furthermore, it is readily observed that the PoincareCartan form Hi is the
pull-back of the canonical Liouville form
E = pdt + pidq
x
on the cotangent bundle T'Q by the associated Legendre morphism (4.8.21). Ac-
cordingly, the presymplectic form
dH
L
=d7r, A dq* - d^tf -C)Adt (4.8.23)
on the jet manifold J
1
Q is the pull-back of the canonical symplectic form
dz. =dp Adt + dpi A dq
l
on the cotangent bundle T'Q of the configuration space Q.
Remark 4. 8. 8. The presymplectic formdH
L
(4.8.23), together with the 1-form dt,
define a copresymplectic structure on the configuration space J
l
Q [37, 86]. Let us
take the exterior product
(dH
L
)
m
Adt = det(7Ty)( ]T dql A dq*) A dt.
I
(4.8.24)
If it is nowhere vanishing, the pair (dHi,dt) is called a cosymplectic structure on
the (2m +l)-dimensional manifold J
1
Q. A glance at the expression (4.8.24) shows
that the pair (dHi, dt) is cosymplectic if and only if the Lagrangian L is regular. A
cosymplectic structure (dHi,dt) on J
l
Q defines the isomorphism
TJ
l
Q 9 H v\dH
L
+ {v\dt)dt e T'J
l
Q.
In particular, let T be a subbundle of the tangent bundle TJ
l
Q J
1
Q. One can
define the orthocomplement
T
1
= {v TJ
l
Q : v\dH
L
+ (v\dt)dt Ann (T)}
of T in TJ
l
Q with respect to (dH
t
, dt).
194 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
Note that the classification of Lagrangians on the product R x M by the proper-
ties of the intersection KerdHi n Kei dt has been suggested [86]. This classification
remains also true for Lagrangians on an arbitrary fibre bundle Q R, though not all
its constructions (e.g., the tangent-valued forms S and ~S (see [86])) are maintained
under time-dependent transformations.
Given a Lagrangian L, let the image Z
L
of the configuration space J
l
Q by the
Legendre morphism Hi (4.8.21) be an imbedded subbundle
k Z
L
*-* T'Q
of the cotangent bundle T'Q. It is provided with the pull-back De Dondei form
=.
L
= i'
L
E.
By analogy with the Cartan equations (4.8.18), the corresponding Hamilton-De
Donder equations for sections r of the fibre bundle Zi R are written as the
condition
T'{u\dE
L
) = 0 (4.8.25)
where u is an arbitrary vertical vector field on ZL > R. To obtain these equations
in an explicit form, one should substitute the solutions
q
,
t{.t,q
i
,p
j
), {t,q>,pj,p)
of the equations
Pi = *,{t,q
j
,q{),
p = C{t,q>,4) - ni(t,q>,qi)q\ (4.8.26)
in the Cartan equations (4.8.18).
If a Lagrangian L is regular, the equations (4.8.26) have a unique solution. Then
the Hamilton-De Donder equations take the coordinate form
d
t
r = -d
l
r,
d
t
T
{
= dir,
r = C(t,q\g'
t
{t,g
]
,p
J
))-p
l
q'
t
(t,q
i
,p
1
),
(4.8.27)
and are equivalent to the Cartan equations. If a Lagrangian L is degenerate, the
equations (4.8.26) may admit different solutions, or no solutions at all. More can
be said in the following case.
4.9. NEWTONIAN SYSTEMS 195
PROPOSI TI ON 4.8.2. Let the Legendre morphismH
L
: J
l
Q Z
L
be a submersion.
Then a section c of the jet bundle J
l
Q > K is a solution of the Cartan equations
(4.8.18) if and only if HL c is a solution of the Hamilton-De Donder equations
(4.8.25) [63]. a
Remark 4.8.9. In the case of a regular Lagrangian L, the Hamilton-De Donder
equations (4.8.27) can be seen as the equations of motion in variant (B) of the
Dirac system on the symplectic manifold T'Q in the case of: (i) a zero Hamiltonian,
(ii) the primary constraint space N = Z
L
(4.8.26), and (iii) with the additional
condition that a solution v of the equation (3.6.1) has the component v = 1. The
latter condition guarantees that the motion parameter is t.
In conclusion, let us touch briefly on the case of conservative Lagrangians. Let
us suppose that, given a trivialization
ip:Q =R x M
in the coordinates (,<f), a Lagrangian L is independent of t. With respect to these
coordinates, the configuration space J
l
Q admits the presymplectic form
u/L = ctn, A dq
x
(cf. (3.3.5)), and the equation (4.8.14) for a Lagrangian connection fx, may be
written in the form
(,L\UL = -d{-Kiq\ - C).
Thus, a conservative Lagrangian system reduces to the presymplectic Hamiltonian
system whose Hamiltonian is the energy function (^gj C) (see Examples 3.3.2 and
3.4.2). In particular, if a Lagrangian is degenerate, one can apply the procedure from
Section 3.6 to investigate it [138]. We refer the reader to [31, 37, 111] for extension
of this procedure to time-dependent degenerate Lagrangian systems on the product
R x M. We will investigate degenerate Lagrangian systems in the framework of
Hamiltonian formalism (see Section 5.5).
4.9 Newtonian systems
Let L be a Lagrangian on a velocity phase space J
1
Q and L the Legendre map
(4.8.9). Due to the vertical splitting (1.1.6) of VV'Q, the vertical tangent map VL
196 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
to L reads
VL : V
Q
J
X
Q - . V'Q x V'Q.
Q
It yields the linear fibred morphism
b
d
5=
f
(I d
J 1Q
,pr
2
oVL) : V
Q
J
l
Q - V ^' Q,
b : d[ i-> ffyd^,
(4.9.1)
where {dg?} are bases for the fibres of the vertical tangent bundle VA J
l
Q >J 'Q.
The morphism (4.9.1) defines the mapping
J
l
Q -* VAJ'Q 9 VAJ
l
Q
W
J'Q *
and, due to the splitting (4.1.9), also the mapping
m:J
l
Q >V QV Q,
"*ij = P,j om = 7Ty,
where (,(?', p
tJ
) are holonomic coordinates on V'QV'Q. Thus, n^ = m
tJ
are
2 ___
components of the W'Q-val ued field m on the velocity phase space J Q. It is
called the mass tensor.
Let a Lagrangian L be regular. Then the mass tensor is non-degenerate, and
defines a fibre metric, called mass metric, in the vertical tangent bundle VQJ^Q
J
l
Q. Let us recall that, if a Lagrangian L is regular, there exists a unique Lagrangian
connection & for L which is holonomic in accordance with the equation (4.8.15a),
and obeys the equation
mutt* = -dtn - djTTi^t + d,. (4.9.2)
This holonomic connection defines the dynamic equation (4.8.20). At the same time,
the equation (4.9.2) leads to the commutative diagram
VQJ'Q -*- VAJ'Q
D
KL
\/e
L
J
2
Q
S
L
= \>OD
U
,
> = m
xk
{q'!
t
-(,
k
L
),
(4.9.3)
4.9. NEWTONIAN SYSTEMS
197
where D^
L
is the covariant differential (4.1.17) relative to the connection fa. Fur-
thermore, the derivation of (4.9.2) with respect to q\ results in the relation
fa\dmij +771^7* +mj
k
nff =0,
(4.9.4)
where
li 2 SL
are coefficients of the symmetric dynamic connection 7^(4.3.10) corresponding to
the dynamic equation fa.
Thus, each regular Lagrangian L defines the dynamic equation fa, related to the
Euler-Lagrange operator
L
by means of the equality (4.9.3), and the non-degenerate
mass tensor m^, related to the dynamic equation fa by means of the relation (4.9.4).
This is a Newtonian system in accordance with the following definition.
DEFI NI TI ON 4.9.1. Let Q >R be a fibre bundle together with
a (non-degenerate) fibre metric m in the fibre bundle VQJ
1
Q >J
Y
Q:
m:J
l
Q-+VQVQ,
Q
m = -mijdq' V dq
3
,
satisfying the symmetry condition
dlrriij =d'rriik,
(4.9.5)
and a holonomic connection f (4.1.25) on the jet bundle J
l
Q K, related to
the fibre metric m by the compatibility condition (4.9.4).
The triple (Q, m, ?) is called a Newtonian system. This is not the terminology of
[31].
Note that the compatibility condition (4.9.4) can also be introduced in an in-
trinsic way as
V
?
m =0,
198 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
where by V is the covariant derivative with respect to the connection 7^on the
vertical cotangent bundle VqJ
l
Q * J
1
Q, which is dual of the connection \ (4.3.23)
on the vertical tangent bundle VQ J
l
Q > J
l
Q.
Definition 4.9.1 generalizes the second Newton law of point mechanics. Indeed,
the dynamic equation for a Newtonian system is equivalent to the equation
m,
k
(qu - *) =0.
(4.9.6)
There are two main reasons for considering Newtonian systems.
From the physical viewpoint, with a mass tensor, we can introduce the notion
of an external force. Note that, in the universal unit system, the mass tensor m
is dimensional. For instance, the dimension of a mass tensor of a point mass with
respect to Cartesian coordinates q' is [length]
-1
, while that with respect to the angle
coordinates is [length].
DEFI NI TI ON 4.9.2. An external force is defined as a section of the vertical cotangent
bundle VQJ
1
Q > J
l
Q. Let us also bear in mind the isomorphism (4.1.9).
Note that there are no canonical isomorphisms between the vertical cotangent
bundle VQ J
l
Q and the vertical tangent bundle VQ J 'Q of J
l
Q. Therefore, one should
distinguish forces and accelerations which are related by means of a mass metric (see
also Remark 4.9.2 below).
Let (Q, m, ) be a Newtonian system and / an external force. Then
o=r+(m-n
(4.9.7)
is a dynamic equation, but the triple (Q, m, /) is not a Newtonian system in general.
As follows from direct computation, if and only if an external force possesses the
property
af/i +fl}/i =o, (4.9.8)
then / (4.9.7) fulfills the relation (4.9.4), and (Q, m, /) is also a Newtonian system.
Example 4. 9. 1. For instance, the Lorentz force
}i = eF^q?,
9? =1,
where
(4.9.9)
F\ = dxA^ - d^Ax
(4.9.10)
4.9. NEWTONIAN SYSTEMS 199
is the electromagnetic strength, obeys the condition (4.9.8). Note that the Lorentz
force (4.9.9), just as other forces, can be expressed in the relative velocities qr with
respect to an arbitrary reference frame T:
*-($*-*+4
where q are the coordinates adapted to the reference frame T, and J P is the electro-
magnetic strength, written with respect to these coordinates.
Remark 4. 9. 2. The contribution of an external force / to a dynamic equation
qit-S
l
= (m-
l
)
ik
fk
of a Newtonian system obviously depends on a mass tensor. It should be empha-
sized that, besides external forces, we have a universai force which is a holonomic
connection
C = ^ A < A ,
A
- Q? = 1,
associated with a symmetric linear connection K (4.4.3) on the tangent bundle
TQ >Q. From the physical viewpoint, this is a non-relativistic gravitational force,
including an inertial force, whose contribution to a dynamic equation is independent
of a mass tensor.
From the mathematical viewpoint, the equation (4.9.6) is the kernel of an Euler-
Lagrange-type operator (see (4.9.25) below). By an appropriate choice of a mass
tensor, one may hope to bring it into Lagrange equations. We have seen that a
non-degenerate Lagrangian system is necessarily a Newtonian one, while a converse
statement is not generally true.
Exampl e 4. 9. 3. Let us consider a non-degenerate quadratic Lagrangian
= ^( ^) *rf + w) ? ; + <K<n>
(4.9.11)
where the mass tensor m
tJ
is a Riemannian metric in the vertical tangent bundle
VQ Q (see the isomorphism (4.1.8)). Then the Lagrangian L (4.9.11) can be
written as
= -^Sa^gfgf, <?? = !,
(4.9.12)
200 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
where g is the metric
Poo = -2<t>,
9oi ~k<>
9i
3
= - " >)
(4.9.13)
on the tangent bundle TQ. The corresponding Lagrange equations take the form
q\
t
= -{m-
l
)
,k
UMtll 9? =1,
(4.9.14)
where
{\nv} = Adxg^u +d^g^x - d^gx,,)
are the Christoffel symbols of the metric (4.9.13). Let us assume that this metric is
non-degenerate. By virtue of Corollary 4.4.4, the dynamic equation (4.9.14) gives
rise to the geodesic equation (4.4.9) on the tangent bundle TQ, which reads
9
fl
=0,
9 =1,
* = {AVHV - <AAO,}?V.
Let us now bring the Lagrangian function (4.9.11) into the form
= ^
y
(O(?;-r)te-r' )
+
0'(
9
"), (4.9.15)
where T is a Lagrangian frame connection onQ- >R which takes its values into the
kernel of the Legendre map L (see Section 5.6). This connection defines an atlas
of local constant trivializations of the fibre bundle Q R and the corresponding
coordinates (t,5*) on Q such that the transition functions q
1
q" are independent
oft, and P =0 with respect to (,<f). In these coordinates, the Lagrangian (4.9.15)
reads
=^m+wof ))
(4.9.16)
Let us assume that <f>' is a nowhere vanishing function on Q. Then the Lagrange
equations (4.9.14) take the form
Qu = {x\}t?
t
, 3? = l,
where {>'} are the Christoffel symbols of the metric (4.9.13), whose components
with respect to the coordinates ((,?") read
goo = -20',
9oi =0, 9x] = -?Ry.
(4.9.17)
4.9. NEWTONIAN SYSTEMS 201
Then the spatial part of the corresponding geodesic equation
5 = 0,
9 =1 ,
Q ={x /}? q
(4.9.18)
on the tangent bundle TQ is precisely the spatial part of the geodesic equation with
respect to the Levi-Civita connection for the metric (4.9.17) on TQ.
This example shows that a mass tensor may be treated sometimes as a field
variable.
A Newtonian system (Q, rh, ) is said to be standard, if rh is the pull-back on
VQJ
1
Q of a fibre metric in the vertical tangent bundle VQ Q in accordance with
the isomorphisms (4.1.8) and (4.1.9), i.e., the mass tensor rh is independent of the
velocity coordinates q\.
It is readily observed that any fibre metric rh in VQ Q can be seen as a mass
metric of a standard Newtonian system, given by the Lagrangian
c = \nrnWM - n(4 - n,
(4.9.19)
where T is a reference frame. If rh is a Riemannian metric, one can think of the
Lagrangian function (4.9.19) as being a kinetic energy with respect to the reference
frame T.
Exampl e 4.9.4. Let us consider a system of n distinguishable particles with masses
(mi ,..., m
n
) in a 3-dimensional Euclidean space R
3
. Their positions (ri ,..., r)
span the configuration space R
3n
. The total kinetic energy is
Z
A=l
that corresponds to the mass tensor
Tn
A
Btj = S/igSiftrtAi
A,B = l ,...n, i,j = 1,2,3,
on the configuration space R
3n
. To separate the translation degrees of freedom, one
performs a linear coordinate transformation
(ri ,...,r
n
) (- (7i ,...,p_i ,R),
202 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
where R is the centre of mass, while the n - 1 vectors (p
u
.. ., p-i) are mass-
weighted Jacobi vectors (see their definition below) [119, 120). The J acobi vectors
PA are chosen so that the kinetic energy about the centre of mass has the form
i n - l
Z
A=l
(4.9.20)
that corresponds to the Euclidean mass tensor
TKABi] = ^AB^ij,
A,B = l , . . . n- l , i,j = 1,2,3,
on the translation-reduced configuration space R
3n
~
3
. The usual procedure for defin-
ing Jacobi vectors involves organizing the particles into a hierarchy of clusters, in
which every cluster consists of one or more particles, and where each J acobi vector
joins the centres of mass of two clusters, thereby creating a larger cluster. A J acobi
vector, weighted by the square root of the reduced mass of the two clusters it joins,
is the above-mentioned mass-weighted J acobi vector. For example, in the four-body
problem, one can use the following clustering of the particles:
Pi = y/fH(T2-n),
P
2
= \/^2(
r
4 - r
3
),
- . /m
3
r
3
+m
4
r
4
mi r
1
+ m
2
r
2
\
P
3
= V^3 I I ,
\ m
3
+ m
4
mi + m
2
/
Hi mi m
2
'
_L J_ J_
p.2 m
3
m
4
'
J_ 1 1
u.
3
m\ + m.2 m
3
+ m
4
Different clusterings lead to different collections of J acobi vectors, which are related
by linear transformations. Since these transformations maintain the Euclidean form
(4.9.20) of the kinetic energy, they are elements of the group 0(n 1), called the
"democracy group".
Now let us turn to the conditions for a Newtonian system to be a Lagrangian one.
This is the well-known inverse problem formulated for time-dependent mechamcs.
We will investigate it as the particular inverse problem for dynamic systems on fibre
bundles [57].
The equation (4.9.6) is the kernel of the second order differential Euler-Lagrange
type operator
:J
2
Q^ V'Q,
= mik(Z
k
- q
k
tt
)dq
x
.
(4.9.21)
4.9. NEWTONIAN SYSTEMS 203
One can write such an operator as a 2-form
= e
i
e
i
A dt, (4.9.22)
where 6
l
are contact forms (1.3.6). Obviously, the class of Euler-Lagrange type
operators includes the Euler-Lagrange operators (4.8.6). Thus, the inverse problem
reduces to the conditions for the Euler-Lagrange type operators to be the Euler-
Lagrange ones.
Given a fibre bundle Q R, we have the so called variationaJ sequence
0^R ^O( Q) ^D'
1
( J
1
Q) ^Q
ll
(J
2
Q)
6
-^0
%l
(J
3
Q) ^, (4.9.23)
where O
0,l
(J
1
Q) is the space of horizontal densities Cdt on J
l
Q, while D
1,l
(J
2
Q)
is the space of 2-forms id8' A dt on J
2
Q, and so on [46, 57, 179]. In particular,
the elements of D
0,1
(J
l
Q) are the Lagrangians Cdt, while those of D
l,1
(J
2
Q) are
the Euler-Lagrange-type operators written as 2-forms (4.9.22). The key point lies
in the fact that the variationaJ sequence (4.9.23) is a complex. It implies that
S o d
H
=0,
<5
2
o 8 = 0,
(4.9.24)
(4.9.25)
where
d
H
(f) = dtfdt,
feO{J
l
Q),
is the horizontal exterior differential,
6(Cdt) = (di - dtdDLP A dt (4.9.26)
is the Euler-Lagrange map (or the variationai operator), and
^( ^ A dt) = [(2d; - djdj +dldf^P A 6' +
{d)8i + dlSj - 2dtdfi)6\ A 9
j
+ {df
{
- df^)^ A 6*} A dt =0
is the Helmholtz-Sonin map.
A glance at the variationai operator (4.9.26) shows that the image Im<5 of the
variationai operator (4.9.26) consists of the Euler-Lagrange operators, and Im<5 C
Ker<52 in accordance with the equality (4.9.25). It follows that
6
2
() =0
(4.9.27)
204 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
is the necessary condition for an Euler-Lagrange type operator to be an Euler-
Lagrange one
= 6(L)
(4.9.28)
for some Lagrangian L. The condition (4.9.27) takes the coordinate form
d
3
, - di
}
+ fkiBtSj - Sfa) = 0,
d)i + d\j - 2dtdf
x
= 0,
dfi - dfi = 0.
(4.9.29a)
(4.9.29b)
(4.9.29c)
The obstruction preventing the condition (4.9.27) from implying the equality (4.9.28)
is topological [46, 179). If
Q =R
m+1
->R
(in particular, locally), the variational sequence (4.9.23) is exact, i.e., Im6 = Ker6
2
.
and the equality (4.9.27) is also a sufficient condition for an Euler-Lagrange type
operator to be an Euler-Lagrange one.
Remark 4.9.5. A glance at the equality (4.9.24) shows that the Lagrangian L in
the equality (4.9.28) is defined modulo the variationally trivial Lagrangians
L = ckfdt (4.9.30)
where / is a function on Q.
Applying the condition (4.9.27) to the operator (4.9.21), we restate the well-
known Helmholtz conditions (see [34, 78, 81, 143, 156] and references therein). It
is readily observed, that the condition (4.9.29c) is satisfied since the mass tensor
is symmetric. The condition (4.9.29b) holds due to the equality (4.9.4) and the
property (4.9.5). Thus, it suffices to verify the condition (4.9.29a) for a Newtonian
system to be a Lagrangian one.
Example 4.9.6. Let be a free motion equation which takes the form (4.6.9)
with respect to a reference frame (i.g
1
), and let m be a mass tensor which depends
only on the velocity coordinates q\. Such a mass tensor may exist in accordance
with affine coordinate transformations (4.6.10) which maintain the equation (4.6.9).
Then and m make up a Newtonian system. This system is a Lagrangian one if
4.9. NEWTONIAN SYSTEMS 205
m is constant with respect to the above-mentioned reference frame (t,if). Relative
to arbitrary coordinates on a configuration space Q, the corresponding Lagrangian
takes the form (4.9.19), where T is the connection associated with the reference
frame {t,q
i
).
Exampl e 4.9.7. Let us consider the 1-dimensional motion of a point mass mo
subject to friction. It is described by the equation
77ioj
H
= -kq
t
, fc>0, (4.9.31)
on the configuration space R
2
K, coordinated by [t, q). This mechanical system
is characterized by the mass function m = mo and the holonomic connection
k
f =8
t
+q
t
d
q
q
t
fr
m
(4.9.32)
but it is neither a Newtonian nor a Lagrangian system. The conditions (4.9.29a)
and (4.9.29c) are satisfied for an arbitrary mass function m(t,q,q
t
), whereas the
conditions (4.9.4) and (4.9.29b) take the form
kqtd^m km + d
t
m + q
t
d
q
m = 0. (4.9.33)
The mass function m = const, fails to satisfy this relation. Nevertheless, the
equation (4.9.33) has a solution
' Jfc
m = mo exp t .
[mo
(4.9.34)
The mechanical system characterized by the mass function (4.9.34) and the holo-
nomic connection (4.9.32) is both a Newtonian and a Lagrangian system with the
Havas Lagrangian
, 1 [
k
1 2
C = -mo exp t q,
2 [m
0
J
(4.9.35)
[151]. The corresponding Lagrange equations are equivalent to the equation of
motion (4.9.31).
Exampl e 4.9.8. [143]. The dynamic equation
Ztt - Vt = 0, Vtt - y = 0
on the configuration space R
3
K is not equivalent to any system of Lagrange
equations.
206 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
4.10 Holonomic constraints
Let (Q, m, ) be a Newtonian system, where m is a Riemannian metric in the vertical
tangent bundle VQJ
1
Q J
l
Q. This Section is devoted to the restriction of (<5, m, )
to a closed imbedded fibred submanifold i
N
: N - Q of the configuration bundle
Q R, which is treated as a holonomic constraint. We refer the reader to [24]
and references therein for the geometric description of holonomic constraints in
conservative mechanics.
With a holonomic constraint N C Q, we have the following imbedding diagrams
VN <-. VQ
I I
N -+ Q .
\ /
R
y
2
/v -> J
2
<2
I I
I I
N ^-> Q
\ /
R
Furthermore, there exists an open tubular neighbourhood U of the submanifold N,
which is a fibred manifold over TV [109]. Therefore, one can provide the configuration
space Q with the atlas of bundle coordinates (t, a
T
, q
l
) such that
<?' \N= 0, ql |J I J V= o, 9M \J*N= 0.
We will continue to use the notation q
x
for the whole coordinate collection (t, <r
r
, <j').
Let n be the induced Riemannian fibre metric in the vertical tangent bundle
VNJ
1
N > J*N. Its components are
nrs = m
TS
\ji
N
.
Since
d ^ , = d^rrir, \ji
N
,
the Riemannian metric n satisfies the symmetry condition (4.9.5).
Using the Riemannian metric m in VQJ
1
Q >J 'Q, we can define the orthocom-
plement V - J
l
N of the subbundle V
N
J
l
N C V
Q
J
1
Q \JI
N
and the corresponding
splitting
VQJ'Q \JI
N
= V
N
J'N V (4.10.1)
4.10. HOLONOMIC CONSTRAINTS 207
together with the canonical projections
Prj : V
Q
J
l
Q \
fiN
-> VNJ'N,
P^i(d
t
r)=d
t
r
, pt
x
(flf) = n"m
si
dl (4.10.2)
and
pr
2
: VQJ'Q |
J I A f
- V,
pr
2
(5*) =0, pr
2
(9j) =
d\ - n
T
'm
3l
d'
r
= tf
t
.
Since the restriction J
2
Q \j\
N
> J
Y
N is an affine bundle modelled over the vector
bundle VQJ
1
Q \JI
N
, the splitting (4.10.1) defines the corresponding decomposition
J
2
Q UAT= J
2
N 0 V.
J*N
Then the dynamic equation : J
l
Q - J
2
Q of our Newtonian system splits in
the following way:
f |j 'w=(IN + r, (4.10.3)
where
&v J
l
N ~* J
2
N,
C
N
= (t;
T
+n
r
'm
si
e)\ji
N
, (4.10.4)
and
r = Z% : J
l
N V. (4.10.5)
It follows that the dynamic equation f on the configuration space Q induces the
dynamic equation N (4.10.3) on the holonomic constraint N. In an equivalent way,
the induced dynamic equation ^is characterized by the relation
KTSCN = ("V.' +"hi?) Ij'AT
(4.10.6)
PROPOSI TI ON 4.10.1. Let (<2,m,) be a Newtonian system on a configuration space
Q, and N C Q a holonomic constraint. Then (Af, n,jv) is a Newtonian system.
Proof. The proof of the compatibility condition (4.9.4) for
N
and n follows from
direct computation. QED
208 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
Let 7
?
be the dynamic connection (4.3.10) associated with a dynamic equation
of the Newtonian system in question. It can be represented by the tangent valued
form (1.4.4) which reads
7{
: J
l
Q - T'J'Q V
Q
J
l
Q,
7 =(dq\ - 7i V) 9, +(d<- 7XV) &
Let us consider the map
7 : J*N -. J 'Q IJIJV ^ T ' J ' Q VQJ
1
*? |
7
,
W
>
rv'jv vw
1
^,
where the dual morphism (1.1.3)
(./ %): T V ' Q |j .
w
-> TV'TV
and the morphism pr, (4.10.2) are utilized. We obtain
% = {do
r
t
-%dt-r
3
do)d
T
,
(4.10.7)
lo =(7o +
r
'"i5.7o) \j*N, 7p =(7p +n
TS
m
st
j
l
p
) \ji
N
It follows that the map 7^is a connection on the affine jet bundle J N N.
PROPOSI TI ON 4.10.2. The dynamic connection 7^(4.10.7), induced on the jet
bundle J
l
N N of the holonomic constraint N by the dynamic connection 7^,
defines a dynamic equation on N which is precisely
N
(4.10.4), induced on N by
the dynamic equation {,
Proof. The proof is straightforward. QED
It is readily observed that the dynamic connection 75 differs in a torsion from
the symmetric dynamic connection -y
N
, associated with the dynamic equation #.
We have the relation
% = -\TZ + INI
r, = \Ti+y
N
i
77 =-d^n^m^e ~ n
rh
dlm
ht
C- (4.10.8)
Let now a Newtonian system (Q, m,) be a Lagrangian system for a regular
Lagrangian L on the velocity phase space J*Q. The pull-back Ls = i*
N
L of L by
4.10. HOLONOMIC CONSTRAINTS 209
the imbedding J
l
i
N
is a Lagrangian on the holonomic constraint N. We come to
the following assertion.
PROPOSI TI ON 4.10.3. The dynamic equation
LN
, associated with the induced
Lagrangian L^ on N, coincides with the dynamic equation /,#(4.10.4) induced on
N by the dynamic equation
L
, associated with the Lagrangian L.
Proof. We obtain the relation
(mr.a +mrt&) \jx
N
= -d
t
d
l
r
L
N
- a
t
d,d
r
L
N
+ d'
r
L
N
= n
r3
l
(see (4.9.2)), where
n
TS
= m
rs
\ji
N
= dffiLrf.
This is precisely the relation (4.10.6). QED
It is readily observed that the Lagrange equations for a Lagrangian L^ on the
holonomic constraint N are locally equivalent to those of the local Lagrangian
L + Xiy' on Q, where X
{
are the Lagrange multipliers. Consequently, the dynamic
equation L
M
= LH describes a motion on an ideal holonomic constraint N. This
fact also remains true for an arbitrary Newtonian system. Indeed, the splitting
(4.10.3) shows that one can think of the dynamic equation #as being the dynamic
equation at the points of N plus the additional acceleration r (4.10.5), treated as
a constraint reaction acceleration. Then, bearing in mind the isomorphism (4.1.8),
it is readily observed that this acceleration is orthogonal to the holonomic constraint
N with respect to the mass metric m, i.e.,
m(u, s) = m^v!
1
^ m/
lr
ii''n
rs
m
SJ
'
7
=0
for any local vertical tangent vector field u
h
dh on N K.
Thus, we generalize the D'Alembert principle (the principle of virtual displace-
ments) for an arbitrary Newtonian system. In accordance with this principle, a
constraint reaction acceleration of a Newtonian system on an ideal holonomic con-
straint is orthogonal to the constraint manifold with respect to the mass metric of
this system.
One can say more when (Q,m,) is a standard Newtonian system, i.e., m is a
Riemannian metric in the vertical tangent bundle VQ >Q. This metric induces
210 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
the Riemannian metric n on the vertical tangent bundle VN N. I ts components
are
TV, = m
r s
\
N
.
We have the splitting
TQ\
N
=TNV
N
with the canonical projections
pr
t
(a
r
) = a
r
,
Wi(9i) = n
r
'm
si
d
r
,
pr
2
(d
r
) =0, pr
2
(9i) =d\ - rfm^dr = d
u
and the corresponding splitting of the first order jet manifold
J
X
Q\
N
= J
l
NV.
N
In particular, any reference frame T on a configuration space Q defines a reference
frame r \ on the holonomic constraint TV in accordance with the decomposition
r \N= r
N
+ a
v
,
nr
s
r% = ( m
r s
r
S
+ TTViP) I AT . (4.10.9)
Moreover, the expression (4.10.8) shows that, in the case of a standard Newtonian
system, the dynamic connection 7
4
(4.10.7) is precisely the dynamic connection y^
N
,
associated with the dynamic equation fa on the holonomic constraint N.
Example 4. 10. 1. Let (Q,m,) be a Newtonian system for a free motion equation
. Then, fa fails to be a free motion equation on a holonomic constraint N in
general. In particular, if L is a Lagrangian (4.9.19), its restriction to a holonomic.
constraint N reads
LN = 2
n
T-(l ~
r
N)(o! ~ T
s
w
) + ^n{o
v
,o
v
),
where I \ is the induced frame (4.10.9).
4.11. NON-HOLONOMIC CONSTRAINTS 211
4.11 Non-holonomic constraints
Here we are concerned with the geometric theory of non-holonomic constraints in
time-dependent mechanics. We refer the reader to [25, 30, 112, 113, 124, 183, 184]
and references therein for the geometric description of non-holonomic constraints
in conservative mechanics. The key problem is that the method of the Lagrange
multipliers is not appropriate to non-holonomic constraints in general (see [7, 25,
112, 114]).
Let the jet manifold J
l
Q be a velocity phase space of time-dependent mechanics
on a configuration bundle Q R. The most general non-holonomic constraints
considered in the literature are given by codistributions S or, accordingly, by distri-
butions Ann(S) on the jet manifold J
l
Q [55, 132]. In conservative mechanics on a
configuration space M, this is the case of a codistribution on TM [183]. Submani-
folds of the jet manifold J
l
Q [88, 106] and distributions on a configuration space Q
[114] can also be seen as non-holonomic constraints.
In connection with non-holonomic constraints in time-dependent mechanics, one
usually studies the following problem. Given a configuration space Q, let be a
dynamic equation on Q, and S a codistribution on J
l
Q whose annihilator Ann (S)
is treated as a non-holonomic constraint. Similarly to the case of holonomic con-
straints, the goal is a decomposition
= I + r,
(4.11.1)
where is a dynamic equation obeying the condition
|cA nn( S) .
Then one can think of the dynamic equation as describing a mechanical system
subject to the non-holonomic constraint S, while ( r) is said to be the constraint
reaction acceleration.
Let us assume that the codistribution S has dimension n and is locally spanned
by the 1-forms
s" =s
a
0
dt + sdq' + sdq\
on the jet manifold J
l
Q. Then a dynamic equation is compatible with the non-
holonomic constraint S if
S
a
()=Z]s
a
= S
a
o
+ Sql + Se = 0.
212 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
This equation is algebraically solvable for n components of if and only if the nxm
matrix s(g\g|) has everywhere maximal rank n < m. Therefore, we will restrict
our consideration to the non-holonomic constraints, called admissible, such that
dimS =dim?(S),
where v* is the vertical endomorphism (4.1.14)
Example 4. 11. 1. Note that holonomic constraints can also be seen as the non-
holonomic ones which, however, are not admissible.
If a non-holonomic constraint is admissible, there exists a local m x n matrix
sJ,(g
A
,<?!) such that
4*J = <t
(4.11.2)
Then, the local decomposition (4.11.1) of a dynamic equation can be written in
the form
g =g +JJafjQ.
(4.11.3)
The global decomposition (4.11.1) exists by virtue of the following lemma.
LEMMA 4.11.1. [55]. The intersection
F= J
2
QnAnn(S)
is an affine bundle over J 'Q, modelled over the vector bundle
F = V
Q
J
l
Qn Ann (S).
n
Proof. The intersection F consists of the vertical vectors v
x
d\ 6 VQJ
1
Q which fulfill
the conditions
*?(*,*)* = 0.
Since the non-holonomic constraint S is admissible, every fibre of F is of dimension
m-n, i.e., F is a vector bundle, while F is an affine bundle. QED
Since the intersection J
2
Q n Ann (S) >J
1
Q is an affine bundle, it has always a
global section f by virtue of Theorem 1.1.2.
4.11. NON-HOLONOMIC CONSTRAINTS 213
To construct the global decomposition (4.11.1), one should perform a splitting
of the vertical tangent bundle
V
Q
J
l
Q = F V (4.11.4)
in order to obtain the corresponding splitting of the second order jet manifold
J
2
Q = F V (4.11.5)
which is modelled over (4.11.4). Here, V J 'Q should be interpreted as the bundle
of possible constraint reaction accelerations.
If an admissible non-holonomic constraint S is of dimension n =m, a dynamic
equation is decomposed in a unique fashion. If n < m, the decomposition (4.11.1)
is indeterminate. Different variants of this decomposition lead to different con-
straint reaction accelerations which, from the physical viewpoint, characterize dif-
ferent types of non-holonomic constraints.
Let us turn now to some important particular cases of non-holonomic constraints.
Exampl e 4. 11. 2. Let N be a closed imbedded submanifold of the velocity phase
space J
l
Q, defined locally by the equations
/
a
(g\<?l) =o,
a = 1,... ,n < m. (4.11.6)
One can treat AT as a non-holonomic constraint at points N C J 'Q, given by the
codistribution S =Ann(TN) on J
l
Q \s- This codistribution is locally spanned by
the 1-forms
s
a
=df
a
= d
t
f
a
dt +d
]
}
a
dq> + aj/d^.
The non-holonomic constraint ,/V is admissible if and only if the matrix (9|/) is
of maximal rank n. It follows that N is a fibred submanifold of the affine jet
bundle J
i
Q Q. Moreover, it is possible to separate locally some velocities <,
a = 1,..., n, expressed as functions of the remaining coordinates (q
x
, qf , , <?[")
such that the equations (4.11.6) are brought into the form
Qt ~9 {Q ,Qt i . It ) = 0-
(4.11.7)
Then we have
sg = -d
t
g
a
,
? = ~dw
a
,
S? =6f - dig",
(4.11.8)
214 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
where d
l
b
g
a
= 0 for all indices a, b =1,..., n. Now, applying the above procedure to
the codistribution (4.11.8), one can obtain a dynamic equation on a submanifold N,
compatible with the constraint Ann (TN). For instance, one can use the particular
solution s\ = 6
l
a
of the equations (4.11.2), where the matrix s? is given by the
expression (4.11.8) [106]. Then the (local) dynamic equation (4.11.3), compatible
with the constraint (4.11.7), reads

, /a
= e,
={ _
s
*() =d
k
g
a
e + d
t
g
a
+ qfrg*. (4.11.9)

Example 4.11.3. A non-holonomic constraint N is said to be linear if it is an affine
subbundle of the affine jet bundle J
l
Q >Q, which is given by the local equations
r=/oV)+/f(<?
A
)<?; = o,
(4.11.10)
where the matrix /
t
a
is of maximal rank. A linear non-holonomic constraint is always
admissible. Since N is an affine subbundle of J
l
Q >Q, it has a global section V
which is a connection on the configuration bundle Q > R, called the constraint
reference frame. With this connection F, the constraints (4.11.10) take the form
f(q
x
M - n =o.
(4.11.11)
We may say that the linear constraint is stationary with respect to the constraint
reference frame T. Then, one can think of
<t ~ it - r ,
satisfying the equation (4.11.11), as virtual velocities relative to the linear constraint
Example 4. 11. 4. Let a configuration space Q admit a composite fibration Q
E ->R, where
7TQE : Q E
is a fibre bundle, and let (t, a
T
, g
a
) be coordinates on Q, compatible with this fibra-
tion. Given a connection
B =dt (d
t
+ B
a
d
a
) + da
T
(d
T
+ B?d
a
)
(4.11.12)
4.11. NON-HOLONOMIC CONSTRAINTS 215
on the fibre bundle Q E, we obtain the corresponding horizontal splitting (1.4.1)
of the tangent bundle TQ. Restricted to the jet manifold J
X
Q C TQ, this splitting
reads
J
1
Q = B(n'
QE
J
1
Yi)V
E
Q,
dt + c\d
T
+ q?d
a
= [(d
t
+ B
a
d
a
) + a
r
t
{d
T
+ B
a
T
d
a
)] +
[tf - B
a
- o\B?\d
a
,
where 7TQ
E
J
l
E is the pull-back of the affine jet bundle J 'E > E onto Q. It is readily
observed that
N = B(n'
QS
J
l
Z)
is an affine subbundle of the affine jet bundle J
l
Q Q, defined locally by the
equations
qf - o
T
t
B(q
X
) ~ B
a
(q
x
) = 0
(cf. (4.11.7)). Then this subbundle yields a linear non-holonomic constraint, given
by the codistribution S = Ann (TN) [162, 163]. This codistribution is locally
spanned by the 1-forms
s
a
= -{d
t
B
a
+ o\d
t
B)dt - (d
a
B
a
+ o
r
t
d
s
B*)do" -
(d
b
B
a
+ a
r
t
d
b
B
a
T
)dq
b
+ dqf - Bdo\.
(4.11.13)
With the connection (4.11.12), we also have the splitting (1.6.10a) of the vertical
tangent bundle VQ of Q >R and the corresponding splitting of the vertical tangent
bundle VQJ
1
Q (see the canonical isomorphism (4.1.21)). The latter splitting reads
V
Q
J
l
Q = F 0 V ,
a[dl + qfdl = a
r
t
(% + B^
a
) + (q? - B
r
V[)^. (4.11.14)
It is readily observed that F |/v consists of vertical vectors which are the annihilators
of the codistribution (4.11.13). The splitting (4.11.14) yields the corresponding
splitting (4.11.5) of the second order jet manifold J
2
Q. Then, given a dynamic
equation f on J
l
Q, we obtain the decomposition (4.11.1) which reads
f = c, e=r - s(o
216 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
(cf. (4.11.9)).
Now we consider Newtonian systems because they provide the vertical tangent
bundle VQJ
1
Q with a non-degenerate fibre metric m. Let us assume that m is a
Riemannian metric. With this metric, we immediately obtain the splitting (4.11.4),
where V is the orthocomplement of F. Then the decomposition (4.11.3) takes the
form
f = |
i
+ m
at
m"5fo
t
(0.
(4.11.15)
where m,^is the inverse matrix of
m
ab
= s's^mv
[55] By definition, the decomposition (4.11.15) satisfies the generalized D'AIembert
principle. The constraint reaction acceleration
- r' = -m
ab
m
lJ
ss
b
(Z) (4.11.16)
is orthogonal to every element of VQj
i
QnArm (S) with respect to the mass metric m.
Since elements of VQ J 'QflAnn (S) can be treated as the virtual accelerations relative
to the non-holonomic constraint S, the constraint reaction acceleration (4.11.16)
characterizes S as an ideai non-holonomic constraint.
The Gauss principle is also fulfilled as follows. Given a dynamic equation and
the above-mentioned fibre metric m, let us define a positive function G(w) on J^Q
as
G(w) =rh ((ir?(uO) - U/,(J T?(IU)) - w) ,
G(q\ q\, q
l
tt
) = m
i}
(q\ <g)(?(q\ <) - q\
t
)(?(q\ <j
t
fc
) - q}
t
).
We say that
IMI =^GH
is a norm of w 6 J
2
Q.
PROPOSI TI ON 4.11.2. [55]. Among all dynamic equations compatible with a
non-holonomic constraint, the dynamic equation defined by the decomposition
(4.11.15) is that of least norm. D
4.11. NON-HOLONOMIC CONSTRAINTS 217
Pr oof. Let be another dynamic equation which takes its values into F. Then
Then we obtain
Kl l - a- f+- C. - + - 0 = llCll + S(C-C.C-0.
i.e., ||Cn >llltl- QED
Now we will show that, in the case of non-degenerate Lagrangian system and lin-
ear non-holonomic constraints, the decomposition (4.11.15) satisfies the traditional
D'Alembert principle.
Let S be an admissible non-holonomic constraint on the velocity phase space
J
X
Q, and L a regular Lagrangian on J
l
Q with a Riemannian mass metric m
l}
= n
l}
.
Since this is a particular Newtonian system, we obtain the dynamic equation
& = ft - f w' ^ ^ gi + jjq* +&] (4.H.17)
which is compatible with the constraint S, treated as an ideal non-holonomic con-
straint. This is the system of Lagrange equations in the presence of a non-Lagrangian
external force
F, = -rH^Afe)
(see the equation (4.12.9) below). It is a constraint reaction force. Let us consider
the energy-momenturn conservation law in the presence of this force. It reads
LpL - frft = -d
(
Tr,
where LfL is the Lie derivative of the Lagrangian L along a reference frame T and
Tr is the energy function relative to this reference frame (see (4.12.15), (4.12.16),
and (4.12.19) below). In particular, if a non-holonomic constraint is linear and T
is a constraint reference frame (see Example 4.11.3), the constraint reaction force
does not contribute to the energy conservation law. It follows that, in this case,
the standard D'Alembert principle holds, while the equation (4.11.17) describes a
motion in the presence of an ideal non-holonomic constraint in the sense of this
D'Alembert principle.
218 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
The constrained motion equation (4.11.17) on a configuration space Q is neither
a system of Lagrange equations nor a dynamic equation of a Newtonian system.
In Section 5.11, we will show that this can be seen as Hamilton equations in the
framework of the Hamiltonian formalism extended to the configuration space VQ.
4.12 Lagrangian conservation laws
Given a Lagrangian system (J 'Q, L), its integrals of motion can be found if a La-
grangian L possesses symmetries. In time-dependent mechanics, these integrals of
motion should be covariant under time-dependent transformations. For instance,
the canonical energy function
EL = mq
x
t
-
(4.12.1)
fails to be such an integral of motion. There are different approaches in order to
obtain the conservation laws in Lagrangian dynamics (see [1, 29, 51, 52, 57, 160] for
the geometric methods). As in field theory, we will use the first variational formula
of the calculus of variations [57, 160].
Let L be a Lagrangian (4.8.1) on the velocity phase space J
l
Q and
u =u
l
dt + u'd
x
, u
l
= 0,1,
(4.12.2)
a projectable vector field (4.8.2) on the configuration bundle Q R. By virtue of
Theorem 1.2.1, this vector field may be treated as the generator of a 1-parameter
group of local automorphisms, called gauge transformations, of the fibre bundle
Q >R. In particular, if u
l
= 0, the vertical vector field (4.12.2) is the generator
of vertical automorphisms of the fibre bundle Q > R projected over the identity
transformation of the base R. If u
l
= 1, the vector field u (4.12.2) is projected over
the standard vector field d
t
on the base R, which is the generator of the group of
translations of R.
Let us apply the first variational formula (4.8.4) to a Lagrangian L and to a
vector field u (4.12.2). On the shell (4.8.7), the identity (4.8.4) is brought into the
weak identity
u\dC -dtl,
(u'd
t
+u% + d
t
u'dl)C w -(f,(7r,(u'g; - u') - u'C),
(4.12.3)
4.12. LAGRANGIAN CONSERVATION LAWS
219
where, by analogy with field theory,
T = -u\H
L
= K
i
{u
t
q\ - u
{
) - u
l
C
(4.12.4)
is said to be the current along the vector field u. The symbol " ~" stands throughout
for weak equalities fulfilled on-shell.
If the Lie derivative L^L of a Lagrangian L along a vector field u vanishes,
i.e., L is invariant under the 1-parameter group generated by u, we have the weak
conservation law
0 -d,[7T,(u'qJ - u*) - ']. (4.12.5)
It is brought into the differential conservation law
0 ~ T[(
W
> c)(u'3[c' - u' o c ) - u'C o c]
at
on solutions c of the Lagrange equations (4.8.8). A glance at this expression shows
that, in time-dependent mechanics, the conserved current (4.12.4) plays the role of
an integral of motion.
Remark 4. 12. 1. Let L be a Lagrangian on the velocity phase space J 'Q. Every
integral of motion (4.12.4) defines a non-holonomic constraint dT = 0 on J
l
Q such
that any Lagrangian connection fx, for L is a dynamic equation compatible with this
constraint.
Remark 4. 12. 2. The conservation law (4.12.5) is broken if a Lagrangian depends
on variable parameters which do not live in the dynamic shell (4.8.7) (see Section
5.9). In order to obtain a conservation law in the presence of such parameters, let
us consider a bundle product
Qtot = Q x Y
(4.12.6)
of a configuration bundle Q > R with coordinates [t, q') and a fibre bundle Y > R
with coordinates (t, y
A
), whose sections are the variable parameters which take the
background values
y
B
= <P
B
(x),
f = d,4>
B
{x).
A Lagrangian L is defined on the total velocity phase space ./'Qtot-
220 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
Let u be a projectable vector field (4.12.2) on Qtot which projects also onto Y
because the transformations of parameters do not depend on the dynamic variables.
This vector field takes the coordinate form
u = u
l
d
t
+u
A
(t, y
B
)d
A
+u'{t,y
B
, <?)%.
(4.12.7)
Substitution of the vector field (4.12.7) into the first variational formula (4.8.4)
leads to the first variationai formula in the presence of parameters
[u'd, +u
A
d
A
+ u'd, + d
t
u
A
d
A
+ dtu'dflC =(u
A
- ytu
l
)d
A
C +
(4.12.8)
v
A
dt(v.
A
- j /
t
V) + (W - qlu^Si - dt[i(uWt ~ *) - *]
Then the weak identity
[u% + u
A
d
A
+u'd, + dtU
A
d
l
A
+ dtU
l
%]C {u
A
- VS)9A^ +
n
A
dt(u
A
- y
A
u
l
) - MM^Ql ~ "') -
u
']
is fulfilled on the dynamic shell (4.8.7).
If a Lagrangian L is invariant under gauge transformations of the product (4.12.6)
whose generator is the vector field u (4.12.7), we obtain the weaic conservation Jaw
in the presence of parameters
(u
A
- y^u
{
)d
A
C + n
A
d
t
(u
A
- y
A
u
l
) d
t
{K,{u
l
q'
t
- u') - u*].

Remark 4.12.3. The first variational formula (4.8.4) can also be utilized when
a Lagrangian possesses symmetries, but a motion equation is seen as Lagrange
equations plus additional non-Lagrangian external forces and reads
[d
i
-d
t
df)C + F
i
{t,qi,<f
t
) = 0. (4.12.9)
Let us substitute =F, from this equality in the first variational formula (4.8.4)
and assume that the Lie derivative of the Lagrangian L along a vector field u van-
ishes. Then, we have the conservation law
(it
1
- q\)F, g - dtMu'qj - u') - u'].
(4.12.10)

4.12. LAGRANGIAN CONSERVATION LAWS 221
It is easy to see that the weak identity (4.12.3) is linear in the vector field u.
Therefore, one can consider superposition of the weak identities (4.12.3) associated
with different vector fields.
For instance, if u and v! are projectable vector fields (4.12.2), projected onto the
standard vector field d
t
on R, the difference of the corresponding weak identities
(4.12.3) results in the weak identity (4.12.3) associated with the vertical vector field
u v!. Conversely, every vector field u (4.12.2), projected onto d
t
, can be written
as the sum
u = Y + ti (4.12.11)
of some reference frame
r = dt + rd
t
(4.12.12)
and a vertical vector field $ on Q.
It follows that the weak identity (4.12.3) associated with an arbitrary vector
field u (4.12.2) can be represented as the superposition of those associated with a
reference frame V (4.12.12) and some vertical vector field d.
If u = d is a vertical field, the weak identity (4.12.3) reads
(^di + dt^dDC^dtin^).
If the Lie derivative of L along -d vanishes, we obtain from (4.12.5) the weak con-
servation law
0 w Mviti')
(4.12.13)
and the integral of motion
1 = -TTjg.
(4.12.14)
By analogy with field theory, (4.12.13) is called the Noether conservation law for
the Noether current (4.12.14).
Example 4. 12. 4. Let assume that, given a trivialization Q =R x M in coordinates
(t, q'), a Lagrangian L is independent of a coordinate g
1
. Then the Lie derivative
of L along the vertical vector field d = d\ equals zero, and we have the conserved
Noether current (4.12.14) which reduces to the momentum X = TTI.. With respect
to arbitrary coordinates (t,^'), this conserved Noether current takes the form
1 = - ^
222
CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
In particular, the free motion Lagrangian admits m conserved Noether currents.
In the case of a reference frame T (4.12.12), where u' = 1, the weak identity
(4.12.3) reads
{d
t
+ rdi + d
(
raj) -Mniqt - V) - c),
(4.12.15)
where
X
r
=7r,(g
t
' - P) - C
(4.12.16)
is said to be the energy function relative to the reference frame T [51, 57, 161).
With respect to the coordinates adapted to the reference frame T, the weak
identity (4.12.15) takes the form of the familiar energy conservation law
d
t
C ss -dt(niq'
t
- ),
(4.12.17)
and Xr coincides with the canonical energy function Ei (4.12.1). It follows that
the canonical energy function Ei is not a unique existent energy function. Each
reference frame defines an energy function.
Example 4.12.5. Let us consider a free motion on a configuration space Q. It is
described by the Lagrangian
= 2
m
'j9t9t-
m = const., (4.12.18)
written with respect to a reference frame (,<?") such that the free motion dynamic
equation takes the form (4.6.1). Let T be the associated connection. Then the
conserved energy function Tr (4.12.16) relative to this reference frame T is precisely
the kinetic energy of this free motion. Relative to arbitrary coordinates ((, q') on Q,
it takes the form
x
r
=m(q\ -r)-C = ImjWM-n(qi - P).
Now we generalize this example for a motion described by the equation (4.12.9),
where is the free motion Lagrangian (4.12.18) and F is an external force. The Lie
derivative of the Lagrangian (4.12.18) along the reference frame T vanishes, and we
have the weak equality (4.12.10) which reads
q\.F
t
dtlr,
(4.12.19)
4.12. LAGRANGIAN CONSERVATION LAWS 223
where q'
r
is the relative velocity. This is the well-known physical law whose left-hand
side is the power of an external force.
Exampl e 4. 12. 6. Let us consider a 1-dimensional motion of a point mass mo
subject to friction on the configuration space R
2
R, coordinated by (t,q) (see
Example 4.9.7). It is described by the dynamic equation (4.9.31) which is the
system of Lagrange equations for the Lagrangian L (4.9.35). It is readily observed
that the Lie derivative (4.8.3) of this Lagrangian along the vector field
T = d
t
-~qd
q
2 mo
(4.12.20)
vanishes. Hence, we have the conserved energy function (4.12.16) with respect to
the reference frame T (4.12.20). This energy function reads
_ 1 [ k , , k . 1 ., mk? o
2T = ^wi
0
exp ' Qt(Qt + Q) = ^rnq
r
- j g ,
2 [m
0
J mo 2 8mo
where m is the mass function (4.9.34).
Since any vector field u (4.12.2) can be represented as the sum (4.12.11) of a
reference frame T (4.12.12) and a vertical vector field $, each current (4.12.4) along
a vector field u (4.12.2) is the sum of a Noether current (4.12.14) along the vertical
vector field t? and the energy function (4.12.16) relative to the reference frame T
[51, 57, 161]. Conversely, energy functions relative to different reference frames V
and T' differ from each other in the Noether current along the vertical vector field
r-r.
In conclusion, we touch on gauge transformations with a generator u (4.12.2),
which preserve the Euler-Lagrange operator SL, but not necessarily a Lagrangian
L. They are called generalized invariant transformations [57, 105, 174]. We use the
formula
Lj2
u

t
L_L
[54, 57], where
J
2
u = u'dt + u% +dtu'dj + dtdttfd?
is the second order jet prolongation of a vector field (4.12.2).
The following two assertions are immediate corollaries of this formula.
224 CHAPTER 4. LAGRANGIAN TIME-DEPENDENT MECHANICS
COROLLARY 4.12.1. If a Lagrangian L is invariant under a 1-parameter gauge group
whose generator is a vector field u (4.12.2), so is the associated Euler-Lagrange
operator EL. D
COROLLARY 4.12.2. If an Euler-Lagrange operator /,, associated with a Lagran-
gian L, is invariant under a 1-parameter gauge group whose generator is a vector
field u (4.12.2), the Lie derivative h^L is a variationally trivial Lagrangian (4.9.30).
D
Let us consider conservation laws in the case of generalized invariant transfor-
mations. Let L be a Lagrangian and EL the associated Euler-Lagrange operator.
Let u be a vector field (4.12.2) which is the generator of a local 1-parameter group
of gauge transformations such that
L/2
u
z, =0.
(4.12.21)
Then we have the equality
LuL = d,/,
where / is a function on Q. In this case, the weak identity (4.12.3) reads
dtf^dti^iiS-qb+u'C),
and we obtain the weak equality
OwdeMu' - <?; ) +' - / ) . (4.12.22)
Example 4.12.7. Let L be the free motion Lagrangian (4.12.18). The correspond-
ing Euler-Lagrange operator
E
L
= -rriijqltdq'
is invariant under the Galilei transformations with the generator
u
i
= v
i
t + a
i
, v' = const., a
1
= const., (4.12.23)
(see (4.6.10)). At the same time, the Lie derivative of the free motion Lagrangian
(4.12.18) along the vector field (4.12.23) does not vanish, and we have
LjjL =mijV
l
(f
t
= d
t
(m
l ;
vV +c), c = const.,


4.12. LAGRANGIAN CONSERVATION LAWS 225
[148]. Then the weak equality (4.12.22) shows that (q\t q
%
) is a constant of motion.
Remark 4. 12. 8. The invariance condition (4.12.21) is generalized as
if one deals with symmetry transformations of the differential equation Si = 0
[93, 94]. For instance, the Galilei transformations (4.6.10) are the symmetry trans-
formations of the free motion equation.
L^u^L ~ 0


Chapter 5
Hami l toni an ti me-dependent
mechanics
This Chapter is devoted to the Hamiltonian formulation of time-dependent mechan-
ics with respect to an arbitrary reference frame. Let us recall that a configuration
bundle Q > R of time-dependent mechanics is isomorphic to the product R x M,
where M is a typical fibre of Q, but this isomorphism is not canonical in general.
Different trivializations Q = R x M correspond to different reference frames. For
this reason, many constructions of Hamiltonian conservative mechanics on a sym-
plectic manifold Z and its extension to the product I x Z can not be applied to
mechanical systems which are subject to time-dependent transformations, including
canonical transformations and reference-frame transformations.
Let us summarize the main peculiarities of Hamiltonian time-dependent mechan-
ics.
A momentum phase space of time-dependent mechanics is provided with the
canonical degenerate Poisson structure. However, Hamiltonian time-dependent
mechanics does not reduce to a Poisson Hamiltonian system, since a Hamilto-
nian on a momentum phase space of time-dependent mechanics is not a scalar
function under time-dependent transformations (see Remark 5.1.1 below).
As a consequence, the evolution equation of time-dependent mechanics is not
expressed in terms of a Poisson bracket, and integrals of motion cannot be
defined as functions in involution with a Hamiltonian. For the same reason,
the familiar procedure of describing constraint Hamiltonian systems in Section
227
228 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
3.6 cannot be applied to time-dependent mechanics.
Hamiltonian and Lagrangian formulations of time-dependent mechanics are
equivalent in the case of hyperregular Lagrangians. A degenerate Lagrangian
requires a set of associated Hamiltonians in order to exhaust all solutions of
the Lagrange equations.
We follow the notation of the previous Chapter, where Q K is a fibre bundle
(4.0.1), coordinated by (t, q
l
), whose typical fibre M is an m-dimensional manifold,
while its base R is equipped with the Cartesian coordinate t possessing the transition
functions t' = t+ const.
5.1 Canonical Poisson structure
As was mentioned, the momentum phase space of time-dependent mechanics is the
vertical cotangent bundle
7r
n
: V'Q - Q (5.1.1)
of a configuration bundle Q >R. This is the particular Legendre bundle II (2.9.7)
over a fibre bundle Q > X when X =R. The vertical cotangent bundle (5.1.1) is
equipped with the coordinates (,<?', p<= q,).
The momentum phase space V'Q is provided with the canonical Poisson struc-
ture as follows (cf. Example 2.3.8). Let us consider the cotangent bundle T'Q of
the configuration space Q, equipped with the coordinates (t,<?',p,Pi). This is the
homogeneous Legendre bundle (2.9.3) over a fibre bundle Q >X when X = R.
The cotangent bundle T'Q admits the canonical Liouville form
3 =pdt + p,dq' (5.1.2)
and the canonical symplectic form
dE =dp A dt 4- dpi A dq'.
(5.1.3)
The corresponding Poisson bracket on the vector space D(T'Q) of functions on
T'Q reads
{/, 9} = Vfdtg - Vgdtf + d'fd
l9
- d'gdj. (5.1.4)
5.1. CANONICAL POISSON STRUCTURE
229
Let us consider the subspace of D(T'Q) which comprises the pull-backs */ on
T'Q of functions / on the vertical cotangent bundle V'Q by the canonical projection
T'Q >V'Q (1.1.7). It is easily seen that this subspace is closed under the Poisson
bracket (5.1.4). By virtue of Proposition 2.3.1, there exists the canonical Poisson
structure
{!,9}v = d
i
fd
i
g-d
i
gdJ (5.1.5)
on momentum phase space V'Q induced by (5.1.4), i.e.,
C{f,9}v = {Cf,Cg}-
The corresponding Poisson bivector field
w{df,dg) = {f,g}
v
on V'Q >R is vertical with respect to the fibration V'Q M, and reads
w'
J
= 0, w
i}
- 0, W*j = 1.
(5.1.6)
A glance at this expression shows that the holonomic coordinates of V'Q are canon-
ical for the Poisson structure (5.1.5). Since the rank of the bivector field w (5.1.6) is
constant, the Poisson structure (5.1.5) is regular. This Poisson structure is obviously
degenerate.
Given the Poisson bracket (5.1.5), the Hamiltonian vector field Of for a function
/ on the momentum phase space V'Q is defined by the relation
{f,9h
V
P
6 D(V'Q).
It is the vertical vector field
0/ =#fdj -
(5.1.7)
on the fibre bundle V'Q R. Hence, the characteristic distribution of the Poisson
structure (5.1.5) is precisely the vertical tangent bundle VV'Q C TV'Q of the fibre
bundle V'Q R.
In accordance with Theorem 2.3.3, the Poisson structure (5.1.5) defines the sym-
plectic foliation on the momentum phase space V'Q, which coincides with the fibra-
tion V'Q * R. The symplectic forms on the fibres of V'Q ->R are the pull-backs
Q
t
= dpi ^dq'
= 4
f
\dg,
d,fd'
230 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
of the canonical symplectic form on the typical fibre T'M of the fibre bundle V'Q
R with respect to trivialization morphisms [27, 74, 161]. Given such a trivialization
V'Q & R x T'M,
(5.1.8)
the Poisson structure (5.1.5) is isomorphic to the direct product of the zero Poisson
structure on R and the canonical symplectic structure on T'M (see Example 2.3.7).
The Poisson structure (5.1.5) can be introduced in a different way [57, 161].
The polysymplectic form (2.9.9) on the Legendre bundle V'Q -* Q (5.1.1) reads
A =dp, Adq
{
Adtd
t
,
and reduces to the canonical exterior 3-form
n =dp, A dq' A dt.
(5.1.9)
This is the exterior differential of the canonical 2-form
0 =Pidq' A dt
(5.1.10)
on the momentum phase space V'Q. The canonical forms (5.1.9) and (5.1.10) are
maintained under any holonomic coordinate transformation of V'Q.
Given the canonical form (5.1.9), every function / on the momentum phase
space V'Q determines the corresponding Hamiltonian vector field $/ (5.1.7) by the
relation
tf,J ft = -dfAdt, (5.1.11)
while the Poisson bracket (5.1.5) is defined by the condition
{f,g}vdt = 4g\4f\n.
Remark 5.1.1. It should be emphasized that, though the momentum phase space
V'Q of time-dependent mechanics is provided with the canonical Poisson structure,
non-conservative mechanical systems are not reduced to the Poisson Hamiltonian
systems of Section 3.2. Given a trivialization (5.1.8), i.e., a reference frame, one
can write the Hamilton equations (5.2.18a) - (5.2.18b) (see below) as the equations
of the Hamiltonian vector field d-n (5.1.7) for a Hamiltonian Ti with respect to the
Poisson structure (5.1.5). Moreover, by very definition of the canonical form ft
(5.1.9), the reference frame transformations are canonical transformations for the
5.2. HAMILTONIAN CONNECTIONS AND HAMILTONIAN FORMS 231
Poisson structure (5.1.5) (see Section 5.3). However, a Hamiltonian H (see (5.2.10)
below) is not a scalar function under reference frame transformations in general.
Therefore these transformations are not the equivalence transformation of a Poisson
Hamiltonian system.
5.2 Hamiltonian connections and Hamiltonian forms
The dynamics of time-dependent mechanics on a phase phase space V*Q is described
by first order dynamic equation.
DEFI NI TI ON 5.2.1. Let
7
= d
t
+ j% +<* # (5.2.1)
be a connection on the fibre bundle V'Q > M. In accordance with Definition 4.2.1,
it defines the first order dynamic equation on the momentum phase space V'Q K
written as
q\ = V, Pu = 7i
with respect to the adapted coordinates
{t,q\Pi,q'
t
,Pti)
on the jet manifold J
1
V'Q. a
Exampl e 5. 2. 1. Let us consider the product
(J = R x M - t I ,
coordinated by (t, q
l
). We have the canonical isomorphism
V'Q =R x TM,
with the coordinates p<=Qj, and the corresponding isomorphism
JW'Q^RXTT'M,
(5.2.2)
with the coordinates q\ = <?\p =j>i- Every vector field
u = u'd
t
+ Uid'
232 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
on the cotangent bundle T'M defines the first order dynamic equation
7 =d
t
+ u% +u,9*
on V'Q (5.2.2). Its projection
q
x
= u\ p, =u
x
onto T'M is an autonomous first order dynamic equation in conservative mechanics
(see Definition 3.1.1).
Let J
2
Q be the second order jet manifold of a configuration space Q K
provided with the adapted coordinates
(*,?', *.)
Recall that it coincides with the sesquiholonomic jet manifold T*Q.
Following the general scheme of polysymplectic formalism in field theory (see
Section 3.8), we say that a connection (5.2.1) on the Legendre bundle V'Q R is
locally Hamiltonian if the exterior form 7jfi is closed, i.e.,
L,n =d(7jn) =o. (5.2.3)
It is readily observed that a connection 7 (5.2.1) on V'Q R is locally Hamiltonian
if and only if it obeys the conditions
ay - tvy = 0,
drf] - djli = 0,
dj-y' + a y =0.
(5.2.4a)
(5.2.4b)
(5.2.4c)
Example 5.2.2. Every connection
r : Q - J'Q C TQ,
r =9, + Fd
it
(5.2.5)
on the configuration bundle Q R gives rise to the locally Hamiltonian connection
r = KT = a
(
+ r
i
a
i
-p,a
J
r
,
a
j
(5.2.6)
(1.6.13) on the momentum phase space V'Q R such that
T\n = dH
r
,
H
r
= pjdtf - pjFdt.
(5.2.7)
5.2. HAMILTONIAN CONNECTIONS AND HAMILTONIAN FORMS 233

Locally Hamiltonian connections constitute an affine space modelled over the
linear space of vertical vector fields fl on the Legendre bundle V'Q >R, which
fulfill the same condition
d(tfjn) =0 (5.2.8)
as (5.2.3). These vector fields are locally the Hamiltonian vector fields in accordance
with the following lemma.
LEMMA 5.2.2. Every closed form-y\fl on the fibre bundle V'Q - R is exact.
Proof. Let us consider the decomposition
7=r+tf, (5.2.9)
where T is a connection on Q >R and f is its lift (5.2.6) onto the fibre bundle
V'Q >R, while -d is a vertical vector field on V'Q satisfying the relation (5.2.8).
It is easily seen that
d\n = aAdt,
where a is a 1-form on V'Q. Using the properties of the De Rham cohomology of a
manifold product, one can show that every closed 2-forma A dt on a manifold V'Q,
diffeomorphic to the product R x T'M, is exact, and so is 7J O [57]. Moreover, in
accordance with the relative Poincare lemma, we can write locally
ti\n = df Adt,
where / is a local function on V'Q. QBD
DEFI NI TI ON 5.2.3. An exterior 1-formH on the momentum phase space V'Q is
said to be a locally Hamiltonian form if
7
j n =dH
for a connection 7 on the fibre bundle V'Q >R.
234 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
By virtue of Proposition 5.2.2, there is one-to-one correspondence between the
locally Hamiltonian connections and locally Hamiltonian forms, considered through-
out modulo closed forms. In particular, the exterior form Hr (5.2.7) is a locally
Hamiltonian form.
DEFI NI TI ON 5.2.4. An exterior 1-form H on the momentum phase space V'Q is
called a Hamiltonian form if it is the pull-back
H = h*S - pidq* - Hit
(5.2.10)
of the Liouville form S (5.1.2) on the homogeneous Legendre bundle T'Q by a
section h of the fibre bundle
C : TQ - V'Q.
(5.2.11)
a
Remark 5.2.3. Note that, with respect to the universal unit system, a Hamiltonian
form is physically dimensionless.
Remark 5.2.4. Given a trivialization Q =R x M, the Hamiltonian form (5.2.10) is
the well-known integral invariant of Poincare-Cartan [6]. Therefore, the coefficient
H of the horizontal part of the Hamiltonian form (5.2.10) is said to be a Hamiltonian.
A glance at the expression (5.2.10) shows that Hamiltonians fail to be functions,
i.e., H $. )(V'Q), but make up an affine space modelled over the linear space of
functions on V'Q. *
For instance, every connection T on a configuration bundle Q R is an affine
section
p o r = -p.r
of the fibre bundle (5.2.11) (see (1.3.12)), and defines the Hamiltonian form H
r
(5.2.7) on the momentum phase space V'Q, where its Hamiltonian is
H =
Pi
F.
It follows that any Hamiltonian form on the momentum phase space V'Q admits
the splitting
H = H
T
- H
r
dt =
Pl
dq
l
- {p,F + Hr)dt, (5.2.12)
5.2. HAMILTONIAN CONNECTIONS AND HAMILTONIAN FORMS 235
where T is a connection on Q - R, and Hr is a real function on V'Q, called the
Hamiltonian function.
Since a connection T on a configuration space Q K is treated as a refer-
ence frame, i.e., as a kinematic object, one may say that, in accordance with the
decomposition (5.2.12), every Hamiltonian form H is split into kinematic and dy-
namic parts, and so is its Hamiltonian H. In Section 5.7, we will show that the
Hamiltonian function H
T
in the splitting (5.2.12) is the Hamiltonian energy func-
tion relative to the reference frame I \ Of course, the splitting (5.2.12) is not unique.
Nevertheless, every Hamiltonian form H admits the canonical splitting (5.2.12) as
follows.
We mean by a Hamiltonian map any fibred morphism
*:VQfJ
l
Q,
5t $= &(p), P e V'Q, (5.2.13)
over Q from the momentum phase space V'Q to the velocity phase space J
1
Q. Its
composition with the canonical morphism A (4.1.4) yields the fibred morphism
9: V'Q ^TQ,
$ =dt + Vd
{
. (5.2.14)
In particular, every connection T on a configuration bundle Q > R defines the
Hamiltonian map
f =T o Trn : V'Q - Q - J
x
Q
t
T = dt + r%.
Conversely, every Hamiltonian map $ (5.2.13) yields the associated connection
r* = $ o 6
on the configuration bundle Q R, where 0 is the global zero section of the vertical
cotangent bundle V'Q Q. For instance, we have
r~ = r
PROPOSI TI ON 5.2.5. Every Hamiltonian map (5.2.14) defines the associated Ha-
miltonian form
H* = - $J 0 = (A o $)*H =pidq
1
-
Pi
&dt
236 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
on V'Q, where 0 is the canonical 2-form (5.1.10).
PROPOSI TI ON 5.2.6. Every Hamiltonian form H on the momentum phase space
V'Q defines the associated Hamiltonian map
H : V'Q -> J
l
Q,
q]oH = d
x
U.

Proof. Let h : V'Q --> T'Q be a section of the fibre bundle (5.2.11), which
corresponds to the Hamiltonian form H, i.e., H = h'E. The vertical tangent map
Vh of the morphism h defines the linear fibred morphism
Vh : VV'Q -> V'Q x T'Q
Q
over V'Q. Therefore, it can be represented by the section
Vh = dp, {dq* - d'Hdt)
of the fibre bundle
VV'Q T'Q -> V'Q.
After natural contractions, this section is brought into the section
Vh = {dq* - d'Hdt) ft
of the pull-back
VQx(T'QVQ)->VQ,
Q
and takes its values into the image of the jet manifold J
1
Q by its canonical imbedding
(1.3.5) into T*QVQ. QED
COROLLARY 5.2.7. Every Hamiltonian formH on the momentum phase space V'Q
determines the associated connection
T
H
=H o 0
5.2. HAMILTONIAN CONNECTIONS AND HAMILTONIAN FORMS 237
on the configuration bundle Q R. It is called a Hamiltonian frame connectioi

In particular, we have
r
Wr
= r,
where H
r
is the Hamiltonian form (5.2.7) associated with the connection T on the
fibre bundle Q - R.
COROLLARY 5.2.8. Every Hamiltonian form H (5.2.10) on the momentum phase
space V'Q admits the canonical splitting
H = Hr
H
Hdt.
a
Let us turn now to the relationship between the locally Hamiltonian forms and
the Hamiltonian ones.
PROPOSI TI ON 5.2.9. Every locally Hamiltonian form on the momentum phase
space V'Q is locally a Hamiltonian form.
Proof. Given two locally Hamiltonian forms H*, and Hy on the momentum phase
space V'Q, their difference
(7 = fly Hy,
Ar = ( 7- V ) j n,
is a 1-form on V'Q such that the 2-form a A dt is closed and, consequently, exact
by virtue of Lemma 5.2.2. Hence, in accordance with the relative Poincare lemma,
we have
<T =fdt + dg,
where / and g are local functions on V'Q. Then, we deduce from the splitting
(5.2.9) that, in a neighbourhood of every point p V'Q, a locally Hamiltonian
form H~! can be written as
H^ = H
r
+ fdt,
238 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
and coincides with the pull-back of the Liouville form E on the cotangent bundle
T*Q by the local section
(t,q',Pi) - (t,q\Pi,P = -P,F + f)
of the fibre bundle (5.2.11).
QED
A converse of Proposition 5.2.9 is the following.
PROPOSI TI ON 5.2.10. Given a Hamiltonian formH on the momentum phase space
V'Q, there exists a unique connection JH on the Legendre bundle V'Q R, called
the Hamiltonian connection, such that
y\n = dH.
(5.2.15)
a
Proof. As in the polysymplectic case, let us introduce the first order Hamilton
operator on the phase space V'Q, which is associated with the Hamiltonian form
H. This Hamilton operator reads

H
: J'V'Q ^ AT'V'Q,

H
d
=dH - AJ fi =[(<?; - d'H)d
Pl
- (p
ti
+ dtTVjdq*) A dt, (5.2.16)
where A is the canonical monomorphism (4.1.4) and J
1
V'Q is the jet manifold of
the fibre bundle V'Q R, coordinated by
{t,q\pq],Pti)
It is readily observed that the kernel of the Hamilton operator E
H
(5.2.16) is an
image of the global section
7w =dt + d'Hdt - d
i
nd
i
(5.2.17)
of the affine jet bundle J
l
V"Q V'Q, which is precisely the desired Hamiltonian
connection (see Remark 5.2.6 below for a different proof). QED
Let us recall that Hamiltonian forms H on the momentum phase space V'Q
make up an affine space modelled over the linear space of horizontal densities fdt
5.2. HAMILTONIAN CONNECTIONS AND HAMILTONIAN FORMS 239
on the fibre bundle V'Q -> R. Then, as follows from the relation (5.2.15), Hamil-
tonian connections 7#constitute an affine space modelled over the linear space of
Hamiltonian vector fields (5.1.11):
H-H' = df,
7 - 7tf' = - $ /
The kernel of the covariant differential D
1H
, associated with a Hamiltonian con-
nection (5.2.17), is a closed imbedded subbundle of the jet bundle J ' V Q - R,
and defines the system of first order differential equations on the momentum phase
space V'Q. These are called the Hamilton equations
g| = d
{
n,
Pti = -diH
(5.2.18a)
(5.2.18b)
for the Hamiltonian formH. Note that, in comparison with the Lagrange equations,
the Hamilton ones are always well defined differential equations. The integral sec-
tions r : R 3 () - V'Q of the Hamiltonian connection (5.2.17) (or equivalently, the
integral curves of the horizontal vector field (5.2.17)) are solutions of the Hamilton
equations (5.2.18a) - (5.2.18b). Moreover, a glance at the equation (5.2.18a) shows
that the relation
J
1
{TTU or) = H or (5.2.19)
is fulfilled for any solution r of the Hamilton equations (5.2.18a) - (5.2.18b).
Remark 5.2.5. The Hamilton equations (5.2.18a) - (5.2.18b) are the particular
case of the first order dynamic equations
qt = Y, Pti = 7.
(5.2.20)
on the momentum phase space V'Q R, where 7 is a connection (5.2.1) on the
fibre bundle V'Q R.
The first order reduction of the equation of a motion of a point mass m subject
to friction in Example 4.9.7 exemplifies first order dynamic equations which are not
Hamilton ones. These equations read
1
Qt = P,
m
0
k
Pt = P
m
0
(5.2.21)
The connection
1 k
7 =d
t
+ pd
q
pd
p
m m
240 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
does not obey the condition (5.2.4c) for a locally Hamiltonian connection. At the
same time, the equations (5.2.21) are equivalent to the Hamilton equations for the
Hamiltonian associated with the Lagrangian (4.9.35).
Note that the Hamilton equations (5.2.18a) - (5.2.18b) can be introduced without
appealing to the Hamilton operator. On sections r of the fibre bundle V'Q R,
these equations
r' =&H o r, r<= -d,H o r
(5.2.22)
are equivalent to the relation
r'{u\dH) =0 (5.2.23)
which is assumed to hold for any vertical vector field u on V'Q R.
Remark 5.2.6. Every Hamiltonian form H on a. momentum phase space V'Q
defines a presymplectic structure. A Hamiltonian form H = h'E, by definition, is
a pull-back of the canonical Liouville form 5 (5.1.2) by means of a section h of the
fibre bundle T'Q >V'Q. Accordingly, its exterior differential
dH =h'dE = {dpi + diHdt) A {dq? - d'Tidt) (5.2.24)
is the pull-back of the canonical symplectic form (5.1.3), and is a presymplectic
form. The presymplectic form (5.2.24) has constant rank 2m since the form
{dH)
m
= {d
Pi
A dg')
m
- m{dp
x
A dq')
m
-
1
AdH/\dt (5.2.25)
is nowhere vanishing. It is also easily seen that
{dH)
m
Adty^O.
It follows that the pair (dH, dt) defines a cosymplectic structure on V'Q (see Remark
4.8.8). Since the presymplectic formdH is of constant rank 2m, its kernel Ker dH is
a 1-dimensional distribution which is spanned by the vector field ^
H
(5.2.17). Hence,
there is a unique vector field u =7// on V'Q such that
u\dH =0, u\dt = 1
(see [37]). This is a different proof of Proposition 5.2.10.
5.2. HAMILTONIAN CONNECTIONS AND HAMILTONIAN FORMS 241
PROPOSI TI ON 5.2.11. The Hamiltonian form (5.2.6) is a contact form on the mo-
mentum phase space V'Q if the function
j
H
\H = p
i
d
i
H-H=[H] (5.2.26)
nowhere vanishes [116].
Proof. Since a Hamiltonian connection 7#(5.2.17) is a nowhere vanishing vector
field, the condition
H A {dH)
m
0
for a Hamiltonian form H to be a contact form is equivalent to the condition
lH
\{H A (dH)
m
) = {
lH
\H){dH)
m
=[H\(dHr 0.
The result follows because the form (dH)
m
(5.2.25) is nowhere vanishing. QED
Note that one may try to add some exact form, e.g., the form cdt, c = const.,
to a Hamiltonian form H in order to make the function \H] nowhere vanishing. For
instance, the Hamiltonian form i /
r
(5.2.7) fails to be a contact one since [Hr] =0,
but the equivalent Hamiltonian form Hr - dt, where [Hr dt] = 1, is a contact
form.
If a Hamiltonian form H is a contact one, the corresponding Reeb vector field
(2.2.5) reads
E
H
= [WTV
(5.2.27)
By virtue of Proposition 2.2.6, one can then introduce the J acobi bracket defined
by the vector field (5.2.27) and by the bivector field w
H
on V'Q derived from the
relations (2.2.8) which read
w
H
(<P,.)\H = 0,
w
H
(<t>,.)}dH=-(4>-(E
H
\<t>)H),
whenever 0 is a 1-form on V'Q. We find
w
H
{<P, tf) =0Vi - o
l
4>i + Pia'E
H
\(p - p
{
(j>'E
H
\a,
242 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
where 4>, a axe arbitrary 1-forms on V'Q. The corresponding J acobi bracket on the
momentum phase space V'Q is
{J,9}H = w
H
(4>,a) + E
H
\(fdg-gdf) =
V,9}v + [HT'fobtfJtf - lfh
H
\dg),
where {/, g}
v
is the Poisson bracket (5.1.5) and
[/]=?.>'/- /, [g]=Pid
i
g-g.
Remark 5.2.7. Given the Poisson bracket (5.1.5) on the momentum phase space
V'Q, one can introduce the generalized Poisson bracket {., .}
w
(2.8.3) on the exterior
algebra Q'(V'Q), and the bracket {., .}
d
(2.8.5) on the quotient
0'(VQ)/dD'(VQ).
In particular, the generalized Poisson bracket (2.8.3) of Hamiltonian forms H and
H' reads
{H, H'}
w
=
Pi
(d"H' - d'H)dt.

5. 3 Canoni cal t rans f ormat i ons
Canonical transformations in time-dependent mechanics are not compatible with
the fibration V'Q Q of the momentum phase space in general.
DEFI NI TI ON 5.3.1. By a canonical automorphism is meant a vertical automorphism
p of the fibre bundle V'Q R, which preserves the canonical Poisson structure
(5.1.5) on the momentum phase space V'Q, i.e.,
{fP,gp}v = {f,g}vP-
D
It is easily seen that an automorphism p of V'Q R is canonical if and only if
it preserves the canonical form SI (5.1.9) on V'Q, i.e.,
n = p'n.
5.3. CANONICAL TRANSFORMATIONS 243
DEFI NI TI ON 5.3.2. The bundle coordinates of the fibre bundle V'Q - R are called
canonical if they are canonical for the Poisson structure (5.1.5).
It is readily observed that canonical coordinate transformations satisfy the rela-
tion
dp, dp
k
dp
k
dp,
dpUW
dqi dq
k
dq
k
dqi '
dp
k
dqi dq> dp
k
3
'
By very definition of the canonical formil, the holonomic coordinates of the vertical
cotangent bundle V'Q Q are the canonical coordinates. Accordingly, holonomic
automorphisms
(<?\P.)~ (<?",?: =f^
Pj
)
(5.3.1)
of V'Q Q, induced by the vertical automorphisms of the configuration bundle
Q R are also canonical.
PROPOSI TI ON 5.3.3. Locally Hamiltonian connections are transformed into each
other by canonical automorphisms, and so are locally Hamiltonian forms.
Proof. If 7 is a locally Hamiltonian connection for H, we have
7>(
7
)jn =(p-'rwn) =d[fjr
l
YH\
and Tp(y) is also a locally Hamiltonian connection. QED
PROPOSI TI ON 5.3.4. Let 7 be a complete locally Hamiltonian connection on V'Q >
R, i.e., the vector field (5.2.1) is complete. There exist canonical coordinates on V'Q
such that 7 =d
t
.
Proof. A glance at the relation (5.2.3) shows that each locally Hamiltonian con-
nection 7 is the generator of a local 1-parameter group of canonical automorphisms
of the fibre bundle V'Q - R. Let V
0
'Q be the fibre of V'Q ->R at 0 G R. Then
= 0,
= 0,
= *
i+
D
244 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
canonical coordinates of the symplectic manifold VQ = T'M dragged along inte-
gral curves of the complete vector field 7 determine the desired canonical coordinates
on V'Q. QED
In other words, a complete locally Hamiltonian connection 7 on the momentum
phase space V'Q in accordance with Proposition 4.1.2 defines a trivialization
QED
0 : V'Q - R x V
0
'Q
(5.3.2)
of the fibre bundle V'Q > R such that the corresponding coordinates of V'Q,
compatible with this trivialization, are canonical. However, it should be emphasized
that, although the fibre V
0
*Q is diffeomorphic to T'M, the trivialization (5.3.2) fails
to be a trivialization of the type V'Q =R x T'M since the trivialization morphism
rp is not a bundle morphism of the fibre bundle V'Q Q.
In particular, let H be a Hamiltonian form (5.2.12) such that the correspond-
ing Hamiltonian connection -y
H
(5.2.17) is complete. By virtue of Proposition 5.3.4,
there exists a trivialization of the phase space V'Q with respect to the global canon-
ical coordinates {q
A
,PA) (where q
A
are not coordinates on Q in general) such that
ft = dp
A
A dq
A
A dt,
1H = d
t
, dH = dp
A
A dq
A
,
and H reduces to the Hamiltonian form
H = p
A
dq
A
with the Hamiltonian H = 0. Then the corresponding Hamilton equations take the
form of the equilibrium equations
qt = 0,
PtA = 0
(5.3.3)
such that q
A
(t,q
t
,p
l
) and PA(t,q
l
,Pi) are constants of motion.
Accordingly, any Hamiltonian Ti can be locally brought into zero, and the cor-
responding Hamilton equations are reduced to the equilibrium ones (5.3.3) by local
canonical coordinate transformations.
Exampl e 5.3.1. Let us consider the 1-dimensional motion with constant acceler-
ation a with respect to the coordinates (t,q). Its Hamiltonian on the momentum
phase space R
3
R reads
P
2
5.3. CANONICAL TRANSFORMATIONS 245
The associated Hamiltonian connection is
1H =d
t
+ pd
q
+ ad
p
.
This Hamiltonian connection is complete. The canonical coordinate transformations
at
2
q =q-pt + , p' =p at (5.3.4)
bring it into fn =d
t
. Then, the functions q'(t, q
%
p) and p'(t,p) (5.3.4) are constants
of motion.
Exampl e 5. 3. 2. Let us consider the 1-dimensional oscillator with respect to the
same coordinates as in the previous Example. Its Hamiltonian on the momentum
phase space R
3
* K reads
K=\{p
2
+ q
2
)-
The associated Hamiltonian connection is
lH = dt+ pd
q
- qd
p
,
which is complete. The canonical coordinate transformations
q' = qcost psi nt, p' = p cos t + q sin t (5.3.5)
bring it into JH = d
t
. Then, the functions q^(i, q,p) and p'(t,q,p) (5.3.5) are con-
stants of motion.
It should be emphasized that, in general, canonical automorphisms do not trans-
form Hamiltonian forms into Hamiltonian forms, but only locally.
Let H be a Hamiltonian form (5.2.10) on a momentum phase space V'Q equipped
with the coordinates (t,q
x
,Pi). Given a canonical automorphism p, we have
d(p*H - H) =0,
where
p'H = pidp
1
-Ho pdt. (5.3.6)
Therefore, we can write locally
H -p'H = dS, (5.3.7)
246 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
where S(t, q',Pi) is a local function on a momentum phase space V'Q.
It should be emphasized that, if a configuration space Q is a contractible mani-
fold, the equality (5.3.7) is fulfilled globally, and the function S is defined everywhere
on V'Q.
Substituting (5.3.6) in (5.3.7), we obtain the relations
diS = Pi- Pjdtp',
d
l
S = -pjd
x
(P,
H-H' = p,d
t
p' - d
t
S.
Taken on the graph
A = {(p,p(p)) 6 V'Q x V'Q, p e V'Q}
of the canonical automorphism, the function S plays the role of a generating function
of canonical transformations.
Example 5.3.3. Given a canonical automorphism p, let
det( dV ) ^0.
(5.3.8)
Then, if the graph A can be coordinated by
(t,q',q
n
= q
t
op),
we obtain the familiar relations
dS
Pi=
dq~''
, _ dS
p
'~ do"'
H-H' = d
t
S(t,q\q").

(5.3.9)
Example 5.3.4. For instance, the generating function of a holonomic automor-
phism (5.3.1), where det(<9i<j'
J
) ^ 0, is
S(t,q'
3
,
Pi
) = -q
l
(t, q'
j
)
Pl
.

Let H be a Hamiltonian form whose associated Hamiltonian connection is com-
plete. By virtue of Proposition 5.3.4, there exists a canonical automorphismp which
bring its Hamiltonian into zero. If the condition (5.3.8) holds, then (5.3.9) is the
Hamilton-Jacobi equation.
5.4. THE EVOLUTION EQUATION 247
5.4 The evolution equation
Given a Hamiltonian formH (5.2.12) and the corresponding Hamiltonian connection
7H (5.2.17) on the momentum phase space V'Q, let us consider the Lie derivative
of a function / e D(V'Q) along the horizontal vector field 7/f, which reads
!*/ =7J4f =(d
t
+VHdi - W#)J.
(5.4.1)
This equality is the evolution equation in time-dependent mechanics. Substituting a
solution r of the Hamilton equations (5.2.22) in (5.4.1), we obtain the time evolution
of / along this solution:
L
-m/
or
= J
t
U
r
)-
It should be emphasized that, in comparison with the evolution equation (3.2.5)
in symplectic mechanics, the right-hand side of the evolution equation (5.4.1) is not
reduced to the Poisson bracket since a Hamiltonian Hona momentum phase space
of time-dependent mechanics is not a function, i.e., 7i D(V'Q). Of course, one
can define locally the Poisson bracket {H,f}v of a Hamiltonian H with a function
/ on V'Q. However, being equal to zero with respect to some coordinates, this
Poisson bracket does not necessarily vanish with respect to other coordinates.
Given the splitting (5.2.12) of a Hamiltonian form H, the evolution equation
(5.4.1) is written as
L
7
/ =d
t
f+(r% - diVp^f+iHrJh-
(5.4.2)
In particular, the following two consequences of this form of the evolution equation
should be pointed out.
Since the evolution equation (5.4.2) is not reduced to the Poisson bracket,
the integrals of motion in time-dependent mechanics cannot be defined as
functions on a momentum phase space, which are in involution with a Hamil-
tonian. Therefore, to obtain conservation laws in Hamiltonian time-dependent
mechanics, we follow the methods of field theory (see Section 5.7).
The second (kinematic) term in the right-hand side of this equality plays an
essential role under quantization. It makes the quantization procedure depen-
dent on a reference frame. In Appendix B, we will show that quantizations
under different reference frames fail to be equivalent in general.
248 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
Note that the kinematic term in the evolution equation (5.4.2) can be eliminated
at least locally by means of canonical transformations. Let a connection T in the
splitting of a Hamiltonian form (5.2.12) be complete. With respect to the coordinate
system (t, q') adapted to the reference frame T, the configuration bundle Q is triv-
ialized, and the corresponding holonomic coordinates (t, q',pi) on the momentum
phase space V'Q are canonical. With respect to these coordinates, the evolution
equation (5.4.2) takes the familiar form
L^
H
f = d
t
f + {H,f}
v
.
5.5 Degenerate systems
This Section is devoted to the relationship between Lagrangian and Hamiltonian
formalisms of time-dependent mechanics. This relationship is characterized by the
diagram
V'Q * JIQ
z\ _ j .
J
X
Q ^-V'Q
which fails to commute in general. Its jet extension is
jiy'Q
J
^J
l
J
l
Q
J'L | I PL
J
l
J
l
Q i^-J
l
V'Q
where the jet prolongations of Hamiltonian and Legendre maps read
J
l
H:J
l
V'Q^J
l
J
l
Q,
(<?\ q\, it), lit) oj'H = (q\ d'H, q\, dt&H),
J
x
l:J
l
J
l
Q^J
l
V'Q,
(<?\P.. q't.Pti) J
1
L = (g\7r q\
ty
d,7r,).
As we will show below, in the case of a hyperregular Lagrangian when the Leg-
endre map L (4.8.9) is a diffeomorphism, Lagrangian and Hamiltonian formalisms
are equivalent. Conversely, let H be a Hamiltonian form and JH the corresponding
5.5. DEGENERATE SYSTEMS 249
Hamiltonian connection (5.2.17) on the momentum phase space V'Q K. Let us
consider the composition of morphisms
J
l
Ho
lH
:V'Q^J
2
QdJ
l
J
l
Q,
(<M
0
.<4Wtf 7/f = (d'Kd'KjHlMd'H)).
(5.5.1)
If the Hamiltonian map H is a diffeomorphism, then J
l
H 07^0 H~
l
is a holonomic
connection on the jet bundle J
l
Q M, which defines a dynamic equation on the
velocity configuration space J
1
Q. This dynamic equation is equivalent to the system
of Lagrange equations for the Lagrangian function
= H-
U
[H]
which is the pull-back on J
X
Q of the function [H\ (5.2.26) by the morphism //"'.
Let us consider more general conditions for solutions of Hamilton equations on
a momentum phase space to be solutions of Lagrange equations and second order
dynamic equations on a velocity phase space, and vice versa.
Following the general scheme of polysymplectic Hamiltonian formalism [57, 158,
159], we say that a Hamiltonian form H on the momentum phase space V'Q is
associated with a Lagrangian L on the velocity phase space J
l
Q if H obeys the
conditions
LHoL = L,
H = H& + LoH
(5.5.2a)
(5.5.2b)
[161]. A glance at the condition (5.5.2a) shows that the composition L o H is the
projection operator
p,(p) =d, (t, ^, c^(p)), | peQ, (5.5.3)
from V'Q onto a submanifold
N = L{J
X
Q) C V'Q, (5.5.4)
called the Lagrangian constraint space. Given a Hamiltonian formH associated wit!
the Lagrangian L, this constraint space is defined by the relation (5.5.3). Accord-
ingly, the composition HoL is the projection operator from J
i
Q onto a submanifold
H(N) C J
l
Q. The relation (5.5.2b) takes the form
CoH = [H] (5.5.5)
250 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
everywhere on V'Q.
Remark 5. 5. 1. Unless otherwise stated, we regard N as a subset of the momentum
phase space V'Q, without a manifold structure. All objects are defined on the whole
phase space V'Q, and their restriction to N means only that their values at the
points of N C V'Q are considered.
PROPOSI TI ON 5.5.1. A Hamiltonian formH associated with a Lagrangian L obeys
the relation
H \s= H'HL \N,
(5.5.6)
where Hi is the Poincare-Cartan form (4.8.5).
Proof. The proof follows from direct computation. QED
Acting on both sides of the equality (5.5.5) by the exterior differential, we obtain
the relations
QED
d
t
H(p) = -(d
t
)(t,q>,d>H(p)), peN,
d
l
H{p) = -{d
i
C){t^,&H{p)), peN,
(
P
, - (a^)(t,^,^H))a'dw = o.
(5.5.7)
(5.5.8)
(5.5.9)
A glance at the relation (5.5.9) shows that:
the condition (5.5.2a) is a corollary of the condition (5.5.2b) if the Hamiltonian
map H is regular, i.e.,
det(a'^W) yt 0
at all points of the Lagrangian constraint space N;
the Hamiltonian map H is always degenerate outside the Lagrangian con-
straint space N.
Example 5.5.2. Let L = 0 be the zero Lagrangian. In this case, the Lagrangian
constraint space is N =0(Q), where 0 is the canonical zero section of the Legendre
bundle V'Q Q. The condition (5.5.2a) is fulfilled trivially for every Hamiltonian
map, while the condition (5.5.2b) takes the coordinate form
H = Pid
{
H.
5.5. DEGENERATE SYSTEMS 251
Any Hamiltonian form Hr (5.2.7) obeys this condition, and is associated with the
Lagrangian L = 0.
If a Lagrangian L is hyperregular, there exists a unique associated Hamiltonian
form
H = #_, + (L-
l
)'L (5.5.10)
such that
H = L~\
Pi=n
l
(q\d^H(q
X
,
Pk
)), g{ saw,*,(**,rf}).
(5.5.11)
As an immediate consequence of (5.5.11), we have
J
l
H = (J
1
L)-\
PROPOSI TI ON 5.5.2. Let L be a hyperregular Lagrangian, and H the associated
Hamiltonian form. The following relations hold:
H
L
= L'H, (5.5.12)
e
I
= (j
1
ly
H
, (5.5.13)

B
= {J
l
H)'
T
, (5.5.14)
where EH is the Hamilton operator (5.2.16) for H, and j is the Euler-Lagrange-
Cartan operator (4.8.16) for L. O
Proof. The proof follows from direct computation. QED
A glance at the relations (5.5.6) and (5.5.12) shows that the Poincare-Cartan
form is the Lagrangian counterpart of a Hamiltonian form, whereas it follows from
(5.5.13) and (5.5.14) that the Lagrangian counterpart of the Hamilton operator is
the Euler-Lagrange-Cartan operator j (4.8.16).
In particular, if 7// is a Hamiltonian connection for the associated Hamiltonian
formH (5.5.10), then, by virtue of the equality (5.5.14), the composition J
l
H o^
H
(5.5.1) takes its values into the kernel of the Euler-Lagrange-Cartan operator ^or.
more exactly, in the kernel of the Euler-Lagrange operator L- Then the composition
QED
J
1
/ f o
7
oL : ; ' ( j 3( t ,
l /
i
,
g
; )
M
(t, q\ q\ = VH o L, q\
t)
= d'H o L, q\
t
=
lH
\d(d'H) o L) =
(t, q\ q\, q\
t)
= <?|, q\t =J
H
\d(d'H) o L) J
2
Q
e
252 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
is a holonomic Lagrangian connection for L. Conversely, if L is a Lagrangian
connection for L, then
y
H
= J
l
L
L
o H
is a Hamiltonian connection for H. This proves the following assertion.
PROPOSI TI ON 5.5.3. Let L be a hyperregular Lagrangian and H the associated
Hamiltonian form (5.5.10).
(i) Let a section r of the fibre bundle V'Q - 1 be a solution of the Hamilton
equations (5.2.18a) - (5.2.18b) for the Hamiltonian form H. Then the section
c =7r
n
o r
of the fibre bundle Q R is a solution of the Lagrange equations (4.8.7) for the
Lagrangian L, while its first order jet prolongation c satisfies the Cartan equations
(4.8.17a) - (4.8.17b).
(ii) Conversely, if a section c of the jet bundle J
1
Q R is a solution of the
Cartan equations for the Lagrangian L, the section
r =L o c
of the fibre bundle V'Q > R satisfies the Hamilton equations (5.2.18a) - (5.2.18b)
for the Hamiltonian form H. O
It follows that, given a hyperregular Lagrangian, there is one-to-one correspon-
dence between the solutions of the Lagrange equations (and, consequently, of the
Cartan equations) and the solutions of the Hamilton equations for the associated
Hamiltonian form.
In the case of regular Lagrangian L, the Lagrangian constraint space N is an
open subbundle of the Legendre bundle V'Q Q. U N jt V'Q, an associated
Hamiltonian form fails to be defined everywhere on V'Q in general. At the same
time, an open constraint subbundle N can be provided with the appropriate pull-
back structure with respect to the imbedding N V'Q so that we may restrict our
consideration to Hamiltonian forms on N. If a regular Lagrangian is additionally
semiregular (see Definition 5.5.4 below), the associated Legendre morphism is a
diffeomorphism of J
l
Q onto the Lagrangian constraint space N and, on N, we can
restate all results true for hyperregular Lagrangians.
5.5. DEGENERATE SYSTEMS
253
Exampl e 5.5.3. Let Q be the fibre bundle R
2
- R with coordinates (t,q). Its jet
manifold J
l
Q 3 K
3
and its Legendre bundle V'Q = R
3
are coordinated by (t,q,q
t
)
and (tf,g,p), respectively. Put
L = exp(q
t
)dt.
(5.5.15)
This Lagrangian is regular, but not hyperregular. The corresponding Legendre map
reads
po L = expift.
It follows that the Lagrangian constraint space N is given by the coordinate relation
p >0. This is an open subbundle of the Legendre bundle, and L is a diffeomorphism
of J
X
Q onto N. Hence, there is a unique Hamiltonian form H on N which is
associated with the Lagrangian (5.5.15). Its Hamiltonian function reads
H = p{\np- 1).
This Hamiltonian form, however, fails to be smoothly extended to V'Q.
Hereafter, we will restrict our consideration of degenerate systems described by
semiregular Lagrangians. Most physically interesting Lagrangians, including the
quadratic ones, are of this type. On the other hand, there is a comprehensive
relationship between Lagrangian and Hamiltonian formalisms in this case [57, 158,
159, 190].
DEFI NI TI ON 5.5.4. A Lagrangian L is said to be semiregular, if the pre-image
L'
1
(p) of any point p N is a connected submanifold of the velocity phase space
J
l
Q.
PROPOSI TI ON 5.5.5. All Hamiltonian forms H associated with a semiregular La-
grangian L coincide with each other at the points of the Lagrangian constraint space
N, i.e.,
H \N H' |/V .
Moreover, the PoincareCartan form Hi for the Lagrangian L is the pull-back
H
L
= L'H,
faqi - C)dt = H(t,q
i
,n)dt, (5.5.16)
254 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
of any associated Hamiltonian formH by the Legendre map L.
Proof. Let u be a vertical vector field on the jet bundle J 'Q - Q. If u takes its
values into the kernel Ker TL of the tangent morphism to L, it is easy to see that
L
U
H

=0.
Hence, the Poincare-Cartan form H
L
for a semiregular Lagrangian L is constant
on the connected pre-image ~'(p) of each point p N. Then results follow from
(5.5.6). QED
Note that the Hamilton operators for Hamiltonian forms in Proposition 5.5.5
do not necessarily coincide at points of the Lagrangian constraint N because of the
derivatives of these forms.
Let H be a Hamiltonian form associated with a semiregular Lagrangian L. Act-
ing by the exterior differential on the relation (5.5.16), we obtain the equality
QED
{q\ - d"H o L)d-n
x
A dt (5.5.17) - (diC + di(H o L))dq
i
Adt = 0
or, equivalently, the equalities
7r
t>
(^-3>WoL ) =0,
Bi*M ~ &H o ) - (diC + (diH) o L) = 0.
Using the equality (5.5.17), one can extend the relation

z
= (j
l
ly
H
(5.5.18)
(5.5.13), but not necessarily the relation (5.5.14) to semiregular Lagrangians. The
relation (5.5.18) enables us to generalize item (i) of Proposition 5.5.3 for Hamiltonian
forms associated with semiregular Lagrangians.
PROPOSI TI ON 5.5.6. Let a section r of the fibre bundle V'Q > R be a solution
of the Hamilton equations for a Hamiltonian form H associated with a semiregular
Lagrangian L If r lives in the Lagrangian constraint space ./V, the section
c = n
n
or
of the fibre bundle Q > R satisfies the Lagrange equations for L, while its first
order jet prolongation
c =f f or =c
5.5. DEGENERATE SYSTEMS 255
obeys the corresponding Cartan equations (4.8.17a) - (4.8.17b).
Proof. Put c = H o r. Since r(R) C N, then
r = L o c, f= J 'L o J
l
c.
If r is a solution of the Hamilton equations, the exterior formH vanishes at points
of r(R). Hence, the pull-back form

T
= (j
1
ly
H
vanishes at points of c(R). It follows that the section c of the fibre bundle J 'Q > R
obeys the Cartan equations. By virtue of the equation (5.2.18a), we have c = c, and
the section c is a solution of the Lagrange equations, which lives in the submanifold
H(N). QED
In the case of semiregular Lagrangians, item (ii) of Proposition 5.5.3 can be
modified as follows.
PROPOSI TI ON 5.5.7. Given a semiregular Lagrangian L, let a section c of the jet
bundle J
l
Q * R be a solution of the Cartan equations (4.8.17a) - (4.8.17b). Let
H be a Hamiltonian form associated with L so that the corresponding Hamiltonian
map satisfies the condition
QED
HoLo-5= J
l
( f o4
(5.5.19)
Then the section
r = L o c,
r, =TTi(t,C
j
,4),
r* = c\
of the fibre bundle V'Q R is a solution of the Hamilton equations (5.2.18a) -
(5.2.18b) for H, which lives in the Lagrangian constraint space TV. D
Proof. The Hamilton equations (5.2.18a) hold by virtue of the condition (5.5.19).
Using the relations (5.5.17) and (5.5.19), the Hamilton equation (5.2.18b) is brought
into the Cartan equation (4.8.17b):
dtTCi o c = -{diH)oloc =
(d
t
c
t
cDdiKj o c +d{C o c.
-(3-&Hlc)d
t
n
3
oc + diCoc =
256 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
QED
Proposition 5.5.6 shows that, if H is a Hamiltonian form associated with a
semiregular Lagrangian L, every solution of the corresponding Hamilton equations,
which lives in the Lagrangian constraint space N, yields a solution of the Cartan
equations and of the Lagrange equations for L. Thus, the counterpart of degenerate
Lagrangian systems are the constraint Hamiltonian systems.
We will restrict our consideration to the case of a Lagrangian constraint space
N. This plays the role of a primary constraint space. In order that local solutions
of the Hamilton equations for a Hamiltonian form H exists on N, the Hamiltonian
connection 7#must be tangent to TV. This condition is given by the equations
Pi = %C(t,<j>,9iH),
(9, +d>Hd, - djH&)\d{p
t
- d\C{t, q>,d>H)) = 0,
(5.5.20a)
(5.5.20b)
where (5.5.20a) is the equation (5.5.3) of a Lagrangian constraint space, while
(5.5.20b) requires that the horizontal vector field JH is tangent to N at the point
(t, g
1
, Pi). In contrast with the case of autonomous constraint systems, the left-hand
side of the equation (5.5.20b) is not reduced to the Poisson bracket. Therefore, we
cannot follow the familiar procedure of constructing the final constraint space in
Section 3.5.
There is another possibility. Given a solution of the Lagrange equations, one can
try to find an associated Hamiltonian form H such that this solution is a solution
of the Hamilton equations for H. It may happen that different solutions of the
Lagrange equations require different Hamiltonian forms in general. Thus, degenerate
Lagrangian systems are described as multi-Hamiltonian systems.
Note that, in the case of semiregular Lagrangians, the condition (5.5.19) is the
obstruction for a solution c of the Cartan equations to be a solution of the Hamilton
equations and the Lagrange equations. At the same time, one can try to find a
family of associated Hamiltonian forms such that each solution of the Lagrange
equations is a solution of the Hamilton equations for some Hamiltonian form from
this family.
DEFI NI TI ON 5.5.8. We will say that a family of Hamiltonian forms H, associated
with a Lagrangian L, is complete if, for each solution c of the Lagrange equations,
there exists a solution r of the Hamilton equations for a Hamiltonian form H from
5.5. DEGENERATE SYSTEMS
257
this family so that
c =7Tnr, r = Lo c (5.5.21)
(see the relation (5.2.19)).
Let L be a semiregular Lagrangian. Then, by virtue of Proposition 5.5.7, such
a complete family of associated Hamiltonian forms exists if and only if, for every
solution c of the Lagrange equations for L, there is a Hamiltonian formH from this
family such that
H o L o c = c. (5.5.22)
Exampl e 5.5.4. Let L = 0. This Lagrangian is semiregular. Its Lagrange equations
reduce to the identity 0 = 0. Every section c of the configuration bundle Q R
is a solution of this equation. Given a section c, let T be a connection on the fibre
bundle Q >R such that c is an integral section of T. The Hamiltonian form Hr
(5.2.7) is associated with L = 0, and the Hamiltonian map H^ obeys the relation
(5.5.22). The corresponding Hamilton equations read
?! = r
1
,
Pu = -PjdiT
3
.
They have a solution
r =L o c,
r' = c\
n = o,
which lives in the Lagrangian constraint space p, =0.
The example below shows that a complete family of associated Hamiltonian
forms may exist when a Lagrangian is not necessarily semiregular.
Exampl e 5.5.5. Let Q be the fibre bundle R
2
R
1
in Example 5.5.3 with coor-
dinates (t,q). Put
= i(*)
3
-
The associated Legendre map reads
poL = <fi.
(5.5.23)
258 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
The corresponding Lagrangian constraint space N is given by the coordinate relation
p >0. It fails to be a submanifold of the momentum phase space V'Q. There exist
two associated Hamiltonian forms
H
+
= pdq - -p
3
'
2
dt,
H-=pdq+-p
3/2
dt,
on N, which correspond to the two different solutions
<7t = VP, Qt = -y/p
of the equation (5.5.23). The Hamiltonian forms H
+
and #_ make up a complete
family.
Remark 5.5.6. Let 7#be a Hamiltonian connection for a Hamiltonian form H
associated with a semiregular Lagrangian L. By virtue of the relation (5.5.18), the
composition J
l
H o -y
H
(5.5.1) takes its values into the kernel of the Euler-Lagrange
operator {,. Since
HoL\
S(Q)
= UH(Q),
the morphism
j'Ho-yol: j ' Q g f t / o i l H
(t, q\ qt = d'H o L, q\
t)
=Shi L, q\
t
= y
H
\d(d>H) o L)
restricted to H(Q) is the holonomic section over H(Q) C J
l
Q of the affine jet
bundle J
2
Q >J
l
Q. Let H(Q) be a closed submanifold, e.g., when a Lagrangian L
is almost regular (see Definition 5.5.9 below). Then the section J
l
H o^
H
oL can
be extended to a holonomic connection on the jet bundle J
l
Q R, which defines
a dynamic equation on the configuration space Q. In this case, given a solution r
of the Hamilton equations for H which lives in the Lagrangian constraint space ,/V,
its projection 7rn o r is a solution of both the Lagrange equations and the dynamic
one.
DEFI NI TI ON 5.5.9. A semiregular Lagrangian L is called almost regular if: (i) the
Lagrangian constraint space N is a closed imbedded subbundle i^ : N ^-> V'Q of
5.5. DEGENERATE SYSTEMS ZiO\y
the Legendre bundle V'Q Q and (ii) the Legendre map L : J
X
Q N is a fibre
bundle. D
PROPOSI TI ON 5.5.10. [57, 190]. Let L be an almost regular Lagrangian. On an
open neighbourhood in V*Q of each point q N, there exists a complete family of
local Hamiltonian forms associated with L. O
Let L be an almost regular Lagrangian L. Given an associated Hamiltonian form
H, the Hamiltonian map
H\
N
:N^ J
l
Q
restricted to N is a section of the fibre bundle L : J
l
Q N, and its image H(N) is
a closed imbedded submanifold of the configuration space J
l
Q. Therefore, it follows
from Remark 5.5.6 that each solution of the Lagrange equations for L, which is a
solution of Hamilton equations, is a solution of a dynamic equation.
Since solutions of the Lagrange equations for a degenerate Lagrangian may cor-
respond to solutions of different Hamilton equations, we can conclude that, roughly
speaking, the Hamilton equations involve some additional conditions in comparison
with the Lagrange equations. Therefore, let us separate a part of the Hamilton
equations which are defined on the Lagrangian constraint space Q in the case of an
almost regular Lagrangian.
Let
IT def . Tf
"N
l
N
be the restriction of a Hamiltonian form H associated with an almost regular La-
grangian L to the Lagrangian constraint space N. By virtue of Proposition 5.5.5,
this restriction, called the constrained Hamiltonian form, is uniquely defined. More-
over, the PoincareCartan form is the pull-back He = L'Hs of the constrained
Hamiltonian form HN. On sections r of the fibre bundle Q R, we can write the
constrained Hamilton equations
r'{u
N
\dH
N
) = 0,
(5.5.24)
where us is an arbitrary vertical vector field on the fibre bundle N >R [56, 57, 161].
For the sake of simplicity, we can identify a vertical vector field u
N
on TV > R with
260 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
its image Tz^(u^) and can bring the constrained Hamilton equations (5.5.24) into
the form
r'{u
N
\dH) =0,
(5.5.25)
where r is a section of the fibre bundle TV > R, and u^ is an arbitrary vertical vector
field on N R. It should be emphasized that these equations fail to be equivalent
to the Hamilton equations (5.2.23) restricted to the Lagrangian constraint space N.
If an almost regular Lagrangian admits associated Hamiltonian forms (see Propo-
sition 5.5.10), the following three assertions together with Proposition 5.5.7 give
the relationship between Cartan, Hamilton-De Donder, Hamilton, and constrained
Hamilton equations [57, 161].
PROPOSI TI ON 5.5.11. For any Hamiltonian form H associated with an almost
regular Lagrangian L, every solution r of the Hamilton equations which lives in the
Lagrangian constraint space N is a solution of the constrained Hamilton equations
(5.5.25).
Proof. For any vertical vector field Ufj on the fibre bundle N * K, the vector field
Tiw(ww) is obviously a local vertical vector field on the fibre bundle V'Q R.
Then we have
r'(u
N
\dH
N
) = r'(u
N
\i%dH) = T'[Ti
N
{u
N
)\dH) =0
if r is a solution of the Hamilton equations (5.2.23) for the Hamiltonian form H.
QED
PROPOSI TI ON 5.5.12. A section c of the jet bundle J
l
Q ->R is a solution of the
Cartan equations (4.8.18) if and only if Loc is a solution of the constrained Hamilton
equations (5.5.25).
Proof. Let u
N
be a vertical vector field on the fibre bundle N R. Since L is a
submersion, there exists a vertical vector field v on the jet bundle J
l
Q > R such
that
TL ov = us-
5.5. DEGENERATE SYSTEMS 261
For instance, v is the horizontal lift of u by means of a connection on the affine jet
bundle J
l
Q Q. Let a section c : M J
l
Q be a solution of the Cartan equations
(4.8.18). Then we have
(Loc)*(u
N
\dH
N
) = c*{v\dH
L
) = 0. (5.5.26)
It follows that the section L oc : R ^* N is a. solution of the equations (5.5.25).
The converse is obtained by running (5.5.26) in reverse, bearing in mind that the
restriction of any vector field v on J
l
Q to c(R) is projectable by L. QED
PROPOSI TI ON 5.5.13. The constrained Hamilton equations (5.5.24) are equivalent
to the Hamilton-De Donder equations (4.8.25).
Pr oof. Let H be a Hamiltonian form associated with an almost regular Lagrangian
L, and let h be the corresponding section of the fibre bundle
C : T*Q - VQ
(see Definition 5.2.4). This section yields the morphism
h
N
=h o i
N
: N - T*Q
which does not depend on the choice of H. This is a local section of the fibre bundle
T*Q - V*Q over N C VQ, i.e,
C,oh
N
= lAN. (5.5.27)
Moreover, we have
H
N
= i*
N
H = i*
N
(h*E) = h*
N
E,
whenever H is a Hamiltonian form associated with the Lagrangian L. In accordance
with the relation (5.5.16),
HL = h
N
o L,
(5.5.28)
where HL is the Legendre morphism (4.8.21) associated with the Poincare-Cartan
form Hi. Substituting (4.8.22) in (5.5.28), we obtain
H
L
= h
N
oC,o H
L
.
262 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
It follows that
h
N
C \z
L
Wi i (Z
L
),
(5.5.29)
where i
L
(Z
L
) = H
L
(J
l
Q) is the image of the Legendre morphism H
L
. A glance at
the relations (5.5.27) and (5.5.29) shows that there is the bundle isomorphism
Z
L

N
over Q. Since H^ h*
N
E and E
L
ilH, we have
H
N {i
L
l
o h
N
yz
L
, ~L {chyH
N
.
Hence, the Hamilton-De Donder equations (4.8.25) are equivalent to the constrained
Hamilton equations (5.5.24). QED
5.6 Quadratic degenerate systems
As an important illustration of Propositions 5.5.6 and 5.5.7, let us describe the
complete families of Hamiltoman forms associated with almost regular quadratic
Lagrangians. The Hamiltonians of these Hamiltonian forms are quadratic in mo-
menta.
Remar k 5.6.1. Since Hamiltonians in time-dependent mechanics are not scalar
functions on a momentum phase space, one cannot apply to them the well-known
analysis of the normal forms [18] (e.g., quadratic Hamiltonians [6] in symplectic
mechanics).
Given a configuration bundle Q R, let us consider a quadratic Lagrangian
C
-a
ijq
\4, + b
iq
l +c, (5.6.1)
where a, b and c are local functions on Q. This property is global due to the
transformation law of the velocity coordinates q\. The associated Legendre map
reads
Pi o L = a^ql + bi.
(5.6.2)
5.6. QUADRATIC DEGENERATE SYSTEMS 263
LEMMA 5.6.1. The Lagrangian (5.6.1) is semiregular.
Pr oof. For any point p of the Lagrangian constraint space N (5.5.4), the system of
linear algebraic equations (5.6.2) for q\ has solutions which make up an afHne space
modelled over the linear space of solutions of the homogeneous linear algebraic
equations
0= Oy^,
where q> are the coordinates on the vertical tangent bundle VQ. This affine space
is obviously connected. QED
Let us assume that the Lagrangian L (5.6.1) is almost regular, i.e., the matrix
a is of constant rank.
The Legendre map (5.6.2) is an affine morphism over Q. It defines the corre-
sponding linear morphism
L:VQ ^VQ,
p
{
oL = a
i:j
q>,
whose image N is a linear subbundle of the Legendre bundle V*Q ^ Q. Accordingly,
the Lagrangian constraint space N, given by the equations (5.6.2), is an affine
subbundle, modelled over N, of the momentum phase space VQ Q. Hence, the
constraint fibre bundle N >Q has a global section. For the sake of simplicity, let us
assume that this is the canonical zero section 0(Q) of the vertical cotangent bundle
VQ -> Q. Then N = N.
The kernel
KerL = L-
1
(0(Q))
of the Legendre map is an affine subbundle of the affine jet bundle J
X
Q >Q, which
is modelled over the vector bundle
KeiL = L~\0(Q))cVQ.
Then there exists a connection
T : Q - Ker L,
(5.6.3)
aijP + h = 0,
(5.6.4)
264
CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
on the configuration bundle Q- R, which takes its values into KerL. It is called
a Lagrangian frame connection. With this connection, the quadratic Lagrangian
(5.6.1) can be brought into the form
c
= 2
a
y9r?f +
c
'>
* = 4 - r'-
For instance, if the quadratic Lagrangian (5.6.1) is regular, there is a unique solution
(5.6.3) of the algebraic equations (5.6.4).
PROPOSI TI ON 5.6.2. There exists a linear map
<r : V*Q -> VQ,
(5.6.5)
q
l
oa = a
t]
p
i
,
over Q such that
L o a o i
N
= i^.
D
Proof. The map (5.6.5) is a solution of the algebraic equations
dija^akb = (lib-
(5.6.6)
The matrix ay is symmetric. After diagonalization, let this matrix have non-
vanishing components a
AA
, A & I. Then a solution of the equations (5.6.6) takes
the form
1

AA =
^ X'
OAA' = 0,
A, A
1
6 /, A? A',
while the remaining components are arbitrary. In particular, there is a solution
1

AA
= ^ u -
OAA' =0, OCB = 0, B*I,
(5.6.7)
which satisfies the relation
a = a o L o a,
(5.6.8)
<T'J = a
ik
a
kb
^
b
.
5.6. QUADRATIC DEGENERATE SYSTEMS 265
QED
From now on, by a is meant the solution (5.6.7). If the Lagrangian (5.6.1) is
regular, the map (5.6.5) is uniquely determined by the equation (5.6.6).
The Lagrangian frame connection (5.6.3) and the map (5.6.5) play a prominent
role in the construction below.
PROPOSI TI ON 5.6.3. The matrix a defines the splitting
J
1
Q = K erZeI m(aoZ),
Q
(5.6.9)
4 = [qj - <?
k
{a.kd + h)} + \o
ik
{akj<& +Ml-
of the velocity phase space J
l
Q. D
In particular, it follows that, since L and a are linear morphisms, their compo-
sition L o a is a fibration of the momentum phase space V*Q over the Lagrangian
constraint space N.
PROPOSI TI ON 5.6.4. There is also the following splitting of the momentum phase
space
V*Q Ker aN,
Q
Pi
\pi - aijcr^pk] + [aija
3k
p
k
].
a
COROLLARY 5.6.5. Every vertical vector field
u =u'di + Uid'
on the momentum phase space V*Q > M admits the decomposition
U (u Upf) + UN, (5.6.10)
m = [Ui - aija
3k
u
k
] + [atja^Uk],
where
u
N
= u
l
di + aijt7
jk
u
k
d'
266 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
is a vertical vector field on the Lagrangian constraint space N >R.
Given the linear map a (5.6.5) and the Lagrangian frame connection T (5.6.3),
let us consider the affine Hamiltonian map
$ =T +a : V'Q -> J'Q,
(5.6.11)
$' = F(g)+<7
y
Pi
,
and the Hamiltonian form
H = H* + Ldt,
H = ntf - [r(
Pi
- -h) + -OKPS - c]dt.
(5.6.12)
It is readily observed that H = $ and, by virtue of (5.6.8), the Hamiltonian form
H is associated with the quadratic Lagrangian (5.6.1).
We aim to show that the Hamiltonian forms (5.6.12), parameterized by the
Lagrangian frame connections T (5.6.3), constitute a complete family.
Given the Hamiltonian form (5.6.12), let us consider the Hamilton equations
(5.2.18a) for sections r of the fibre bundle V'Q - R. They read
c = (f +a) o r, c = 7Tn o r, (5.6.13)
or
V
(
r
i
= a
ij
r
jt
where
y
t
r' =<V - ( T o c)
1
is the covariant derivative relative to the connection T. With the splitting (5.6.9),
we have the surjections
S = pr
l
: J
l
Q-*KetL,
S : q\ ql - o
ik
{a
kj
qi + b
k
),
and
T = pr
2
: J
l
Q - Im(cr o Z),
F:qi->a
ik
(a
kj
qi + b
k
).
5.6.
QUADRATIC DEGENERATE SYSTEMS 267
With respect to these surjections, the Hamilton equations (5.6.13) are split into the
following two parts:
S o c =T o c,
V
t
r
{
=a
ik
(a
kj
d
t
ri + b
k
),
(5.6.14)
and
T o c = a o r,
(5.6.15)
a
ik
{a
kj
d
t
T
J
+ h) = a
ik
r
k
.
The Hamilton equations (5.6.14) are independent of the canonical momenta r
k
and play the role of gauge conditions. Moreover, for every section c of the fibre
bundle Q R (in particular, for every solution of the Lagrange equations for the
Lagrangian L), there exists a Lagrangian frame connection T (5.6.3) such that the
equation (5.6.14) holds. Indeed, let F* be a connection on the fibre bundle Q > R
whose integral section is c. Put
T^SoT
1
,
r = r '
i
-<7
i l c
(a
fcJ
r '
i
+ 6
fc
).
In this case, the Hamiltonian map (5.6.11) satisfies the relation (5.5.22) for c, i.e.,
$ o L o c = c.
Hence, the Hamiltonian forms (5.6.12) constitute a complete family. The Hamil-
tonian forms from this family differ from each other only in the Lagrangian frame
connections T (5.6.3) which lead to the different gauge conditions (5.6.14). It is
readily observed that a Lagrangian frame connection V is also a Hamiltonian frame
connection for the corresponding Hamiltonian form (5.6.12).
PROPOSI TI ON 5.6.6. For every Hamiltonian form H (5.6.12), the Hamilton equa-
tions (5.2.18b) and (5.6.15) on sections of the constraint fibre bundle N > R are
equivalent to the constrained Hamilton equations (5.5.25).
Proof. In accordance with the decomposition (5.6.10) of a vertical vector field u on
the momentum phase space V*Q R, the constrained Hamilton equations (5.5.25)
take the form
r*(a
ij
a
:
>
k
d
i
\dH)=0, (5.6.16a)
r*{d
t
\dH) = 0.
(5.6.16b)
268 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
The equations (5.6.16b) are obviously the Hamilton equations (5.2.18b) for H. Bear-
ing in mind the relations (5.6.4) and (5.6.8), one can easily bring the equations
(5.6.16a) into the form (5.6.15). QED
It follows that the equations (5.6.14) are the additional conditions which make
the Hamilton equations differ from the constrained Hamilton equations (5.5.25) and
the Lagrange equations in the case of a quadratic Lagrangian.
COROLLARY 5.6.7. By virtue of Proposition 5.5.12, a section c of the jet bundle
J
l
Q - 1 is a solution of the Cartan equations (4.8.17a) - (4.8.17b) for the almost
regular Lagrangian (5.6.1) if and only if Loc is a solution of the constrained Hamilton
equations (5.2.18b) and (5.6.15).
It follows that the equations (5.6.14) are responsible for the obstruction condition
(5.5.19) for solutions c of the Cartan equations to be solutions of the Hamilton
equations and of the Lagrange equations for the Lagrangian (5.6.1). It is readily
seen that the equations (5.6.15) do not contribute to the obstruction condition
(5.5.19). If r = L o c, these equations hold for solutions c of the Cartan equations
(4.8.17a).
PROPOSI TI ON 5.6.8. Let c be a solution of the Cartan equations for the almost
regular quadratic Lagrangian L (5.6.1). Let Co be a section of the fibre bundle
VQ R, which takes its values into Ker L and projects onto c =7TQ O C. Then the
sum c +Co over Q is also a solution of the Cartan equations.
Proof. The proof is an immediate consequence of the relation
r =L oc =L o(c +Co).
QED
Since Hamiltonian forms associated with an almost regular quadratic Lagrangian
(5.6.1) constitute a complete family, that is, each solution of the Lagrange equations
is also a solution of some Hamilton equations, we can conclude that each solution
of the Lagrange equations is a solution of some dynamic equation.
5.7. HAMILTONIAN CONSERVATION LAWS 269
5.7 Hamiltonian conservation laws
As was mentioned above, the notion of integrals of motion as functions in involution
with a Hamiltonian can not be extended to time-dependent mechanics because its
Hamiltonian is not a scalar function under time-dependent transformations.
In Section 4.12, we have studied the conservation laws in Lagrangian time-
dependent mechanics. Therefore, in order to discover the conservation laws in
Hamiltonian time-dependent mechanics, let us consider the following construction.
Given a Hamiltonian form H (5.2.10) on a momentum phase space V*Q, let us
consider the Lagrangian
L
H
d
=(Piqi-H)dt (5.7.1)
(cf. (4.4.6)) on the jet manifold J
l
V*Q of the momentum phase space V*Q * M.
Note that the Lagrangian (5.7.1) plays a prominent role in the functional integral
approach to mechanics (see Remark 5.11.4 below). Here, our utilization of this
Lagrangian is based on the following two assertions.
LEMMA 5.7.1. The Poincare-Cartan form for the Lagrangian LH (5.7.1) coin-
cides with the pull-back onto J
X
V*Q of the Hamiltonian form H by the projection
J
1
V*Q ~* V*Q, while the Euler-Lagrange operator for LH coincides with the pull-
back onto J
2
V*Q of the Hamilton operator for H by projection J
2
V*Q > J
l
V*Q.
a
Pr oof. The proof is straightforward. QED
It follows from this Lemma that the Lagrange equations for the Lagrangian L
H
(5.7.1) are equivalent to the Hamilton equations for the Hamiltonian form H.
LEMMA 5.7.2. Let u be a vector field (4.12.2) on a configuration space Q, and
u = u
l
d
t
+ u'di diU
j
pjd'
(5.7.2)
its lift onto the momentum phase space V*Q in accordance with the formula (1.6.13).
Then, there is the equality
L ;H = Lji~L
H
,
where J
l
u is the jet prolongation (1.3.7) of the vector field u onto J
l
VQ.
270
CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
Proof. The proof follows from direct computation.
QED
Thus, we can use the general procedure of constructing Lagrangian conservation
laws in Section 4.12 for the Lagrangian L
H
in order to obtain the Hamiltonian
conservation laws.
Let us apply the first variational formula (4.8.4) to the Lagrangian (5.7.1). Then
the weak identity (4.12.3) reads
-u%H - u'dtH + Pidttf ~ -dt%,
(5.7.3)
X = -pitf + u'H.
(5.7.4)
In the case of a vertical vector field u, when u
(
= 0, the weak identity (5.7.3)
leads to the weak equality
L J I ^H ~dt{piU
%
).
If \JJI~LH =0, we have the weak conservation law
0 pa dt(piu')
of the Noether current
% = -piU*. (5.7.5)
For a reference frame Y (5.2.5) on the configuration bundle Q > K, the weak
identity (5.7.3) takes the form
-d
t
H - rdiH+PidtV -dtH
T
,
where H? =Ji Pi P is the Hamiltonian function in the splitting (5.2.12).
In the case of semiregular Lagrangians, there is the following relationship between
Lagrangian and Hamiltonian conserved currents.
PROPOSI TI ON 5.7.3. Let H be a Hamiltonian form on the momentum phase space
V*Q, which is associated with a semiregular Lagrangian L on the velocity phase
space J
1
^. Let r be a solution of the Hamilton equations (5.2.18a) - (5.2.18b) for
H, which lives in the Lagrangian constraint space N, and c the associated solution
of the Lagrange equations for the Lagrangian L so that the conditions (5.5.21) are
5.8. TIME-DEPENDENT SYSTEMS WITH SYMMETRIES 271
satisfied. Let u be a vector field (4.12.2) on the configuration space Q > R, Then
we have
T{r) = 1{Hor),
X(Zoc)=X(c),
where X is the current (4.12.4) on the configuration space J
X
Q, and X is the current
(5.7.4) on the phase space V*Q.
Pr oof. The proof is straightforward. QED
In accordance with Proposition 5.7.3, the Hamiltonian counterpart of the Lagran-
gian energy function Xr (4.12.16) relative to a reference frame T is the Hamiltonian
function HT in the splitting (5.2.12). Therefore, one can think of the Hamiltonian
function Hr as being the Hamiltonian energy function relative to the reference frame
T. In particular, if P =0, we obtain the well-known energy conservation law
d
t
H d
t
H
relative to the coordinates adapted to the reference frame V. This is the Hamiltonian
variant of the Lagrangian energy conservation law (4.12.17).
5.8 Time-dependent systems with symmetries
In mechanics, just as in field theory, systems with symmetries can be described in
terms of bundles with structure groups.
Let 7Tp : P R be a principal bundle with a structure Lie group G. Let
Q = (P x M)/G (5.8.1)
be a P-associated fibre bundle.
By gauge transformations of a principal bundle P are meant its vertical auto-
morphisms $ which are equivariant with respect to the canonical action (1.5.4) of
the structure group G on P on the right, i.e.,
$ o r
g
r
g
o $, VgeG. (5.8.2)
Due to the property (5.8.2), gauge transformations $ of a principal bundle yield the
gauge transformations of a P-associated fibre bundle
$
Q
: (P x M)/G - ($(P) x M)/G.
272 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
Thus, one can treat a fibre bundle Q (5.8.1) as a configuration space of a mechanical
system with the gauge symmetry group G.
Every gauge transformation of the configuration bundle Q- l gives rise canoni-
cally to the holonomic vertical automorphisms (5.3.1) of the momentum phase space
VQ > M. They are canonical automorphisms for the canonical Poisson structure
(5.1.5) on VQ. Moreover, this group action is Hamiltonian as can be seen in the
following way (see Definition 3.7.1).
Any 1-parameter group of gauge transformations of a principal bundle P defines
a vector field (its generator) on P which is invariant under the canonical action of
the group G on P. There is one-to-one correspondence between these right-invariant
vector fields on P, called principal vector fields, and the sections of the gauge algebra
bundle VQP (1.5.6), whose typical fibre is the right Lie algebra jj
r
.
Accordingly, any 1-parameter group of gauge transformations of a configuration
space Q, associated with P, defines a principal vertical vector field
C = C(g)d,
on the configuration space Q. This vector field gives rise to the vector field
Z^cat-atfpjtf
on the momentum phase space (see (5.7.2)). It is readily observed that this is the
Hamiltonian vector field for the Noether current
J = Pi(q)
(5.8.3)
(5.7.5) which is defined globally on VQ (see Example 3.7.4). It follows that, given
a finite dimensional Lie group of gauge transformations, one may apply the familiar
theorems of reduction, e.g., Theorem 3.7.9 to time-dependent mechanics, but not
Theorem 3.7.10 and other results which involve the condition of an invariance of a
Hamiltonian under a group action. Hamiltonians in time-dependent mechanics fail
to be invariant under time-dependent transformations, including gauge transforma-
tions.
Remar k 5.8.1. It should be emphasized that gauge transformations make up a
group whose suitable Sobolev completion is a Banach Lie group. The possible ex-
tension of the above-mentioned theorems to this infinite-dimensional Lie group in
the spirit of [36] seems promising. At the same time, we will show below that a Ha-
miltonian form may possess a finite-dimensional Lie group of gauge transformations.

5.8. TIME-DEPENDENT SYSTEMS WITH SYMMETRIES
273
Let us restrict our consideration to the case of a vector bundle Q > R. A
principal vector field on this vector bundle reads
= ar(t)e
m
i
i
q'a
i
, (5.8.4)
where e
m
'j are generators of the action of the group G on the typical fibre M, and
a
m
(t) are local functions on R. The vector field (5.8.4) gives rise to the vector field
| =a
m
(t)E
m
y9, - a
m
(t)e
m
i
jPl
d>
on the momentum phase space VQ. It is readily observed that this vector field
is the Hamiltonian vector field for the Noether current p;' (5.8.3) relative to the
canonical Poisson structure on VQ. Moreover, these Noether currents constitute
the Lie algebra with respect to the Poisson bracket
{p
i
e,pd'
j
}v=Pi[^']
i
.
(5.8.5)
In particular, let p
f
* and Pi" be conserved Noether currents for a Hamiltonian
form H on the momentum phase space VQ. Since
K,e] = [f,r], J
l
[U'\ = [J
1
lJ
1
Z]
and
Lj i j i /f =0,
Lji^Lff =0,
we obtain
L
J M'1
L H =
'
i.e., the Noether current (5.8.5) is also conserved. Thus, conserved Noether currents
in time-dependent mechanics constitute a Lie algebra (cf. Example 3.7.4). This
algebra obviously is finite-dimensional, and corresponds to some finite-dimensional
Lie group of gauge symmetries of a Hamiltonian form H. As was shown, an action
of this group on a momentum phase space VQ is Hamiltonian, with an equivariant
momentum mapping.
This result remains true for Noether currents (4.12.14) in Lagrangian mechanics.
Given a Lagrangian L, the Noether currents (4.12.14) [conserved Noether currents
(4.12.14)] along vector fields (5.8.4) constitute a Lie algebra with respect to the
Poisson bracket (4.8.13).
274 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
5.9 Systems with time-dependent parameters
Let us consider a configuration space which is a composite fibre bundle
Q^E - >R ,
(5.9.1)
coordinated by (, a
m
, q') where (t, a
m
) are coordinates of the fibre bundle E > R.
We will treat sections h of the fibre bundle R as time-dependent parameters,
and say that the configuration space (5.9.1) describes a mechanical system with
time-dependent parameters. Let us call E - R the parameter bundle. Note that
the fibre bundle Q E is not necessarily trivial.
Let us recall that, by virtue of Proposition 1.6.2, every section h of the parameter
bundle E >R defines the restriction
Q
h
= h'Q
(5.9.2)
of the fibre bundle Q >E to h(R) C E, which is a subbundle i
h
: QH *- Q of
the fibre bundle Q R. One can think of the fibre bundle Q
h
> R as being a
configuration space of a mechanical system with the background parameter function
h(t).
The velocity momentum space of a mechanical system with parameters is the
jet manifold J
X
Q of the composite fibre bundle (5.9.1) which is equipped with the
adapted coordinates
{t,o ,q ,a
t
,q
t
).
Let the fibre bundle Q >E be provided with a connection
A =dt (d
t
+ A\d
{
) +da
m
<g>(d
m
+<,<%). (5.9.3)
Then the corresponding vertical covariant differential
D:J
X
Q^ V
Z
Q,
D = (
q
i-A\- A^o?)*, (5.9.4)
(1.6.11) is defined on the configuration space Q. Given a section h of the parameter
bundle E >R, its restriction to Pihi^Qh) C J
l
Q is precisely the familiar covariant
differential on Qh corresponding to the restriction
A
h
=d
t
+ {A
i
m
d
t
h
m
+ (Aoh)\)d
i
(5.9.5)
5.9. SYSTEMS WITH TIME-DEPENDENT PARAMETERS 275
of the connection A to h(R) C E. Therefore, one may use the vertical covari-
ant differential D in order to construct Lagrangians for a mechanical system with
parameters on the configuration space Q (5.9.1).
We will suppose that a Lagrangian L depends on derivatives of parameters a
only through the vertical covariant differential D (5.9.4), i.e.,
L = C{t, a
m)
q\q\ ~ A* - A^o?)*. (5.9.6)
Such a Lagrangian is obviously degenerate because of the constraint condition
n
m
+ A^i = 0.
As a consequence, the total system of the Lagrange equations
(di - dt%)C =0, (5.9.7)
(d
m
- dt&JC =0 (5.9.8)
admits a solution only if the very particular relation
(d
m
+ A
i
m
d
i
)C + TT
i
d
t
Ai
n
= 0 (5.9.9)
holds. However, we believe that parameter functions are background, i.e., inde-
pendent of a motion. Therefore, only the Lagrange equations (5.9.7) should be
considered.
In particular, let us apply the first variational formula (4.12.8) in order to obtain
conservation laws of a mechanical system with time-dependent parameters. Let
u = %% + u
m
{t, a
h
)d
m
+ u
{
(t, a
k
, q
j
)d
i
, u* = 0,1, (5.9.10)
be a vector field which is projectable with respect to both the fibration Q- * B and
Q E. Then the Lie derivative L^L of a Lagrangian L along the vector field u
(5.9.10) reads
[u% + u
m
d
m
+ v.% + dtiTdl + dtu'dl}L =
(u
m
- a^u
l
)d
m
C +7r
m
d,(u
m
- aj V) +
(' - <?>% - *[ri(*flj - u*) - u'],
dt = d
t
+ <r?d
m
+ qtdi + a^d
l
m
+ &%.
On the shell (5.9.7), the weak identity
[u% + u
m
d
m
+ u% + dtu
m
d
l
m
+ dtu'dfiC
(u
m
- ofu^drnC + ^mdt{u
m
- ffjV) - *[&% ~ ') ~ ^C]
276 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
is fulfilled. If the Lie derivative L^L vanishes, we obtain the conservation law
0 w (u
m
- aT^)d
m
C + n
m
dt(u
m
- a?u
l
) - ^[^(u'gj - u*) - u]
(5.9.11)
for a system with time-dependent parameters.
Now, let us describe such a system in the framework of Hamiltonian formalism.
Its momentum phase space is the vertical cotangent bundle V*Q. Given a connection
(5.9.3), we have the splitting (1.6.10b) which reads
V*Q = A{VZQ){QxV*Y,),
Pidq
1
+ p
m
da
m
=
Pi
(dq' - A^da"
1
) + (p
m
+ A^da.
Then VQ can be provided with the coordinates
Pi=Pi, Pm=Pm+A
l
rn
p
i
,
compatible with this splitting. However, these coordinates fail to be canonical in
general. Given a section h of the parameter bundle S- *K , the submanifold
{a = h(t),
Pm
= Q}
of the momentum phase space V*Q is isomorphic to the Legendre bundle V*Qh of
the subbundle (5.9.2) of the fibre bundle Q >R, which is the configuration space
of a mechanical system with the parameter function h(t).
Let us consider Hamiltonian forms on the momentum phase space VQ which
are associated with the Lagrangian L (5.9.6). Given a connection
r = d
t
+ r
m
d
m
(5.9.12)
on the parameter bundle E * K, the desired Hamiltonian form
H = {p
x
dq
{
+ Pmda
m
) - faA" + p
m
T
m
+H(t, a
m
, q^p^dt (5.9.13)
can be found, where
B = d
t
+ r
m
d
m
+ B'd
{
, B
1
= A
i
+ y^I ,
is the composite connection (1.6.8) on the fibre bundle Q >K, which is defined by
the connection A (5.9.3) on Q E and the connection F (5.9.12) on E R. The
Hamiltonian function H satisfies the conditions
n
i
(t,q
i
,a
m
,d
i
H(t,a
m
,q
i
,n
i
))=7T
h
Pid'n-n^ c^ q^ ^ ^ d'n)
5.9. SYSTEMS WITH TIME-DEPENDENT PARAMETERS 111
which axe obtained by substitution of the expression (5.9.13) in the conditions
(5.5.2a) - (5.5.2b).
The Hamilton equations for the Hamiltonian form (5.9.13) read
ql = A' + A^ r + dm (5.9.14a)
Pu = -Pj{diA
j
+^ r ) - W, (5.9.14b)
_m run
(5.9.14c)
Pt
m
= -Pi(d
m
A
{
+r
n
a
m
4) - d
m
n, (5.9.14d)
whereas the Lagrangian constraint space is given by the equations
p
i
= TT
i
(t,q\(T
m
,d
i
H(t,CT
m
,q\p
i
)), (5.9.15)
p
m
+ AIJH = 0. (5.9.16)
The system of equations (5.9.14a) - (5.9.16) are related in the sense of Section
4.5 to the Lagrange equations (5.9.7) - (5.9.8). Because of the equations (5.9.14d)
and (5.9.16), this system is overdetermined. Therefore, the Hamilton equations
(5.9.14a) - (5.9.14d) admit a solution living in the Lagrangian constraint space
(5.9.15) - (5.9.16) if the very particular condition, similar to the condition (5.9.9),
holds. Since the Lagrange equations (5.9.8) axe not considered, we also can ignore
the equation (5.9.14d). Note that the equations (5.9.14d) and (5.9.16) determine
only the momenta p
m
and do not influence other equations.
Therefore, let us consider the system of equations (5.9.14a) - (5.9.14c) and
(5.9.15) - (5.9.16). Let the connection T in the equation (5.9.14c) be complete and
admit the integral section h(t). This equation together with the equation (5.9.16)
defines a submanifold V'Qh of the momentum phase space V*Q, which is the mo-
mentum phase space of a mechanical system with the parameter function h(t). The
remaining equations (5.9.14a) - (5.9.14b) and (5.9.15) are the equations of this
system on the momentum phase space V*Qh, which correspond to the Lagrange
equations (5.9.7) in the presence of the background parameter function h(t).
Conversely, whenever h(t) is a parameter function, there exists a connection T
on the parameter bundle E R such that h(t) is an integral section T. Then
the system of equations (5.9.14a) - (5.9.14b) and (5.9.15) - (5.9.16) describes a
mechanical system with the background parameter function h(t). Moreover, we can
locally restrict our consideration to the equations (5.9.14a) - (5.9.14b) and (5.9.15).
The following examples illustrate the above construction.
278 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
Exampl e 5.9.1. Let us consider the 1-dimensional motion of a probe particle in
the presence of a force field whose centre moves. The configuration space of this
system is the composite fibre bundle
M
3
->M
2
->E,
coordinated by (t, a, q) where a is a coordinate of the field centre with respect to
some inertial frame and q is a coordinate of the probe particle with respect to the
field centre. There is the natural inclusion
QxTY,B {t,a,q,i,a) ^ {t,o,i,a,y = -a) eTQ
which defines the connection
A = dt d
t
+ da <g>(d - d
q
)
on the fibre bundle Q >E. The corresponding vertical covariant differential (5.9.4)
reads
D = (q
t
+ a
t
)d.
This is precisely the velocity of the probe particle with respect to the inertial frame.
Then the Lagrangian of this particle takes the form
L=[^(q
t
+ a
t
)
2
-V(q)}dt. (5.9.17)
In particular, we can obtain the energy conservation law for this system. Let us
consider the vector field u = d
t
. The Lie derivative of the Lagrangian (5.9.17) along
this vector field vanishes. Using the formula (5.9.11), we obtain
0 TT
q
(7
u
dt[ir
q
q
t
],
or
0 ~d
q
Cat ~ dt[ir
q
{qt + o
t
) - ],
where
X =7r,(ft+ <r
t
)-
is the energy function of the probe particle with respect to the inertial reference
frame.
5.9. SYSTEMS WITH TIME-DEPENDENT PARAMETERS 279
Exa mpl e 5. 9. 2. Let us consider the 1-dimensional oscillator whose frequency is
a time-dependent parameter. In accordance with the adiabatic hypothesis, there
exists a coordinate q such that this oscillator differs from the standard one only in
kinetic energy.
Let us take the configuration space
Q = R x R
+
x R - R
+
x R - +R ,
provided with the coordinates (t, a > 0, q), where a is the frequency parameter. The
corresponding momentum phase space V*Q is coordinated by {t,cr,q,p
a
,p
g
). The
oscillator is characterized by the Hamiltonian form
H = p
a
da + p
q
dy Hdt,
H=p
a
T + \{oy
q
+ q\ (5.9.18)
The corresponding Hamiltonian frame connection
T
H
= d
t
+ T
t
d
a
(5.9.19)
on the configuration bundle Q R is the composite connection (1.6.8) generated
by some linear connection
r =dt (d
t
+ T(t)od
a
)
on the parameter bundle
S = K
+
x l - . R
and by the trivial connection
A = dt d
t
+ da d
a
on the fibre bundle Q > E.
The Hamiltonian form (5.9.18) is associated with the Lagrangian
L=\{o-
2
c?
t
-qi)dt
which describes the oscillator with time-dependent frequency a = h{t) in accordance
with the adiabatic hypothesis.
280 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
The Hamilton equations (5.9.14a) - (5.9.14c) for the Hamiltonian form (5.9.18)
read
qt = ^
2
P
(5.9.20a)
Pt = -9,
(5.9.20b)
c
t
= T{t)a.
(5.9.20c)
They also describe the oscillator with the variable frequency a = h{t) if we put
r =h~
l
d
t
h. The Lagrangian constraint space equations (5.9.15) are trivial.
Let us consider the canonical transformation
q = aq',
1 ,
, 1 i ,
Po=V
a
QPq-
a
Relative to the new coordinates (t,a, ^.p'^.p',), the connection (5.9.19) is written
r
H
= d
t
+ T{d
a
- q'd
q
,),
while the Hamiltonian form (5.9.18) reads
H =p'
a
da + p'
q
dq' - Hdt,
n = p^-p'
qq
'T + \(pf + a^).
Accordingly, the Hamilton equations (5.9.20a) - (5.9.20c) take the form
Pty = \PT - -V,
(5.9.21a)
q'
t
= -
q
>T+p'
q
, (5.9.21b)
<>t = 17-
(5.9.21c)
Substitution of a = h(t) and the equation (5.9.21c) in the equations (5.9.21a) and
(5.9.21b) leads to the Hamilton equations corresponding to the Hamiltonian form
H =p'
q
dq' - Hdt,
H = -i<Wp;+^(p' ,
2
+*v
2
).
5.9. SYSTEMS WITH TIME-DEPENDENT PARAMETERS 281
This Hamiltonian form describes the well-known Berry oscillator where the connec-
tion (5.9.5) is the classical variant of the Berry connection
A
h
= d
t
d
t
aq'd
q/
a
[135, 144].
Exa mpl e 5.9.3. Let us consider an n-body as in Example 4.9.4. Let R
3

-3
be the
translation-reduced configuration space of the mass-weight J acobi vectors {p
A
}-
Two configurations {p
A
} and {p
A
} are said to define the same shape of the n-
body if p
A
= Rp
A
for some rotation R 50(3). This introduces the equivalence
relation between configurations, and the shape space S of the n-body is defined as
the quotient
S = R
3n
-
3
/50(3)
[119, 120]. Then we have the composite fibre bundle
1 x R
3n_3
- R x S - R, (5.9.22)
where the fibre bundle
K
3
""
3
_>S
(5.9.23)
has the structure group 50(3). The composite fibre bundle (5.9.22) is provided with
the bundle coordinates(t, a
m
,q
l
), where q', i = 1,2,3, are some angle coordinates,
e.g., the Eulerian angles, while a
m
, m = 1,... ,3n 6, are said to be the shape
coordinates on S. A section Q
A
{o
m
) of the fibre bundle (5.9.23), called a gauge
convention, determines an orientation of the n-body in a space. Given such a section,
any point {p
a
} of the translation-reduced configuration space R
3

-3
is written as
PA = R(qTQ
A
{
m
)-
This relation yields the splitting
~p
A
=diRcfQji + d
m
~e
A
d
m
of the tangent bundle TR
3n 3
, which determines a connection A on the fibre bundle
(5.9.23). This is also a connection on the fibre bundle
R x R
3
"
- 3
- R x 5,
282 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
with the components A
t
= 0. Then the corresponding vertical covariant differential
(5.9.4) reads
D = (4 - /&?)*.
With this vertical covariant differential, the total angular velocity of the n-body
takes the form
0 =a^D* = u +A
m
<77\
(5.9.24)
where &i are certain kinematic coefficients and 3 is the angular velocity of the n-
body as a rigid one. In particular, we obtain the phenomenon of a faffing cat if
0 =0 so that
u = -A
m
<r
t
m
The coefficients A
m
in the expression (5.9.24) are called gauge potentials, and the
mechanics of deformable bodies is sometimes termed gauge mechanics.
5.10 Unified Lagrangian and Hamiltonian formalism
The relationship between Lagrangian and Hamiltonian formalisms in Section 5.5 is
broken by canonical transformations if the transition functions q' <r" depend on
momenta. The following construction enables us to overcome this difficulty.
Given a configuration bundle Q M, let V*J
l
Q be the vertical cotangent bundle
of the velocity phase space J
l
Q R, with coordinates
(*,?',?!, ft,?'),
and let J
X
V*Q be the jet manifold of the phase space V*Q M, with coordinates
{t,q\Pi,q\,Pti)-
PROPOSI TI ON 5.10.1. There is the isomorphism
n = V* J
X
Q & J
l
V"Q,
ft *> Pu, 4t < P i ,
(5.10.1)
over J
l
Q.
5.10. UNIFIED LAGRANGIAN AND HAMILTONIAN FORMALISM 283
Pr oof. The isomorphism (5.10.1) can be proved by an inspection of transition
functions of the coordinates {%,$) and (puPti)- QED
Due to the isomorphism (5.10.1), one can think of I I as being both the momen-
tum phase space over the velocity phase space J
l
Q and the velocity phase space
over the momentum phase space V*Q. Hence, the space I I can be utilized as the
unified phase space of joint Lagrangian and Hamiltonian formalism.
Rema r k 5. 10. 1. In connection with this, note that, according to [10, 127, 180], the
dynamics of autonomous mechanical system described by a degenerate Lagrangian
L on TM is governed by a differential equation on T*M generated by dL, due to
the canonical diffeomorphism between TT'M and T*TM.
The unified phase space I I is equipped with the holonomic coordinates
{t,q\q
l
t,PtuPi),
(5.10.2)
where {q',pu) and {q\,Pi) are canonically conjugate pairs. It is endowed with the jet
prolongation of the canonical 3-form (4.8.10), which with respect to the coordinates
(5.10.2) reads
ft = (dpti A dq' + dpi A dq\) A dt = dt{dpi A dq' A dt), (5.10.3)
where
dt = d
t
+ q\di + p
ti
d'
is the total derivative on I I . The corresponding Poisson bracket (5.1.5) takes the
form
{/,ff}n
df_ 0 df_ dg_ _ dg_dl _ dg_dl
dp
ti
dq
i
dpidqt dpudq* dpidq\
(5.10.4)
It is readily observed that the canonical form (5.10.3) and, consequently, the
Poisson bracket (5.10.4) are invariant under the transformations of the unified phase
space I I , which are the jet prolongation of the canonical automorphisms of the phase
space V*Q.
Let
H Ptidq'+ Pidq\ - H
u
(t,q',q\,p
u
,Pi)dt
284
CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
be a Hamiltonian form on I I . The corresponding Hamilton equations (5.2.18a)
(5.2.18b) read
OPti
(5.10.5a)
, , dHn
dpi
(5.10.5b)
, dHn
(5.10.5c)
dHn
(5.10.5d)
Substitution of (5.10.5a) in (5.10.5b) and of (5.10.5c) in (5.10.5d) leads to the equa-
tions
mn
=
&Hn
dpu dpi '
(5.10.6a)
&Hn _ dHn
* dq\ " dq'
(5.10.6b)
which look like the Lagrange equations for the "Lagrangian" Tin- Though Hn is
not a true Lagrangian, one can put
H = -C + dt(
Pi
r), (5.10.7)
whenever L is a Lagrangian on the velocity phase space J Q. Then the equations
(5.10.6a) - (5.10.6b) are equivalent to the Lagrange equations for a Lagrangian L
on J
l
Q. However, their solutions fail to be solutions of the corresponding Ha.mil-
ton equations (5.10.5a) - (5.10.5d) for the Hamiltonian (5.10.7) in general. This
illustrates the fact that solutions of the Hamilton equations (5.10.5a) - (5.10.5d)
are necessarily solutions of the Lagrange equations (5.10.6a) - (5.10.6b), but the
converse statement is not true.
To construct the joint Lagrangian-Hamiltonian formalism, let us consider the
Hamiltonian form
H = pudg' +Pidgj - (dtH +{ml -H)- C)dt (5.10.8)
on the unified phase space I I , where L is a semiregular Lagrangian on the velocity
phase space J
l
Q and H is an associated Hamiltonian form on the momentum phase
5.11. VERTICAL EXTENSION OF HAMILTONIAN FORMALISM 285
space V*Q. The Hamilton equations (5.10.5a) - (5.10.5d) for H (5.10.8) read
<w
&H
dpi'
(5.10.9a)
dtq]
2
m
OPi
on
+
dpi
(5.10.9b)
dtPi
dq*
dC
(5.10.9c)
dtPu -dt
an an dc
dq
i +
dq
i +
dtf
(5.10.9d)
Using the relations (5.5.2a) and (5.5.7), one can show that solutions of the Hamil-
ton equations (5.2.18a) - (5.2.18b) for the Hamiltonian form H, which live in the
Lagrangian constraint space L^Q) C V*Q are solutions of the equations (5.10.9a)
- (5.10.9d).
Now let us consider the Lagrange equations (5.10.6a) - (5.10.6b) for the Hamil-
tonian form (5.10.8). They read
<k<t
dpi
o,
(5.10.10a)
dtPi
_dU _
dqi
d
dq
r
(5.10.10b)
In accordance with Proposition 5.5.7, every solution of the Lagrange equations for
the Lagrangian L such that the relation (5.5.22) holds is a solution of the equations
(5.10.10a) - (5.10.10b).
In particular, if the Lagrangian L is hyperregular, the equations (5.10.10a) -
(5.10.10b) and the equations (5.10.9a) - (5.10.9d) are equivalent to the Lagrange
equations for L and the Hamilton equations for the associated Hamiltonian form.
5.11 Vertical extension of Hamiltonian formalism
Let Q R, coordinated by {t,q
l
), be a configuration space of time-dependent
mechanics. Let us consider the vertical tangent bundle VQ of the fibre bundle
Q >R, coordinated by (t,q',q'). It can be seen as the new configuration space,
called the vertical configuration space. Here, we will be concerned with the following
three applications of Hamiltonian mechanics on this configuration space:
mechanical systems with linear deviations (see Example 5.11.1 below),
286 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
Hamiltonian description of non-holonomic constraints (see Example 5.11-3 be-
low),
Hamiltonian BRST mechanics (see Appendix A).
Given VQ, the corresponding velocity phase space is the first order jet manifold
J
l
VQ of the fibre bundle VQ - R. It is canonically isomorphic (see (1.3.8)) to
the vertical tangent bundle VJ
l
Q of the ordinary velocity phase space J
X
Q - M,
coordinated by
{%<?,$ AA)-
We call VJ
l
Q the vertical velocity phase space.
The Legendre bundle of the vertical configuration space VQ is the vertical cotan-
gent bundle V'VQ of the fibre bundle VQ >M. It plays the role of a momentum
phase space, called the vertical momentum phase space, of mechanical systems on
the configuration space VQ. The vertical momentum phase space is canonically
isomorphic (see (1.1.10)) to the vertical tangent bundle VV'Q of the ordinary mo-
mentum phase space V*Q >K, coordinated by
(t,q',Pi,q',Pi).
It is easily seen from the transformation laws
Pi = djlPv
., dp\ . dp[ dq> . dtf
that (q\fii) and {q',Pi) are canonically conjugate pairs.
Thus, we have observed that the vertical velocity and momentum phase spaces,
corresponding to the configuration space VQ, are the vertical tangent bundles of
ordinary velocity and momentum phase spaces, respectively.
The momentum phase space VV'Q is endowed with the canonical 3-form
Qv = [dpi A dq
l
+ dpi A dq^] A dt.
(5.11.1)
For brevity, one can write
flv = dyl,
5.11. VERTICAL EXTENSION OF HAMILTONIAN FORMALISM 287
where fi is the canonical 3-form (5.1.9) on the momentum phase space V*Q, and
the formal operator of the vertical derivative
dy =q'di + pjd'
(1.1.4) is utilized. This operator possesses the properties
dv(4> ACT) =dv(<f>) ACT +</)A9V((T), ^ae&iV'Q),
dv{dcj>) = d{dv{cj>)).
The canonical 3-form (5.11.1) provides VVQ with the canonical Poisson struc-
ture, given by the Poisson bracket
{/, g}vv = frfdjg +d'fdjg - VgdJ - frgdj.
Recall the notation (1.2.4).
The notions of a Hamiltonian vector field, a Hamiltonian connection, a Hamil-
tonian form and others on the vertical momentum phase space VV*Q = V*VQ are
introduced in accordance with the general formalism of Hamiltonian time-dependent
mechanics, expressed in Section 5.2. In particular, any Hamiltonian form on VV*Q
reads
Hy =Pidq
1
+ Pidq' Hydt.
Since a Hamiltonian form is defined modulo exact forms and the function p{<? is
globally defined on VV*Q, we will write a Hamiltonian form on the momentum
phase space VVQ as
Hy =pidq
1
cfdpi Ttydt. (5.11.2)
The corresponding Hamilton equations read
V =ql = d'Hv, (5.11.3a)
li =Pti = -diH
v
,
(5.11.3b)
f =4 = d'Hv,
(5.11.3c)
7i =Pti = -diHv,
(5.11.3d)
where
7 =d
t
+ J% + "fid* + j % + jid*
288
CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
is a Hamiltonian connection on the momentum phase space VV'Q > R.
We have the following relationship between Hamiltonian formalism on the mo-
mentum phase space V'Q and that on VV'Q.
PROPOSI TI ON 5.11.1. Let 7#be a Hamiltonian connection on the ordinary mo-
mentum phase space V*Q R for a Hamiltonian form
H = pidq* - Hdt.
(5.11.4)
Then the connection Vj
H
(1.6.12) on the vertical momentum phase space VV'Q
R is a Hamiltonian connection for the Hamiltonian form
H
v
= Ptdq' q'dpi dyTidt, (5.11.5)
d
v
H = (q% + pj&yH,
called the vertical extension of the Hamiltonian form H. D
Proof. A direct computation shows that, given a Hamiltonian connection
lH = d
t
+ 'y
i
d
i
+ 'y
i
d',
f =d'H,
7
= -diH,
on the fibre bundle V'Q >R, the connection
V-TH = d
t
+ j'di +
7i
d
l
+ fdi + j
{
d\
7
1
=dv-y', 7i = dvlu
on the fibre bundle VV'Q - R satisfies the Hamilton equations (5.11.3a) - (5.11.3d)
for the Hamiltonian form (5.11.5), i.e.,
y =titty =d'H,
(5.11.6a)
7i =-mv = -diH,
(5.11.6b)
f =d'Hv = dvd
l
n,
(5.11.6c)
7i = -diH
v
= -dvdiH.
(5.11.6d)
QED
In particular, given a splitting
H = p,-r + H
T
5.11. VERTICAL EXTENSION OF HAMILTONIAN FORMALISM 289
of a Hamiltonian Tt with respect to a connection Y on Q >R, there is the corre-
sponding splitting
Hv =ftP - tfi-pidjF) + dvH
T
of the Hamiltonian Hv (5.11.5) with respect to the connection T (5.2.6) on the fibre
bundle V*Q -^R.
It is easily seen that the Hamilton equations (5.11.6a) - (5.11.6b) for the Hamil-
tonian form Hv (5.11.5) are precisely the Hamilton equations (5.2.18a) - (5.2.18b)
for the Hamiltonian form H.
Exampl e 5.11.1. In order to clarify the physical meaning of the Hamilton equations
(5.11.6c) - (5.11.6d), let us suppose that Q >R is a vector bundle. Due to the
splitting
VQ^QQ,
we can think of the vertical configuration space VQ as the configuration space of
the initial mechanical system and its small linear deviations
3* + eg*, R,
where is a small parameter. Accordingly, the vertical momentum phase space
VVQ = VQVQ
can be seen in a similar way as the momentum phase space of the initial Hamiltonian
mechanical system and its small linear deviations
q
i
+ eq\ Pi + epi-
Let if be a Hamiltonian form on the momentum phase space V*Q of the initial
Hamiltonian system, and let r{t) be a solution of the Hamilton equations (5.11.6a)
- (5.11.6b). Let r() be a Jacobi field, i.e.,
r(t) +ef(t)
is also a solution of the same Hamilton equations modulo terms of order two in e.
Then it is readily observed that the J acobi field f(t) satisfies the Hamilton equa-
tions (5.11.6c) - (5.11.6d) for the Hamiltonian form H
v
(5.11.5). Substituting a
solution r(t) of the Hamilton equations (5.11.6a) - (5.11.6b) in the Hamilton equa-
tions (5.11.6c) - (5.11.6d), we obtain the system of homogeneous linear differential
290 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
equations for the J acobi field F(t), which are actually the differential equations of
linear deviations.
Returning to the definition of a Hamiltonian form H
v
(5.11.5), let us consider
the vertical tangent bundle VT*Q of the cotangent bundle T*Q - E. It is equipped
with holonomic coordinates
{t,q',Pi,P,q',Pi,p)
and endowed with the canonical form
Ey = pdt + pidq
1
q
l
dpi.
One may also utilize another canonical form
E
v
+ d{q'pi)
since the exact form d(<j*p,*) is globally defined. By very definition, a Hamiltonian
formH (5.11.4) on the ordinary momentum phase space V*Q is the pull-back H =
h'E of the Liouville formE (5.1.2) on the cotangent bundle T*Q by a section h of
the fibre bundle T'Q -> V'Q. Then we have
H
v
= (VhYEv,
where
Vh : VVQ - VT'Q
is the vertical tangent map to h.
We can also generalize Proposition 5.11.1 as follows.
PROPOSI TI ON 5.11.2. Any connection 7 on the momentum phase space V'Q
M gives rise to a Hamiltonian connection on the vertical momentum phase space
VVQ - >R. D
Proof. Let us consider the Hamiltonian form
H
v
= pijdq
1
- fdt) - ^(dpi - -gdi) = p
i
dg
i
- tfdpj - (pg* - tf-y^dt
on the vertical momentum phase space VV*Q. The corresponding Hamiltonian
connection on VVQ > R is given by the Hamilton equations (5.11.3a) - (5.11.3d)
and reads
f = 7\ 7i = 7t,
V-fcay-a'aS,
7i = -Pj -8i 7
J
+?
3
9i 7j .
(5.11.7)
5.11. VERTICAL EXTENSION OF HAMILTONIAN FORMALISM 291
In particular, if 7 is a locally Hamiltonian connection on the fibre bundle V*Q >
R, i.e., satisfies the relations (5.2.4a) - (5.2.4c), then the Hamiltonian connection
(5.11.7) is precisely the connection V7 (1.6.12). QED
It follows that every first order dynamic equation (5.2.20) on the momentum
phase space VQ can be seen as the Hamilton equations (5.11.3a) - (5.11.3b) for a
suitable Hamiltonian form on the vertical momentum phase space.
Exa mple 5.11.2. The equations of a motion (5.2.21) of a point mass m subject to
friction in Example 5.2.5 are the Hamilton equations (5.11.3a) - (5.11.3b) for the
Hamiltonian
Wv =p{p-\ q)
m m
on the vertical momentum phase space.
Similarly to the relationship between Hamiltonian formalisms on ordinary and
vertical momentum phase spaces, that between Lagrangian formalisms on ordinary
and vertical velocity phase spaces can also be considered.
Given a Lagrangian L = Cdt (4.8.1) on an ordinary velocity phase space J
l
Q,
let us consider the vertical tangent morphism
VC : J
l
Q -* R x R
to the Lagrangian function C, and the Lagrangian
Ly = (pr
2
V)dt = Cydt, (5.11.8)
Cy=dyC= (fdj + q\d\)C,
on the vertical velocity phase space VJ
l
Q. It is called the vertical extension of the
Lagrangian L. The corresponding Lagrange equations read
(4 - dtdt)c
v
= 0, (5.11.9a)
{d, - dtdt)
v
=0.
(5.11.9b)
The Lagrange equations (5.11.9a) are precisely the Lagrange equations for the initial
Lagrangian L. If a configuration space Q >M is a vector bundle, the Lagrange
equations (5.11.9b), just as the Hamilton equations (5.11.6c) - (5.11.6d) do, describe
small linear deviations of the initial mechanical system.
292 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
Let us turn now to the relationship between Lagrangian and Hamiltonian for-
malisms on the vertical configuration space.
Each Lagrangian L
v
(5.11.8) on the vertical velocity phase space VJ
l
Q yields
the Legendre map
L
V
= VL:VJ
1
Q ^VV'Q,
VQ
(5.11.10)
Pi = d\C
V
= TTi, Pi = dy^i,
(5.11.11)
which is the vertical tangent map to the Legendre map L.
Each Hamiltonian form H
v
(5.11.5) on the vertical momentum phase space
VV*Q yields the Hamiltonian map
H
V
= VH : VV'Q -> VJ
X
Q,
VQ
(5.11.12)
q
i = ffnv = am,
q\ = dvd'H,
which is the vertical tangent map to the Hamiltonian map H.
PROPOSI TI ON 5.11.3. If a Hamiltonian formH on the momentum phase space VQ
is associated with a Lagrangian L on the velocity phase space J
l
Q, then its vertical
extension Hv (5.11.5) on the vertical momentum phase space VV'Q is associated
with the vertical extension L
v
(5.11.8) of the Lagrangian L on the vertical velocity
phase space V J
X
Q. D
Pr oof. By virtue of the relations (5.11.10) and (5.11.12), the condition (5.5.2a) for
the Hamiltonian formHv reads
VL o VH o VL = VL,
and follows at once from the condition (5.5.2a) for the Hamiltonian form H. The
condition (5.5.2b) for the Hamiltonian form H
v
, written in the coordinate form
(5.5.5), follows immediately from the condition (5.5.5) for the Hamiltonian form H.
QED
The equalities (5.11.11) show that, if a Lagrangian L on the velocity phase space
J
l
Q is semiregular, so is its vertical extension L
v
(5.11.8) on the vertical velocity
phase space VJ
l
Q. Therefore, Proposition 5.11.3 can also be formulated for a
semiregular Lagrangian. In particular, if the relation (5.5.16) is fulfilled for H and
L, it is true for the vertical extensions H
v
and Ly-
5.11. VERTICAL EXTENSION OF HAMILTONIAN FORMALISM 293
Exa mpl e 5. 11. 3. Let us consider the Hamiltonian counterpart of the constrained
motion equation (4.11.15) in the case of a hyperregular Lagrangian and a non-
holonomic constraint. By virtue of Proposition 5.11.2, it can be represented as the
Hamilton equations (5.11.3a) - (5.11.3b) for a suitable Hamiltonian form on the
vertical phase space VV*Q.
Let L be a hyperregular Lagrangian with a Riemannian mass metric m and
S a codistribution on the velocity phase space J
X
Q, whose annihilator Ann (S) is
an admissible non-holonomic constraint on J
l
Q. In the case of a hyperregular
Lagrangian, Hamiltonian and Lagrangian formalisms of time-dependent mechanics
are equivalent, and we have
7H = J ' i o^of f .
Let introduce the notation M
,J
=d'&>H. There are the relations
M
ik
(m
kj
o H) =8}, m
kj
(M
ik
o I) = 6), TTlij = Tty.
It follows that M is a fibre metric in the vertical tangent bundle VQV'Q of the fibre
bundle VQ ~* Q.
Given the above-mentioned codistribution S on J
X
Q, let us consider the pull-back
codistribution H*S on V*Q spanned locally by the 1-forms
/? = H*s
a
= (sg +s^dtd>n)dt + (s? +s'di&HW + sM
ij
d
Pj
=
ffidt + (3fdq' + $
ai
d
Pi
.
This codistribution defines a non-holonomic constraint on the momentum phase
space V*Q.
Given a Hamiltonian connection 7// (5.2.17), let us find its splitting
1H = 7 +0
(5.11.13)
where 7 is a connection on V*Q R which satisfies the condition
7C Ann(i TS). (5.11.14)
The connection 7 (5.11.14) obviously defines a first order dynamic equation on the
momentum phase space V*Q, which is compatible with the non-holonomic con-
straint H*S. The decomposition (5.11.13) is not unique. Let us construct it as
follows.
294 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
Given a Hamiltonian connection 7//, we consider the codistribution SH on V'Q,
spanned locally by the 1-forms dq' - i\jdt. Its annihilator Ann (S
H
) is an affine
subbundle of the affine jet bundle J
l
V'Q ->V'Q modelled over the vertical tan-
gent bundle VQV'Q. The Hamiltonian connection j
H
is obviously a section of this
subbundle. Let us consider the intersection
W = Ann(S
H
)nAnn(#*S).
(5.11.15)
LEMMA 5.11.4. The intersection W (5.11.15) is an affine bundle over V'Q, modelled
over the vector bundle
W = V
Q
V*Q nAnn(H'S).
a
Pr oof. The intersection W consists of elements v of VQV'Q which fulfill the con-
ditions
v
{
p
a
' = 0.
Since the non-holonomic constraint S is admissible and the matrix M
%i
is non-
degenerate, every fibre of W is of dimension m n, i.e., W is a vector bundle, while
W is an affine bundle. QED
Then, using the fibre metric M in VQV'Q, we obtain the splitting
VQV'Q = W e V,
where V is the orthocomplement of W, and the associated splitting
Ann (SH) = W V.
The corresponding decomposition (5.11.13) reads
7 = lH-M
ab
M
tJ
p
ai
p
b
(
lH
)d>,
(5.11.16)
where M
a
t, is the inverse matrix of
M
ab
=
fripMMij.
5.11. VERTICAL EXTENSION OF HAMILTONIAN FORMALISM 295
The splitting (5.11.16) is the Hamiltonian counterpart of the splitting (4.11.15).
We have the relations
~ab
= M
ab
Q
^
/3"(7tf) =s
Q
(a) H,
and as a consequence
7 = J ^of of f .
Note that the above procedure can be extended in a straightforward manner to
any standard Newtonian system seen as a Lagrangian system with the Lagrangian
(4.9.19) and an external force. Following this procedure, one may also study a non-
holonomic Hamiltonian system, without appealing to its Lagrangian counterpart.
The connection (5.11.16) defines the system of first order dynamic equations
{= d*n, Pu = -W - M
o6
M
y
^/3(
7f f
)
(5.11.17)
on the momentum phase space V*Q, which are not Hamilton equations. Neverthe-
less, in accordance with Proposition 5.11.2, one can restate the constrained motion
equations (5.11.17) as the Hamilton equations (5.11.6a) - (5.11.6b) for the Hamil-
tonian form
H
v
= fad? - tfdpi - dvHdt - q
i
M
ab
M
ij
$
a:i
(3
b
('y
H
)dt
on the vertical momentum phase space VV'Q. Its last term can be written in brief
as (<9v/J?9J$7). To exclude the remaining Hamilton equations (5.11.6c) - (5.11.6d),
one can choose their solutions q' = pi = 0.
Rema r k 5.11.4. Let Hy be a Hamiltonian form (5.11.2) on the vertical momen-
tum phase space VV'Q. Similarly to the Lagrangian (5.7.1), let us consider the
Lagrangian
LH = Piq\ - q
l
Pu - 'Hv
(5.11.18)
on the first order jet manifold J
l
VV*Q of the fibre bundle VV*Q > M. It is readily
observed that its Lagrange equations are precisely the Hamilton equations (5.11.3a)
- (5.11.3d) for the Hamiltonian form Hy- In particular, let H be a Hamiltonian
form on an ordinary momentum phase space VQ and "Hv = dy'H- In this case, the
Lagrangian (5.11.18) reads
L
H
=Pi(qi-d
l
H)-q
i
(
Pti
+ d
i
H).
296 CHAPTER 5. HAMILTONIAN TIME-DEPENDENT MECHANICS
It is easily seen that this Lagrangian vanishes on solutions of the Hamilton equations
for the Hamiltonian form H. For this reason, it plays a prominent role for the
functional integral formulation of mechanics [65, 66]
5.12 Appendix. Time-reparametrized mechanics
We have assumed above that the base M of a configuration space of time-dependent
mechanics is parameterized by the standard coordinate t with the transition func-
tions t > t' =t+ const. Here, we consider an arbitrary reparametrization of time
t - a = m
(5.12.1)
which is discussed in some models of quantum mechanics [75, 153, 154]. In this case,
K is not only an affine space as before, but a 1-dimensional manifold. Then the stan-
dard vector field d
t
and the standard 1-formdt on R are not invariant. Therefore,
we should follow the polysymplectic Hamiltonian formalism of Section 3.8. Nev-
ertheless, Hamiltonian formalism of time-reparametrized mechanics possesses some
peculiarities due to R.
1. There exists the invariant tangent-valued 1-form
e
R
= dt dt
on the base R of a configuration space of time-reparametrized mechanics.
2. The velocity phase space J
1
Q of time-reparametrized mechanics is no more
an affine subbundle of the tangent bundle TQ under time reparametrizations. Nev-
ertheless, the momentum phase space U > Q (3.8.1) is isomorphic to the vertical
cotangent bundle V*Q >Q. It follows that the momentum phase space of time-
reparametrized mechanics is provided with the canonical Poisson structure (5.1.5).
Moreover, this Poisson structure is invariant under time reparametrization (5.12.1)
which, consequently, is a canonical transformation.
3. The momentum phase space V*Q is endowed with the canonical polysym-
plectic form
A =dpi A dq' Adt d
t
.
Then the notions of a Hamiltonian connection and a Hamiltonian form are the rep-
etitions of those in Section 3.8. At the same time, since the homogeneous Legendre
5.12. APPENDIX. TIME-REPARAMETRIZED MECHANICS 297
bundle (2.9.3) of time-reparametrized mechanics is the cotangent bundle T*Q of Q,
Hamiltonian forms and Hamilton equations of time-reparametrized mechanics have
the same form as those in time-dependent mechanics. The difference is only that
a Hamiltonian function Tir in the splitting (5.2.12) is a density, but not a function
under the transformations (5.12.1).
4. Since Lagrangians and Hamiltonians of time-reparametrized mechanics are
densities under the transformations (5.12.1), one should introduce a volume ele-
ment on the base M in order to construct them in explicit form. A key problem
of models with time reparametrization lies in the fact that the volume dt on the
time axis M is not invariant under time reparametrization transformations. Another
problem is concerned with mass tensors which also fail to be invariant under time
reparametrizations.
Chapter 6
Relativistic mechanics
Let us note on the following two main peculiarities of relativistic mechanics in com-
parison with non-relativistic one.
A configuration space Q of relativistic mechanics has no fibration Q >R
in general or, at least, it has no preferable fibration over K. Therefore, one
cannot use the familiar formalism of jets of sections of fibre bundles, but must
generalize it for jets of submanifolds.
Hamiltonian relativistic mechanics is described as an autonomous constraint
Dirac system on the hyperboloids of relativistic momenta in the momentum
phase space T*Q.
Relativistic mechanics can be formulated on an arbitrary configuration space Q.
If Q is a 4-dimensional manifold provided with a pseudo-Riemannian metric g, we
come to relativistic mechanics in the presence of a gravitational field g. If Q = R
4
and g is the Minkowski metric, this is the case of Special Relativity.
6.1 J ets of submanifolds
We start from the general notion of jets of submanifolds [57, 104].
DEFI NI TI ON 6.1.1. Let Z be a manifold of dimension m + n. The first order
jet manifold J\Z of n-dimensional submanifolds of a manifold Z comprises the
equivalence classes [S]
l
z
, z Z, of n-dimensional imbedded submanifolds of Z which
299
300 CHAPTER 6. RELATIVISTIC MECHANICS
pass through z Z and which are tangent to each other at z. There is the natural
fibration J\Z ->Z. D
Remark 6.1.1. In fact, the definition of the jet [S}1 of submanifolds involves only
local properties of submanifolds around the point z Z. It is easily extended to
higher order jets of submanifolds. By definition, JZ = Z.
The set of jets J\Z is provided with a manifold structure as follows.
Let Y >X be an (m +n)-dimensional fibre bundle over an n-dimensional base
X and let $ be an imbedding of Y into Z. Then there is the natural injection of
the jet manifold J
l
Y of the fibre bundle Y > X into J\Z such that
7
1
* : J
l
Y -> J\Z, (6.1.1)
J l
S
- [Sjsfgfa;))!
S = I m($os),
where s are sections of Y X.
PROPOSI TI ON 6.1.2. The injection (6.1.1) defines a chart on J ^Z. Such charts
cover the set J^Z, and transition functions between these charts are differentiable.
They provide the set J\Z with the structure of a finite-dimensional manifold.
Proof. The proof is based on the fact that, given a submanifold S C Z which
belongs to the jet [S\l, there exists a neighbourhood U
z
of the point z and the
tubular neighbourhood Us of U
z
D S so that the fibration
u
a
->u,ns
takes place [109]. This means that every jet [S]l belongs to a chart of the above-
mentioned type. We will describe these charts and the corresponding transition
functions in explicit form. QED
It is convenient to use the following coordinate atlases of the jet manifold J\Z
of n-dimensional submanifolds of Z.
Let Z be equipped with a coordinate atlas
Mz*)},
A= 1,... ,n + m.
(6.1.2)
6.1.
JETS OF SUBMANIFOLDS 301
Though JZ, by definition, is diffeomorphic to Z, let us provide JZ with the atlas
where every chart (U; z
A
) on a domain U C Z is replaced with the
(
n + m\ _ (n + m)\
m J n\m\
charts on the same domain U which correspond to different partitions of the collec-
tion (z
1
z
A
) in the collections of n and m coordinates, denoted by
(zV),
A= 1, ,..,!, i = 1,... ,m. (6.1.3)
The transition functions between the coordinate charts (6.1.3) of JZ, associated
with the same coordinate chart (6.1.2) of Z, are reduced simply to an exchange
between coordinates x
x
and y
l
. Transition functions between arbitrary coordinate
charts (6.1.3) of the manifold JZ read
2
A
=ffV*,tf0.
y
i
= f
i
(x" ,y
j
),
(6.1.4)
x* = g*{x?,yl)
1
f = fi^y
3
).
Given the coordinate atlas (6.1.3) of the manifold JZ, the jet manifold J\Z is
provided with the adapted coordinates
(x\y\yi),
X = l , . . . , n, i = 1, . . . , ? i . (6.1.5)
Remark 6.1.2. If S C Z is an imbedded submanifold which belongs to the jet
[S]l, there exists an open neighbourhood U
z
of z Z and a coordinate chart (x
x
,y
l
)
(6.1.3) which covers U
z
so that SfMJ
z
\s given by the coordinate relations
y
{
= 5
{
(x
A
)
and
(x\y\yi)([S]l) = (x
x
(z),S
i
(^(z)),d
x
S
i
(x"(z))).

Using the operators of total derivatives
d\ = d
x
+ y\di,
302
CHAPTER 6. RELATIVISTIC MECHANICS
one can obtain the transition functions of the coordinates (6.1.5) under coordinate
transformations (6.1.4) in explicit form. Given coordinate transformations (6.1.4),
one can easily find that
d~
x
= [d~,g
a
{x^^)]d
xa
.
(6.1.6)
Then we have
t-Kw+K&ypM] ( ^
+
^ )
/ w )
-
(6.1.7)
It is readily observed that the transition functions (1.3.1) are a particular case of
the coordinate transformations (6.1.7) when the transition functions g
a
(6.1.4) are
independent of coordinates y
1
. Note that, in contrast with (1.3.1), the coordinate
transformations (6.1.7) are not affine. It follows that the fibration J\Z >Z is not
an affine bundle.
There is one-to-one correspondence
A
:
[5]^x^+y-([5U)a,)
(6.1.8)
between the jets [S]\ at a point z Z and n-dimensional vector subspaces of the
tangent space T
Z
Z. It follows that the jet bundle J\Z ->Z has the structure group
GL(n, m; M) C GL(m + n, R)
of linear transformations of the vector space M
m+n
which preserve its subspace R
n
.
Its typical fibre is the Grassman manifold
<d{m + n,n) = GL(n + m\ R)/GL{n, m; R)
of n-dimensional subspaces of the vector space R
m+n
. Thus, the jet bundle J\Z >Z
is a Grassman bundle [45, 70]. In particular, if n = 1, the fibre coordinates y'
0
on
J\Z >Z with the transition functions (6.1.7) are precisely the standard coordinates
of the projective space RP
m
.
Exa mple 6.1.3. Let Y >X be an (m + n)-dimensional fibre bundle over an n-
dimensional base X, and J
l
Y the first order jet manifold of its sections. Let JY
be the first order jet manifold of n-dimensional subbundles of Y. Then the injection
J
l
Y <->J^Y (6.1.1) is an affine subbundle of the jet bundle J\Y -+ Y. Its fibre
at a point y &Y consists of the n-dimensional subspaces of the tangent space T
y
Y
whose intersections with the vertical subspace V
V
Y of T
y
Y are the zero vector.
6.2.
RELATIVISTIC VELOCITY AND MOMENTUM PHASE SPACES 303
Rema r k 6.1.4. Generalizing the notion of a connection on a bundle, one may
think of global sections of the jet bundle J\Z - >Zas being preconnections on the
manifold Z [134]. By virtue of the well-known theorem [80], if such a preconnection
T exists, its image T(Z) in the tangent bundle TZ >Z by the correspondence A
(6.1.8) is a vector subbundle of TZ with the structure group GL(n, K). The quotient
TZjY(Z) is also a vector subbundle with the structure group GL(m, M). Thus, we
have the decomposition
TZ = T{Z) e TZ/T(Z)
of the tangent bundle TZ, which can be treated as a horizontal splitting with respect
to the preconnection I \ However, it should be emphasized that, since J\Z Z is
not an affine bundle, preconnections do not make an affine space. Moreover, it may
happen that a preconnection does not exist on a manifold Z.
6.2 Relativistic velocity and momentum phase spaces
Let us now turn to the case of relativistic mechanics, where Z = Q is an (m + 1)-
dimensional configuration space which has not necessarily a fibration or a preferable
fibration over the time axis E. We call Q a reiativistic configuration space in compar-
ison with a non-relativistic configuration space Q endowed with a fibration Q >M.
This configuration space is provided with the coordinate atlas (6.1.3) together with
the transition functions (6.1.4), i.e.,
(q,q%
i = 1,... , m,
-W),
9W( V) -
(6.2.1)
The coordinates q in different charts of Z play the role of a temporal coordinate.
Note that, given a coordinate chart (U; q, q'), we have a local fibre bundle
U 5 (q,q
i
)^q QCR
(6.2.2)
which can be treated as a configuration space of a local non-relativistic mechanical
system.
The velocity phase space of relativistic mechanics on the configuration space Q
is the first order jet manifold J\Q of 1-dimensional subbundles of Q. It is endowed
304 CHAPTER 6. RELATIVISTIC MECHANICS
with the coordinates (q, q\qo) (6.1.5). Their transition functions are obtained as
follows.
Given coordinate transformations (6.2.1), the total derivative (6.1.6) reads
d~
0
djb(q)d
q
o
[dq
+q
dq)
dqO.
In accordance with the relation (6.1.7), we have
I
WK(?)
\d<p
+%
d<?)
The solution of this equation is
9o
I
(8? >9?\
[dqO
+ 9o
dq" )
(6.2.3)
A glance at this transformation law shows that the velocity phase space J\Z * Z
of relativistic mechanics is really a projective bundle.
Exa mple 6.2.1. Let consider the configuration space Q = M
4
, provided with the
Cartesian coordinates
(V).
i 1,2,3.
This is the case of Special Relativity. For instance, let
9
gcha q
1
sha,
f
-<7sha + g'cha, (6.2.4)
<?'* q
2
'
3
-
be a Lorentz transformation of the plane (q, q
1
). Substituting these expressions in
the formula (6.2.3), we obtain
ql
sha +ggcha
cha q^sha
^
3
9o
2,3
go
cha go sha
This is precisely the transformation law of 3-velocities in Special Relativity if we
put
cha
1
sha
v
6.2. RELATIVISTIC VELOCITY AND MOMENTUM PHASE SPACES 305
where v is the velocity of a reference frame moving along the axis q
1
.
Thus, one can think of the velocity phase space J\Q of relativistic mechanics
as the space of 3-velocities v of a relativistic system. For the sake of convenience,
we will call J\Q the 3-velocity phase space, though the dimension of Q is not equal
necessarily to 3 +1. Given a coordinate chart {U;q,q
l
), 3-velocities v = (v
l
0
) of a
relativistic system can be seen as velocities of a local non-relativistic system (6.2.2)
with respect to the corresponding local reference frame T = d
t
on U. However, the
notion of a reference frame in non-relativistic mechanics has no meaning in relativis-
tic mechanics since the relativistic transformations q
l
0
> <$ and P > P
1
are not
affine and the relative velocity gj P is not maintained under these transformations.
To introduce the relativistic velocities, let us consider the tangent bundle TQ of
the configurations space Q. It is equipped with the holonomic coordinates
(9,g\9,<f)-
In accordance with (6.1.8), there is the multivalued morphism A from the 3-velocity
phase space J\Q to the tangent bundle TQ when a point (q,q',q'
0
) J\Q corre-
sponds to a line
A^j^teV^JKTQ
(6.2.5)
in the tangent space to Q at the point (q,q')- Conversely, there is the morphism
Q
TQ-+JIQ,
QoQ
9'
(6.2.6)
such that
go X
IdJlQ.
Indeed, it is readily observed that q'
0
and q
l
/q have the same transformation laws.
It should be emphasized that, though the expressions (6.2.5) and (6.2.6) are singular
at 5 =0, this point belongs to another coordinate chart, and the morphisms A and
g are well defined.
Thus, one can think of the tangent bundle TQ as being the space of relativistic
velocities or 4-velocities of a relativistic system. It is called the4-velocity phase
space.
306 CHAPTER 6. RELATIVISTIC MECHANICS
Remar k 6.2.2. Note that, with respect to the universal unit system, the elements
of TQ are not physically dimensionless, in contrast with the physical relativistic
velocities.
Since the morphism A of J\Q onto TQ is multivalued and the converse morphism
(6.2.6) is a surjection, one may try to find a subbundle W of the tangent bundle TQ
such that
Q:W-J%
is an injection. Let us assume that Q is oriented and endowed with a pseudo-
Riemannian metric g with the signature (+, ...). The pair (Q,g) is called a
relativistic system. This metric g defines the subbundle of veJ ocity hyperboloids
W
g
= {q
x
6 TQ : g^q)?? = 1}.
(6.2.7)
Of course, W
g
is neither a vector nor an affine subbundle of TQ. Let Q be time-
oriented with respect to the pseudo-Riemannian metric g. By definition, this means
that W
g
is a disjoint union of two connected subbundles W
+
and W~ [77, 155].
Then it is readily observed that the restriction of the morphism g (6.2.6) to each of
these subbundles is an injection into J\Q.
Let us consider the image of this injection in a fibre of J\Q at a point q Q.
There are local coordinates (q, q') on a neighbourhood of q Q such that the
pseudo-Riemannian metric g(q) at q comes to the Minkowski metric
g(q) =?7 =di ag(l ,- l , ,-!). (6.2.8)
With respect to these coordinates, the velocity hyperboloid W
q
C T
q
Q is given by
the equation
(9
0
)
2
- E(<n
2
= i-
i
This is the union of the subset W+, where q > 0, and W~, where q < 0. Restricted
to W
+
, the morphism (6.2.6) takes the familiar form of the relations between 3- and
4-velocities
yfl ~ Ulb)
2
\A-n<?i)
2
6.3. KELATTVISTIC DYNAMICS 307
in Special Relativity. The image of each of the hyperboloids W* in the 3-velocity
phase space J{Q by the morphism (6.2.6) is the open ball
B* o)
2
<1- (6.2.9)
i
This relation means that 3-velocities of a relativists system (Q,g) are bounded in
accordance with the relativistic principle.
Let us turn now to the momentum phase space of relativistic mechanics. Given
a chart of the relativistic configuration space Q, the homogeneous Legendre bundle
corresponding to the local non-relativistic system (6.2.2) is the cotangent bundle
T'Q. This fact motivates us to choose T'Q as the reJ ativistic momentum phase
space, equipped with the holonomic coordinates ($\p
A
). The cotangent bundle
T'Q is endowed with the canonical symplectic form
n =4p
M
AdgM, (6.2.10)
called the reJ ativistic symplectic form.
Given a pseudo-Riemannian metric g on the relativistic configuration space Q,
we have the corresponding isomorphism between relativistic velocity and momentum
phase spaces TQ and T'Q. Prom the physical viewpoint, this isomorphism, however,
fails to be correct. With respect to the universal unit system, a metric tensor g is
dimensionless, whereas 4-velocities and relativistic momenta have different physical
dimension. The latter are of dimension [length]"
1
. Therefore, we should use a mass
metric in order to perform the above-mentioned isomorphism.
6.3 Relat ivist ic dynamics
In a straightforward manner, Lagrangian formalism fails to be appropriate to rela-
tivistic mechanics because a Lagrangian
L = Cdq, (6.3.1)
by very definition, is defined only locally on a coordinate chart (6.1.5) of the 3-
velocity phase space J\Q. For instance, the Lagrangian
L
m
= -mil - E(9J )
2
V (6.3.2)
308 CHAPTER 6. RELATIVISTIC MECHANICS
of a free relativistic point mass m in Special Relativity, can be defined on each
coordinate chart of J\Q, but it is not maintained in a straightforward manner by the
Lorentz transformations (6.2.4) because of the termdq. Given a motion q* = c
l
(q)
with respect to a coordinate chart (q, q'), its Lorentz transformation (6.2.4) reads
cf
>
= qcha-c
1
(q
0
)aha,
c
1
(9) = -g
0
(g
0
)sha +c
1
(<
?
0
(g))cha,
^ V ) = c
2
'V(9)).
Then we have the Lorentz invariance of the pull-back c*L
m
of the Lagrangian (6.3.2),
i.e.,
c L
m
c L
m
where dg
0
=d
?
o(5)dg and d
q
o is the total derivative.
Therefore, we concentrate our consideration on Hamiltonian relativistic mechan-
ics.
Let L be a local regular Lagrangian (6.3.1) on a coordinate chart (q, q
x
) of the
3-velocity phase space J\Q- In accordance with Remark 4.8.9, it defines a local
Dirac constraint system on the relativistic momentum phase space T*Q in the case
of a zero Hamiltonian and the primary constraint space given by
dq'
0
dC
dq'o
(6.3.3)
Since the Lagrangian L is regular, the equations (6.3.3) are solvable for
Po =Po(<?",Pi)-
Then a solution of the Lagrange equations is an integral curve of the vector field
tf = d
0
+ Pdi + Ad
{
on the primary constraint space (6.3.3), which fulfills the equation
tfjn
N
= 0,
tfi =diPo, * = d'p
0
,
where fi^is the restriction of the relativistic symplectic form (6.2.10) to the con-
straint space (6.3.3). For instance, the Lagrangian (6.3.2) (which is regular on the
6.3. RELATIVISTIC DYNAMICS 309
ball (6.2.9)) defines a Dirac constraint system on T*Q in the case of a zero Hamil-
tonian and the primary constraint space
2 V^ 2 2
i
(6.3.4)
Therefore, let us describe a relativistic system (Q, g) as an autonomous Hamil-
tonian system on the symplectic manifold T*Q, characterized by a Hamiltonian
H : T'Q - R,
(6.3.5)
called a relativistic Hamiltonian.
Rema r k 6. 3. 1. It should be emphasized that the coordinate po, but not the
relativistic Hamiltonian H, plays the role of a relativistic energy function.
Any relativistic Hamiltonian H (6.3.5) defines the Hamiltonian map
H : T*Q -* TQ,
q = d*H, (6.3.6)
(3.3.4) over Q from the relativistic momentum space T*Q to the 4-velocity phase
space TQ. Since the 4-velocities of a relativistic system live in the velocity hyper-
boloids (6.2.7), we have the constraint subspace
N = WW
g
, (6.3.7)
g^^H ^H = 1,
of the relativistic momentum phase space T*Q. It follows that the relativistic system
(Q, g) can be described as an autonomous Dirac constraint system on the primary
constraint space N (6.3.7) [154]. I ts solutions are integral curves of the Hamiltonian
vector field
u = u'di + Uid
1
on N C T*Q, that obeys the Hamilton equation
u\i*
N
Q = -i*
N
dH.
(6.3.8)
Its solution does not necessarily exist.
310
CHAPTER 6. RELATIVISTIC MECHANICS
Exa mple 6.3.2. The relativistic Hamiltonian of a free relativistic point mass in
Special Relativity is
H = -^ V>
P

(6.3.9)
where r) is the Minkowski metric (6.2.8), while the constraint space N (6.3.7) is given
by the equation (6.3.4). It is readily observed that the restriction of the Hamiltonian
(6.3.9) to this constraint space is a constant function, i.e.,
i"
N
dH =0.
Then the Hamilton equation (6.3.8) takes the form
u\i*
N
Sl = 0.
(6.3.10)
We obtain its solution
pi = Ui = 0,
,
t t
u
V
'
k
p
k
<? =
u
= I . =
sjm? - v
}k
PjPk
and then the familiar expression for 3-velocities
i _ mri
ik
p
k
^Jm
2
- n>
k
pjp
k
Pi = const.

Exa mple 6.3.3. Let us consider a point electric charge e in the Minkowski space in
the presence of an electromagnetic potential Ax- Its relativistic Hamiltonian reads
H = - ^" " ( J V - c4.)(ft "
eA
)<
while the constraint space TV (6.3.7) is
v""{Pn - eA
ll
){p
v
-eA
v
) = m
2
.
As in the previous Example, the Hamilton equation (6.3.8) takes the form (6.3.10).
Its solution is
Pk
(6.3.11) u
k
= u^dkAp,
q
k
fc==
urf
k
(pj - eAj)
Jm
2
- rfUjpi - eAi){
Pj
- eAj)
(6.3.12)
6.4. RELATIVISTIC GEODESIC EQUATIONS 311
The equality (6.3.12) leads to the usual expression for the 3-velocities
Pk I : T -rt-k
Substituting this expression in the equality (6.3.11), we obtain the familiar equation
of motion of a relativistic charge in an electromagnetic field.
Exa mpl e 6.3.4. The relativistic Hamiltonian for a point mass m in a gravitational
field g on a 4-dimensional manifold Q reads
H
-L
9r{
-
q)
*>*
while the constraint space N (6.3.7) is
9*
V
VvP =
m2
-
As in previous Examples, the equation (6.3.8) takes the form (6.3.10).
Examples (6.3.2) - (6.3.4) illustrate the relativistic systems where the restriction
of a relativistic Hamiltonian to the constraint space N (6.3.7) is a constant function.
In this case, the Hamilton equation takes the form (6.3.10). In accordance with
Proposition 3.4.2, this equation has a solution everywhere on the primary constraint
space TV (6.3.7).
Rema r k 6.3.5. Given a symplectic manifold (T*Q,Q) and its submanifold N
(6.3.7), one may apply the reduction procedure described in Section 2.6. For the
relativistic systems in Examples 6.3.2 - 6.3.4, this procedure is not, however, use-
ful since the pull-back Hamiltonian i*
N
Ji on the constraint space N is a constant
function and the associated Hamiltonian vector fields belong to the kernel of the
presymplectic form i*
N
Q on N. *
6.4 Relativistic geodesic equations
Let H be a relativistic Hamiltonian on the relativistic momentum phase space T*Q
and
in
d'Hdi - diHd*
(6.4.1)
312 CHAPTER 6. RELATIVISTIC MECHANICS
the corresponding Hamiltonian vector field whose integral curves are solutions of
the Hamilton equation
frJ fi =-dH.
If the Hamiltonian map H (6.3.6) is a diffeomorphism, the Hamiltonian vector field
(6.4.1) yields the holonomic vector field
e
r Ho^oH- ^^+f i
(6.4.2)
e
(dad^H^H - <9
a
d"Hd
Q
H) o H
1
,
Dn the 4-velocity space 1 Q, where
(*,*, 4", $*)oTH {<f, dH, <f, d
a
dLq
a
+ 09"Hp
o
)
is the tangent morphism to the Hamiltonian map H. The holonomic vector field
(6.4.2) defines a conservative second order dynamic equation
q (d
a
dnd
a
n - d
Q
d"Hd
Q
H) o H
1
, (6.4.3)
called the relativistic dynamic equation, on the relativistic configuration space Q
(see Definition 3.1.2).
For instance, the relativistic dynamic equation (6.4.3) for a point electric charge
in Example 6.3.3 takes the well-known form
q = rf"q
x
F^ (6.4.4)
where F is the electromagnetic strength (4.9.10). The relativistic dynamic equation
(6.4.3) for a point mass m in a gravitational field g in Example 6.3.4 reads
r = {A
M
,}<ZV,
(6.4.5)
where
{A",} ^9
v
'
a
{Q\9oL
V
+d*9a\ - d
a
g\u)
are the Christoffel symbols of the pseudo-Riemannian metric g.
The equations (6.4.4) and (6.4.5) exemplify relativistic dynamic equations which
are geodesic equations
<r W.t fV
(6.4.6)
6.4. RELATIVISTIC GEODESIC EQUATIONS 313
with respect to a connection
K = dq
x
{d
x
+ Kfa)
(6.4.7)
on the tangent bundle TQ * Q (see Definition 3.1.6). We call (6.4.6) the relativistic
geodesic equation. For instance, a connection K in the equation (6.4.5), is the Levi-
Civita connection of the pseudo-Riemannian metric g. In the equation (6.4.4), K is
the zero Levi-Civita connection of the Minkowski metric plus the soldering form
o = rrF
vX
d(?d
ll
.
(6.4.8)
We say that a relativistic geodesic equation (6.4.6) on the 4-velocity phase space
TQ describes a relativistic system(Q, g) if its geodesic vector field does not leave the
subbundle of velocity hyperboloids (6.2.7). It suffices to require that the condition
(dxg^q" +2g^K
x
l
)q
x
q" =0
(6.4.9)
holds for all tangent vectors which belong to W
g
(6.2.7). Obviously, the Levi-Civita
connection {A^I/} of the metric g fulfills the condition (6.4.9). Any connection K on
TQ Q can be written as
^ = {AM?" + *AV,?
A
),
where
a =a$dq
x
d
x
is a soldering form. Then the condition (6.4.9) takes the form
g^ggqV =Q-
(6.4.10)
It is readily observed that the soldering form (6.4.8) in the equation (6.4.4) obeys
this condition for the Minkowski metric r).
Let now compare relativistic and non-relativistic geodesic equations. In physical
applications, one usually thinks of non-relativistic mechanics as being an approxi-
mation of small velocities of relativistic theory. At the same time, the 3-velocities in
mathematical formalism of non-relativistic mechanics are not bounded. It has long
been recognized that the relation between the mathematical schemes of relativistic
and non-relativistic mechanics is not trivial.
314 CHAPTER 6. RELATIVISTIC MECHANICS
Let a relativistic configuration space Q admit a fibration Q >R, where M is a
time axis. One can think of the fibre bundle Q > K as being a configuration space
of a non-relativistic mechanical system.
In order to compare relativistic and non-relativistic dynamics, one should con-
sider pseudo-Riemannian metric on TQ, compatible with the fibration Q R.
Note that M is a time of non-relativistic mechanics. It is one and the same for all
non-relativistic observers. In the framework of a relativistic theory, this time can be
seen as a cosmological time. Given a fibration Q * E, a pseudo-Riemannian metric
on the tangent bundle TQ is said to be admissible if it is defined by a pair (g
R
, T)
of a Riemannian metric on Q and a non-relativistic reference frame T, i.e.,
2rr
R
9=
Jf~f~
9
'
(6.4.11)
I r |
2
= s* FT* = j ^ r ,
in accordance with the well-known theorem [77]. The vector field T is a time-like
vector relative to the pseudo-Riemannian metric g (6.4.11), but not with respect to
other admissible pseudo-Riemannian metrics in general.
There is the canonical imbedding (4.1.4) of the velocity phase space J
l
Q of
non-relativistic mechanics into the affine subbundle
4 =1, ** = &
(6.4.12)
of the 4-velocity phase space TQ. Then one can think of (6.4.12) as the 4-velocities
of a non-relativistic system. The relation (6.4.12) differs from the familiar rela-
tion (6.2.6) between 4- and 3-velocities of a relativistic system. In particular, the
temporal component q of 4-velocities of a non-relativistic system equals 1 (rela-
tive to the universal unit system). It follows that the 4-velocities of relativistic and
non-relativistic systems occupy different subbundles of the 4-velocity space TQ.
Moreover, Theorem 4.4.2 shows that both relativistic and non-relativistic equations
of motion can be seen as the geodesic equations on the same tangent bundle TQ.
The difference between them lies in the fact that their solutions live in the different
subbundles (6.2.7) and (6.4.12) of TQ. At the same time, relativistic equations,
expressed in the 3-velocities q'/q of a relativistic system, tend exactly to the non-
relativistic equations on the subbundle (6.4.12) when q 1, floo * 1, i e., where
non-relativistic mechanics and the non-relativistic approximation of a relativistic
theory only coincide.
6.4. RELATIVISTIC GEODESIC EQUATIONS 315
Let (q, q') be a non-relativistic reference frame on Q compatible with the fibra-
tion of Q R. Given a non-relativistic geodesic equation (4.4.7), we will say that
the relativistic geodesic equation (6.4.4) is the relativization of (4.4.7) if the spatial
parts of these equations are the same.
In accordance with Lemma (4.5.5), any relativistic geodesic equation with respect
to a connection K (6.4.7) is a relativization of a non-relativistic geodesic equation
with respect to the connection
K = dq
x
{d
x
+ {K\-T
i
Ki)d
i
),
where P = 0 is the connection corresponding to the reference frame (q,q'). Of
course, for different reference frames, we have different non-relativistic limits of the
same relativistic equation. The converse procedure is more intricate.
Example 4.9.3 shows that Lagrange equations for any non-relativistic quadratic
Lagrangian (4.9.11) admit a relativization with respect to a Lagrangian frame con-
nection. It follows that inertial forces do not admit relativization. Conversely, let
us consider a geodesic motion
r = {AM?V
(6.4.13)
in the presence of a pseudo-Riemannian metric g on a relativistic configuration space
Q. Let (q
0
,^) be local hyperbolic coordinates such that poo = 1, 9ot = 0. These
coordinates make up a non-relativistic reference frame for a local fibration Q * K.
Then, by virtue of Lemma (4.5.5), the equation (6.4.13) has the non-relativistic limit
(4.9.18) which is the Lagrange equations for the Lagrangian (4.9.16) where $ = 0.
This Lagrangian describes a free non-relativistic mechanical system with the mass
tensor ffiij = gij.
In view of Proposition 4.4.5, the relativization procedure, however, fails to be
unique. Therefore, the relativization (4.9.12) of an arbitrary quadratic Lagrangian
(4.9.11) may lead to confusion. In fact, this relativization procedure is appropriate
to a gravitational Lagrangian (4.9.15), where <jf is a gravitational potential. An
arbitrary quadratic dynamic equation can be written in the form
?So =-(m-
1
)* Wflfoff +Wtf .
9o =1-
where {\k^} are the Christoffel symbols of some pseudo-Riemannian metric g, whose
spatial part is the mass tensor (rn^), while
WteS+W)
(6.4.14)
316 CHAPTER 6. RELATIVISTIC MECHANICS
is an external force. With respect to the coordinates where <?oi = 0, one may
construct the relativistic equation
9* ={AK V +^g\
(6.4.15)
where the soldering forma must fulfill the condition (6.4.10). It takes place only if
QikVj +9ijV
k
= 0,
i.e., the external force (6.4.14) is the Lorentz-type force plus some potential one.
Then, we have
ffg =0,
4 = -</%>&&,
4 = K-
The relativization (6.4.15) exhausts almost all familiar examples. To complete
our exposition, we will point out also another relativization procedure. Let a non-
relativistic equation ^'(x^) be a spatial part of a 4-vector
x
in the Minkowski space.
Then one can write the relativistic equation
a
x
=e - va
0
e<i
a
q
x
-
(6.4.16)
This is the case, e.g., for the relativistic hydrodynamics in gravitation theory. How-
ever, the non-relativistic limit q = 1 of (6.4.16) does not coincide with the initial
non-relativistic equation. There are also other variants of relativistic hydrodynamic
equations [107].
Ap p en d i x A
Geomet r y of BRST mechanics
As is well-known, the BRST formalism is one of the corner stones of contemporary
quantum field theory. This formalism has been extended to mechanics [5, 65, 66, 140,
141]. One can think of BRST mechanics as being a particular case of supersymmetric
mechanics [3, 42, 108], where supercharges are the BRST charges, or mechanics
on supermanifolds [4, 33, 41, 95, 165]. We aim to show that Hamiltonian time-
dependent mechanics on a vertical configuration space VV'Q admits the natural
extension to BRST mechanics on graded manifolds.
We start from the naive BRST extension of Hamiltonian mechanics where the
odd variables are introduced in a formal way.
Let the configuration space Q > K be a vector bundle. Besides the familiar
coordinates (t, q
l
,Pi, q
l
,pi) on the vertical configuration space VV'Q, let us consider
additionally the odd variables
i Cj i ^ Q
(A.l)
which are assumed to be anticommutative. Following the physical terminology, we
will call c and c (A.l) theghosts and antighosts, respectively. Let & and c
1
with
respect to the indices i have the same transition functions as the linear coordinates
q
x
, while c, and Cj have the transition functions of the coordinates p
i;
i.e.,
c"
c,
dq
j
c"
~<
dqi _
dq'
iCj
'
(A.2)
The formal products of the odd variables (A.l) make up the algebra of polyno-
mials over the ring of complex functions on VV*Q. Provided with the unit element,
this is the Z
2
-graded ring V(c,c) of polynomials of the odd variables. One may say
317
318 APPENDIX A. GEOMETRY OF BRST MECHANICS
that such a polynomial is globally defined if it is invariant under the transformations
(A.2). We will follow the convention where the antighosts c are written on the left
of the ghosts c.
For the sake of convenience, we will use the compact notation y
A
for both even
and odd variables. Put {y} =0 for the even variables and {y} = 1 for the odd ones.
Let us consider the left differentiations d/dy of the graded ring V{c,c), which
possess the properties
^
| ( * ) * + (- i )
W{w}
* |; (").
(A.3)
d d
dy
A
dy
B K
' dy
B
dy
A
'
(A.4)
We introduce the exterior forms dy as the dual of the derivatives d/dy such that
{dy} = {y},
(A.5)
^ J *
8
- *
(A.6)
One may define the exterior product of exterior forms dy by the rule
dy A dy
1
= {-l)^^
+1
dy' A dy. (A.7)
With this exterior product, the ring V(c, c) is extended to the (Z, Z
2
)-bi-graded ring
of polynomials V(c, c, dy) which is defined as the algebra of exterior forms dy over
the ring V(c,c), where
ydy' = (-1)WMW)V,
<M <r = ( - 1)
I
*
N +WW
^ A ^, 4>,a e P{c,c,dy).
(A.8)
Accordingly, the interior product (A.6) obeys the relation
^J (0 A a) = (|- j ^) A a + (-l)W+H>0
A (
| . j
f f )
.
(A.9)
The ring V(c, c, dy) is also provided with the exterior differential
*=?
V
^
Wl
(A. 10)
where the derivation
>
4>eV(c,c,dy),
319
fulfills the conditions (A.3), (A.4) and
> = '
As follows from (A.4) and (A.7), the exterior differential (A. 10) is a nilpotent oper-
ator, i.e.,
dod = 0.
It is readily observed that, putting c = 0, c = 0, we restate the familiar exterior
algebra of even variables y.
The main principle of the BRST extension of mechanics is its invariance under
the BRST and anti-BRST transformations whose generators read
o id d . . ^ .. d
& = c
1
- + Ci + iq
l
+ i p;,
oq' dpi oc
1
oct
(Al l )
dq' dpi dc
l l
dci
(A.12)
[65, 66]. The generators (A.11) - (A.12) fulfill the nilpotency rule
M = Q, tftf = 0, && + && = 0,
i.e., for any function f(y), we have
tfJ d(tfJ d/) =0,
0J d(0|d/) =O,
tfJ d(tfJ d/)+tfJ (i(tfJ d/) =0.
Following the criterion of BRST and anti-BRST invariance, let us consider a
Hamiltonian form H (5.2.10) on the momentum phase space V*Q and its vertical
extension Hy (5.11.5) on the vertical momentum phase space VV*Q. We aim to
find a 1-formHe V(c,c, dy) such that the sum
H
s
= Hy + H
c
is BRST-invariant, i.e., its Lie derivative along the vector field i? (A. 11) vanishes:
U{H
V
+ H
c
) = #\d(H
v
+ H
c
) + d($\(H
v
+ H
c
)) = 0.
320 APPENDIX A. GEOMETRY OF BRST MECHANICS
We obtain a desired Hamiltonian form
H
s
= pidq' q*dpi + i(cidc' - c'dci) - b\Hdt - Hcdt,
(A.13)
U
c
= i ^ + c
j
d
j
)(cjd
:
' + Jdj)H,
called the BRST extension of the Hamiltonian form H. It is easily justified that
this form is also anti-BRST-invariant, i.e., its Lie derivative along the vector field d
(A. 12) also vanishes.
In particular, given a splitting of the Hamiltonian
H = p
l
T
i
]
q> + H (A. 14)
with respect to some reference frame T on the configuration space Q >K, there is
the corresponding splitting of the BRST extended Hamiltonian
n
s
p
;
i V - ^( - P A ) +ztfV - ^(-r
jCi
) + dvH +
i{Zjd> + c>d
j
)(c
J
d
j
+ d>d
3
)H
with respect to the connection f (5.2.6) on the momentum phase space V*Q >K.
This splitting shows that the Hamiltonian form (A.13) is globally defined, and that
the connection is given on the odd variables {c',Ci) as on the elements of the fibre
bundle VVQ ->K.
We also can write the BRST extension of the canonical 3-form Q (5.1.9) on the
momentum phase space V*Q. This extension is deduced from the canonical 3-form
flv (5.11.1) on the vertical momentum phase space VV'Q, and reads
fl
s
= [dpi A dq' +dpi A dq
x
+ i{dc\ A dc' - dc
1
A dci)} A dt. (A.15)
The form (A.15) is globally defined. By construction, it is BRST- and anti-BRST-
invariant. The Hamiltonian connection
7
= d
t
+
lt
d< + ya, +
7
<a
i
+ fa, + g, A + f-^ +
9x
j- +
g
'-^ -
for the Hamiltonian formHs (A.13) with respect to the canonical formQ$ satisfies
the Hamilton equations
7* =q\ =Shi, (A. 16a)
7i =Pa = - W , (A. 16b)
f =ql = dhi
s
, (A. 16c)
321
7 = Pu = -9iHs,
(A.16d)
g* = c\ = (cjdi + c
j
d
j
)d
i
H, (A.16e)
g
i
= c
ti
= -(c
j
cV + c>d
j
)d
i
n, (A.16f)
g* =4 =(c,0* + Jd^d'H, (A.16g)
5i =c
ti
= - ( 0 ^+ ^) 9 ^ (A.16h)
for the Hamiltonian form Hs-
It is readily observed that the equations (A. 16a) - (A. 16b) are the Hamilton
equations for the initial Hamiltonian form H. To clarify the meaning of other
equations, let us note that, if y(t) is a solution of the Hamilton equations (A. 16a) -
(A.16h), then the deviations
q^ t) + eecfit) + ec\t) + ec*(t),
Pi(t) + eepi(t) + Ci(t) + Ci{t),
with the odd parameters e and e are solutions of the Hamilton equations for the
Hamiltonian form H extended to the case of even, but not necessarily real variables
Q
l
,Pi-
The following construction plays a prominent role for the BRST quantization
procedure. Let Hbea Hamiltonian function in some decomposition (A. 14) of a Ha-
miltonian 7i. Such a Hamiltonian function becomes an operator under quantization.
Let us consider the operators
n
0
= e
ffH
odo
e
-m = tf _ p(
Cj
&> + (Jdj)n,
/3eR,
Up = er
m
o$oe
im
= d + Pfedt + T?dj)H,
called the BRST and anti-BRST charges, which act on functions f(y) by the law
*(/) = *! #, Hf) = /
It easy to see that these operators are nilpotent, i.e.,
H
p
oH
0
= 0, H
0
oH
p
= 0. (A.17)
Let Hs be the BRST extension of the Hamiltonian function H, which is also treated
as an operator on functions f(y). Then we obtain the relations
" H$ 7~Ls 'Hs Up = 0,
Hf
3
oH
s
-H
s
H
0
= 0,
ipn
s
= H
0
OH
0
+ H
0
O n
0
.
322 APPENDIX A. GEOMETRY OF BRST MECHANICS
These relations together with the relations (A. 17) provide the operators Tip, Tip,
and Hs with the structure of a Lie superalgebra [35], similar to what we have in
supersymmetric mechanics [42, 108].
Now let us restate the above construction in strict mathematical terms. In
general, BRST mechanics can be developed as mechanics on supermanifolds [4, 33].
However, its naive variant shows that it suffices to consider the following particular
case of mechanics on supermanifolds.
First of all, we should restrict our consideration to vector configuration bun-
dles Q * K in order to introduce the generators of BRST and anti-BRST
transformations.
Secondly, the BRST extension TL
S
of a Hamiltonian Ti is a polynomial with
respect to the odd variables. Therefore, we can narrow the class of superfunc-
tions under consideration.
Let us recall a few basic notions. A vector space Q is called a Z
2
-graded vector
space or simply a graded vector space if there is the decomposition
Q = So Qi,
where Qo is called its even subspace, while Qi is said to be the odd subspace. An
element of a graded vector space is called homogeneous if it belongs to <2o or Q\.
The degree of a homogeneous element v e Q is denoted by {v}. A graded vector
space is said to be (m, n)-dimensional if dimQ
0
= m and dim Q
x
= n.
By a graded commutative algebra B is meant an algebra which is a vector graded
space B = B
0
(& Bi such that
a
r
a
s
= (l^ascv G B
T+S
, a
r
e B
T
, a
s
G B
s
, s,r = 0,1.
A graded commutative Banach algebra is a graded commutative algebra if it is a
Banach algebra and the condition
||oo + OiH = |K|| + I M
is fulfilled.
Let V be a vector space, and
A =A V =t 9A V
k
323
its exterior algebra. This is a Z-graded commutative algebra provided with Z
2
-
graded structure as follows:

2 m
.
A
0
=1 A V,
m
2m+l
Ai = A V.
m
It is called the Grassman algebra. Given a basis {c
l
,i /} for the vector space V,
the elements of the Grassman algebra take the form
a = E E ^i-i.e*
1
c**,
(A. 18)
where the sum is over all the collections of indices (i
x
i
k
) such that no two of
them are the permutations of each other. For the sake of simplicity, we will omit
the symbol of the exterior product of elements of the Grassman algebra A. The
Grassman algebra becomes a graded commutative Banach algebra if its elements
(A. 18) are endowed with the norm
NI =E E K-i J -
k (t i-i
f c
)
Rema r k 1. 0. 1. There is a different definition of a Grassman algebra [89], which is
equivalent to the above one only in the case of an infinite-dimensional vector space
V [40].
Let Q be a graded vector space and B a graded commutative algebra with a unit
element. If Q is both a left and a right B-module such that
a
3
v
T
= v
T
a
s
Qs+r, v
T
Q
r
,
a
s
G B
s
,
it is called a graded B-module. Each B-graded module is obviously a Bo-module.
Exa mpl e 1.0.2. Every graded vector space Q can be extended to a graded B-
module which is its graded B-envelope
BQ (BQ)
0
(BQ)! = (B
0
Qo B\ fii) (Bi Qo B
0
Qi).
The graded envelope
gm+n
=
(
f i
m
0
jgn)
0
(_gm
ffi fi
n)
324 APPENDIX A. GEOMETRY OF BRST MECHANICS
of the (m, n)-dimensional graded vector space Q
m,n
is said to be a vector superspace.
Its Bo-submodule
B
m
'
n
=5J
1
0 B?
is also called a vector superspace. If B is a Banach algebra, the vector superspace
B
m+n
is a Banach space provided with the norm
Ikcii
11*11.
i
a; B,
where {c
1
} is a basis for Q
m,n
.
The main ingredient in a theory of supermanifolds is the sheaf B of graded
commutative algebras on a manifold Z. Let (Z,B{U)) be its canonical presheaf
where by B{U) is meant the graded commutative algebra of local sections of the
sheaf B over an open subset U C Z. If the sheaf B fulfills some conditions, called
the Rothstein axioms, it is said to be an R-supermanifold (we refer the reader to
[11, 40, 152] for details). The notion of an .R-supermanifold includes Berezin's graded
manifolds, the supermanifolds of A.Rogers and infinite-dimensional supermanifolds
described by A.J adczyk and K.Pilch.
We restrict our consideration to the following class of .R-supermanifolds [33]. Let
E >Z be a vector bundle with a typical fibre V, and E* >Z the dual of E >Z.
We consider the Grassman fibre bundle
A'=M0(8A*)
Z k Z
(A.19)
over Z. Its typical fibre is the Grassman algebra A* = AV. The sections of the
Grassman fibre bundle make up the sheaf of Grassman algebras B on the manifold
Z, which belongs to the class of graded manifolds. Its sections, i.e., sections of the
fibre bundle (A.19) are called superfunctions. Given bundle coordinates (z
x
, y') with
respect to the bases {CJ} in E >Z, superfunctions take the form
/ = a
il
...
ik
c
i
^ ---c^ ,
k (i i -tfc)
(A.20)
where {c
1
} are the dual bases in E*.
Let us consider the sheaf der B of graded differentiations of the sheaf B. Let U
be an open subset of Z and B(U) the restriction of the sheaf B to U. By a graded
325
differentiation of the sheaf B(U) is meant its endomorphism such that
d : B{U) -* B(U),
&{ab) =d{a)b + (-l)<"><
a
>atf(6), (A.21)
for the homogeneous elements d 6 der 6 and a, b G B(U). The graded differentia-
tions of B(U) constitute the B-module der B(\J), and the presheaf of such a B-module
generates the sheaf der B. One can show that der B is isomorphic to the sheaf of
sections of the fibre bundle
TE AE -> Z.
E
(A.22)
On can think of its elements as being A-valued vector fields, called vector superfields,
on the fibre bundle E. They read
d = aX
j^
+ai
i-
(A.23)
where a
x
, a
1
are superfunctions (A.20) and
_d___d_
dc' dy*
are elements of the holonomic bases in TE. Due to the splitting VE = E x E,
these elements can be identified with the basis elements c, in E. It follows that
the operators d/dc
l
act on superfunctions (A.20) by the law of the exterior product
of vectors and covectors. In accordance with the relation (A.21), they are odd.
The operators d/dz
x
are even, and act on the superfunctions (A.20) as familiar
derivatives.
The dual of the sheaf der B is the sheaf der*S generated by the linear morphisms
der|u ->B(U). (A.24)
This is the sheaf of sections of the fibre bundle
T*EAE->Z.
E
I ts elements take the form
<fi = a.\dz
x
+ aide', (A.25)
where a\, a, are superfunctions (A.20), while dc
1
= dy
x
are elements of the holonomic
bases in T*E. One can think of these elements as being the A-valued 1-forms on the
326 APPENDIX A. GEOMETRY OF BRST MECHANICS
fibre bundle E. Then the morphisms (A.24) can be treated as the interior product
of the vector fields (A.23) and the 1-forms (A.25).
One can introduce the exterior differential of superfunctions
d : B ->der*,
df{d) = (_l )WW^(/) = (_!){/} Wtfjdf,
which is extended to the exterior algebra of A-valued exterior forms on E, called
exterior superforms. These superforms are sections of the fibre bundle
AT*EAE -> Z,
E
(A.26)
where f\T*E is the Grassman fibre bundle induced by T*E,
The vector superfields, seen as sections of the fibre bundle (A.22), and the ex-
terior superforms, represented by sections of the fibre bundle (A.26), are exactly
those mathematical objects which we have considered above on the naive level.
They satisfy the rules (A.3) - (A.10). To obtain the BRST extension of Hamilto-
nian mechanics in terms of vector superfields and exterior superforms, one can put
Z = VV'Q, while E is the vertical tangent bundle of the fibre bundle VVQ - 1
so that V{c,c,dy) = der*B.
Ap p en d i x B
On qua nt um t i me-dependent
mecha nics
Here we are concerned only with a few aspects of quantizing time-dependent me-
chanics.
We will start from the following important remark. Let Q >M be a configura-
tion space of time-dependent mechanics, V*Q its momentum phase space, and H
a Hamiltonian form on V*Q. Let T be a (complete) reference frame together with
the adapted coordinates (t, q') on Q. Since V = 0 relative to these coordinates, the
energy function with respect to this reference frame coincides with the Hamiltonian
H in the decomposition (5.2.10) (see Section 5.7). We have the Hamilton equation
(5.2.18a):
4
d
{
H, (B.l)
where one can think of q\. = q\ as being the relative velocities with respect to the
reference frame T, In another reference frame with the adapted coordinates (, q
11
),
the Hamiltonian
W(t,9*,pJ)
W(*,*(t,flVi),ft(i,9'',j<)) ft(t,fl",rf)*'(t, *,*{)
(B.2)
H(t
i
<?(t,q
H
,p
,
i
),p
i
(t,q'
i
,$)
+
ptr*&8".i0
coincides with the energy function with respect to this new reference frame, while
the energy function with respect to the initial reference frame is
Hr
n{t,<t(t,q'\p
,
%
),p
i
{t,q'
i
,
P
'
i
)) H'(t,q'\p^
rir*(t,9*,?fl
(B.3)
327
328 APPENDIX B. ON QUANTUM TIME-DEPENDENT MECHANICS
(see relation (4.5.2a)). Relative to the coordinates (,</',$)>
t ne
Hamilton equation
(B.l) reads
H
_dH'{t,(p,i/
i
)
qt
~ dp[ '
where 4t
c an
be treated as the relative velocities with respect to the new reference
frame. At the same time, we can rewrite these equations as
a
dH
r
(t,cf,pQ
dpi
where gf. are the relative velocities with respect to the initial reference frame. Thus,
we arrive at the following conclusion. Let we quantize a Hamiltonian system with
a Hamiltonian H, written relative to coordinates (t, q'), and perform a coordinate
transformation. In new coordinates, we quantize the Hamiltonian system with re-
spect to the new reference frame if we use the Hamiltonian (B.2), and quantize
that with respect to the initial reference frame if we use the Hamiltonian function
(B.3). It follows that, in quantum mechanics, a passage from one set of coordinates
to another is not a technical procedure. This explains why most quantization for-
mulations lead to correct result only when a Cartesian coordinate system is used
[47].
Let us point out the two quantization schemes which are invariant under reference
frame transformations.
Let Q * R be a configuration space of time-dependent mechanics and VQ its
momentum phase space, provided with the canonical Poisson structure w (5.1.6).
Let H be a Hamiltonian form on V'Q. Given a reference frame T together with the
adapted coordinates {t,q
%
) on Q, the Hamiltonian connection 7#(5.2.17) for H can
be represented as the sum
1H = d
t
+ 0
n
,
where $ is the Hamiltonian vector field for the Hamiltonian H in the decompo-
sition (5.2.10) with respect to the above-mentioned canonical Poisson structure w.
It follows that, given a reference frame T, a time-dependent Hamiltonian system
with a Hamiltonian form H can be seen as a Poisson Hamiltonian system (w,H).
This Poisson Hamiltonian system can be quantized by the methods of geometric
quantization as follows (we refer the reader to [103, 167, 181, 188] for the geomet-
ric quantization technique). We restrict our consideration to the prequantization
procedure.
329
Let M be the typical fibre of the configuration space Q and T*M the correspond-
ing typical fibre of V*Q, provided with the canonical symplectic formQ, (2.4.2). This
form is exact. Therefore it belongs to the zero integral Chern class 0 H
2
(M,Z).
Then the bundle product
T"M x C - T*M
is the prequa.ntiza.tion bundle over M, provided with the connection
A
dpj & + dq> 0 (dj - 2mpjd
c
),
where c is a coordinate on C, such that its curvature is precisely the 2-form
R iA = -2-irin.
(B.4)
A complete reference frame T define a projection
7r
r
: V*Q -> T'M.
The pull-back of the canonical symplectic form 3 by this projection is the exact
2-form
7rf fi =dpi A dq'
such that
IU
wViW-
Then the pull-back bundle
ir*
r
(T*M x C) =V*Q x C - K*Q (B.5)
is the prequantization bundle over V*Q, provided with the connection
n^A = dp, &+ dq
1
0 (d
t
- 2nipid
c
)
such that its curvature obeys the relation similar to (B.4). It follows that the
Poisson algebra on the Poisson manifold (V*Q,w) is quantized, i.e., any function
/ C D(V
r
*(5) defines the operator
/ : s t->-djf&s + &>f(dj + 2iripj)s (B.6)
330 APPENDIX B. ON QUANTUM TIME-DEPENDENT MECHANICS
on sections s(t, q
j
,Pj) of the prequantization bundle (B.5). Since s are simply com-
plex functions on V*Q, we can rewrite the operator (B.6) in the form
f:s>->{f,s}
v
+
2iripj9' fs.
(B.7)
It is readily observed that this operator is invariant under holonomic transforma-
tions of the momentum phase space V*Q. It follows that prequantization (B.7) is
independent of a reference frame. The result is obvious since the Poisson bivector w
(5.1.6) belongs to the zero element of the second LP-cohomology group Hl
P
(V*Q, w)
(see Example 2.7.5).
Another quantization procedure invariant under holonomic time-dependent trans-
formations is based on the Schrodinger equation. Given a reference frame T together
with the adapted coordinates (t, q') on Q, let H(t, q^Pj) be a Hamiltonian with re-
spect to this reference frame. Then the corresponding Schrodinger equation reads
H(t,q
l
,-ihj-)V. (B.8)
Relative to another reference frame coordinates {t,q"), the Schrodinger equation
(B.8) takes the form
5*
H'(t,q'\-ih^-)V, (B.9)
where H' is the Hamiltonian (B.2). It follows that the Schrodinger equations (B.8)
and (B.9) provide quantization with respect to different reference frames. The fol-
lowing example shows that these quantizations are not equivalent in general.
Exa mple 2.0.3. Let us consider a 1-dimensional motion described by the Ha.mil-
tonian
H

+
^
on the momentum phase space R
3
- R with respect to some inertial reference frame
(t,q,p). The corresponding Schrodinger equation is
V(9)
+
9m
{E-U(q))rl>{q) 0.
Let us consider another reference frame given by a connection T (<?'). The energy
function with respect to the reference frame F reads
UT H-pT.
331
The corresponding Schrodinger equation is
r(q) +f?( - U(q)Mq) - ^Y{qW{q)
0.
Let us substitute
^{q) = p(q)exp(^JT(q)dq
S
j (B.10)
in this equation. We obtain
p"(q) + ^[E- U(q) + ihV(q) + |r
2
(g)] p{q) 0.
(B.H)
For instance, let Y = v = const, be an inertia! reference frame. In this case, the
equation (B.l l ) takes the form
p"
{q
) + ^\E + ^V*~U(q)]p{q) 0.
It is readily observed that energy levels of the quantum system with respect to the
moving inertial reference frames are the shifts
E + -S
of those of the initial quantum system. Moreover, the transformation ip >p (B.10)
is not an automorphism of the Hilbert space of states of the initial quantum system.
This means that quantum systems with respect to different inertial reference frames
are not equivalent.
Bibliogr aphy
1. M.Abbati, R.Cirelli, S. De Santis and E.Ruffini, The second Noether theorem
in the formalism of jet-bundles: Symmetries and degeneration, J. Geom. Phys.
17 (1995) 321.
2. R.Abraham and J .Marsden, Foundations of Mechanics, Second Edition (Ben-
jamin/Cummings Publ. Comp., London, 1978).
3. A.Adrianov, F.Cannata, M.Ioffe and D.Nishnianidze, Matrix Hamiltonians:
SUSY approach to hidden symmetries, J . Phys. A 30 (1997) 5037.
4. S.Albeverio and Shao-Ming Fei, BRST Structures and Symplectic Geometry
on a Class of Supermanifolds, Lett. Math. Phys. 33 (1995) 207.
5. A.Aringazin, BRS and anti-BRS invariant states in path integral approach to
Hamiltonian and Birkoffian mechanics, Phys. Lett B 314 (1993) 333.
6. V.Arnold, Mathematical Methods of Classical Mechanics (Springer, Berlin,
1978).
7. V.Arnold, V.Kozlov and A.Neishtadt, Mathematical aspects of classical and
celestial mechanics, in Dynamical Systems III, (Springer, Berlin, 1988) p.l .
8. J . de Azcarraga, A.Perelomov and J .Perez Bueno, The Schouten-Nijenhuis
bracket, cohomology and generalized Poisson structures, J. Phys. A 29 (1996)
7993.
9. J . de Azcarraga, J .Izquierdo and J . Pfez Buenot, On the higher-order gener-
alizations of Poisson structures, J. Phys. A 30 (1997) L607.
10. F.Barone, R.Grassini and G.Mendella, A generalized Lagrange equation in
implicit form for non-conservative mechanics, J. Phys. A 30 (1997) 1575.
332
BIBLIOGRAPHY 333
11. C.Bartocci, U.Bruzzo, D.Hernandez Ruiperez, V.Pestov, Foundations of su-
permanifold theory: the axiomatic approach, Diff. Geom. and Appl. 3 (1993)
135.
12. M.Bergvelt and E. de Kerf, The Hamiltonian structure of Yang-Mills theories
and instantons, Physica 139A (1986) 101.
13. K.Bhaskara and K.Vismanath, Poisson Algebras and Poisson Manifolds, Pith-
man Research Notes in Mathematics 174 (Longhman Sci., Harlow, 1988).
14. E.Binz, H.Fischer and J .Sniatycki, Geometry of Classical Fields (North-
Holland, Amsterdam, 1988).
15. D.Blair, Contact Manifolds in Riemannian Geometry, Lect. Notes in Mathe-
matics 509 (Springer, Berlin, 1976).
16. G. Bredon, Sheaf theory (McGraw-Hill, N.-Y., 1967).
17. K.Brown, Cohomology of Groups (Springer-Verlag, Berlin, 1982).
18. A.D.Bruno, The normal form of a Hamilton system, Uspehi Matemat. Nauk
43 (1988) Nl , 23 (in Russian).
19. J -L. Brylinski, A differential complex for Poisson manifolds, J . Diff. Geom.
28 (1988) 93.
20. R.Bryant, S.Chern, R.Gardner, H.Goldschmidt, P.Griffiths, Exterior Differ-
ential Systems (Springer, Berlin, 1991).
21. A.Cabras and A.Vinogradov, Extension of the Poisson bracket to differential
forms and multi-vectors, J . Geom. Phys. 9 (1992) 75.
22. S.Campbell, An Introduction to Differential Equations and their Applications
(Wadsworth Inc., Belmont, 1990).
23. D.Canarutto, Bundle splittings, connections and locally principle fibred man-
ifolds, Bull. U.M.I. Algebra e Geometria Serie VI V-D (1986) 18.
24. F.Cardin and G.Zanzotto, On constrained mechanical systems: DAlembert's
and Gauss' princoples, J. Math. Phys. 30 (1989) 1473.
334
BIBLIOGRAPHY
25. F.Cardin and M.Favretti, On nonholonomic and vakonomic dynamics of me-
chanical systems with nonintegrable constraints, J. Geom. Phys. 18 (1996)
295.
26. J .Carinena, J .Gomis, L.Ibort and N.Roman, Canonical transformation theory
for presymplectic systems, J . Math. Phys. 26 (1985) 1961.
27. J .Carinena and M.Ranada, Poisson maps and canonical transformations for
time-dependent Hamiltonian systems, J. Ma.th.Phys. 30 (1989) 2258.
28. J .Carinena, M.Crampin and L.Ibort, On the multisymplectic formalism for
first order field theories, Diff. Geom. and Appl. 1 (1991) 345.
29. J .Carinena, J .Fernandez-Nunez and E.Martinez, A geometric approach to
Noether's second theorem in time-dependent Lagrangian mechanics, Lett.
Math. Phys. 23 (1991) 51.
30. J .Carinena and M.Ranada, Lagrangian systems with constraints. A geometric
approach to the method of Lagrange multipliers, J . Phys. A 26 (1993) 1335.
31. J .Carinena and J .Fernandez-Nunez, Geometric theory of time-dependent sin-
gular Lagrangians, Fortschr. Phys. 41 (1993) 517.
32. J .Carinena, L.Ibort, G.Marmo and A.Stern, The Feynman problem and the
inverse problem for Poisson dynamics, Phys. Rep. 263 (1995) 153.
33. J .Carinena and H.Figueroa, Hamiltonian versus Lagrangian formulations of
supermechanics, J . Phys. A 30 (1997) 2705.
34. G.Caviglia, Helmholtz conditions, covariance, and invariance identities, Int.
J . Theor. Phys. 24 (1985) 377.
35. S.Cecotti and C.Vafa, Topological anti-topological fusion, Nucl. Phys. B367
(1991) 359.
36. P.Chernoff and J .Marsden, Properties of Infinite Dimensional Hamiltonian
Systems, Lect. Notes in Mathematics 425 (Springer, Berlin, 1974).
37. D.Chinea, M.de Leon, and J .Marrero, The constraint algoritm for time-
dependent Lagrangians, J. Math. Phys. 35 (1994) 3410.
BIBLIOGRAPHY 335
38. D.Chinea, M.de Leon, and J .Marrero, The canonical double complex for J a-
cobi manifolds, C. R. Acad. Sci., Paris I 323 (1996) 637.
39. N.Chitaia, S.Goglidze and Yu.Surovtsev, Dynamical systems with first- and
second-class constraints, Phys. Rev. D 56 (1997) 1135; 1142.
40. R.Cianci, I ntroduction to Supermaniifolds (Bibliopolis, Naples, 1990).
41. R.Cianci, M.Prancaviglia and I.Volovich, Variational calculus and Poincare-
Cartan formalism on supermanifolds, J. Phys. A 28 (1995) 723.
42. F.Cooper, A.Khare and U.Sukhatme, Supersymmetry and quantum mechan-
ics, Phys. Rep. 251 (1995) 267.
43. M.Crampin, F.Cantrijn and W.Sarlet, Lifting geometric objects to a cotangent
bundle, and the geometry of the cotangent bundle of a tangent bundle, J .
Geom. Phys. 4 (1987) 469.
44. P.Dazord, A.Lichnerowicz and C-M.Marie, Structure locale des varietes de
J acobi, J. Math. Pures et Appl. 70 (1991) 101.
45. P.Dedecker, On the generalization of symplectic geometry to multiple integrals
in the calculus of variations, in Differential Geometric Methods in Mathemat-
ical Physics, Lect. Notes in Mathematics 570 (Springer, Berlin, 1977), p. 395.
46. P.Dedecker and W.Tulczyjew, Spectral sequences and the inverse problem of
the calculus of variations, in Differential Geometric Methods in Mathematical
Physics, Lect. Notes in Mathematics 836 (Springer, Berlin, 1980), p. 498.
47. P.A.M.Dirac, The Principles of Quantum Mechanics (Oxford Univ. Press,
Oxford, 1976).
48. W.Domitrz and S.J aneczko, Normal forms of symplectic structures on the
stratified spaces, Colloquium Mathematicum LXVI I I (1995) fasc.l, 101.
49. B.Dubrovin, M.Giordano, G.Marmo and A.Simoni, Poisson brackets on
presymplectic manifolds, Int. J. Mod. Phys. 8 (1993) 3747.
50. A.Echeverria Enrfquez, M.Mufioz Lecanda and N.Roman Roy, Geometrical
setting of time-dependent regular systems. Alternative models, Rev. Math.
Phys. 3 (1991) 301.
336
BIBLIOGRAPHY
51. A.Echeverria Enriquez, M.Murioz Lecanda and N.Roman Roy, Non-standard
connections in classical mechanics, J. Phys. A 28 (1995) 5553.
52. F.Fatibene, M.Ferraris and M.Francaviglia, Noether formalism for conserved
quantities in classical gauge field theories, J . Math. Phys. 35 (1994) 1644.
53. R.Fulp, J .Lawson and L.Norris, Generalized symplectic geometry as a covering
theory for the Hamiltonian theories of classical particles and fields, J. Geom.
Phys. 20 (1996) 195.
54. G.Giachetta and L.Mangiarotti, Gauge-invariant and covariant operators in
gauge theories, Int. J. Theor. Phys. 29 (1990) 789.
55. G.Giachetta, J et manifolds in non-holonomic mechanics, J. Math. Phys. 33
(1992) 1652.
56. G.Giachetta and L.Mangiarotti, Constrained Hamiltonian systems and gauge
theories, Int. J. Theor.Phys. 34 (1995) 2353.
57. G.Giachetta, L.Mangiarotti and G.Sardanashvily, New Lagrangian and Ha-
miltonian Methods in Field Theory (World Scientific, Singapore, 1997).
58. D.Gitman and I.Tyutin, Canonical Quantization of Constrained Fields
(Nauka, M., 1986) (in Russian).
59. A.Gomberoff and S.Hojman, Non-standard construction of Hamiltonian struc-
tures, J.Phys. A 30 (1997) 5077.
60. M.Gotay, J .Nester and G.Hinds, Presymplectic manifolds and the Dirac-
Bergman theory of constraints, J . Math.Phys. 19 (1978) 2388.
61. M.Gotay and J .Nester, Presymplectic Lagrangian systems, Ann. l'lnst. Henri
Poincare 30 (1979) 129; 32 (1980) 1.
62. M.Gotay, On coisotropic imbeddings of presymplectic manifolds, Proc. Amer.
Math. Soc. 84 (1982) 111.
63. M.Gotay, A multisymplectic framework for classical field theory and the calcu-
lus of variations. I. Covariant Hamiltonian formalism, in Mechanics, Analysis
and Geometry: 200 Years after Lagrange, ed. M.Francaviglia (Elsevier Science
Publishers B.V., 1991), p. 203.
BIBLIOGRAPHY 337
64. M.Gotay, A multisymplectic framework for classical field theory and the cal-
culus of variations. II. Space +time decomposition, Diff. Geom. and Appl. 1
(1991) 375.
65. E.Gozzi, M.Reuter and W.Thacker, Hidden BRS invariance in classical me-
chanics, Phys. Rev. D40 (1989) 3363.
66. E.Gozzi, M.Reuter and W.Thacker, Symmetries of the classical path integral
on a generalized phase-space manifold, Phys. Rev. D46 (1992) 757.
67. J .Grabowski and PUrbariski, Tangent lifts of Poisson and related structures,
J . Phys. A 28 (1995) 6743.
68. P.Griffiths, Exterior Differential Systems and the Calculus of Variations
(Birkhauser, Boston, 1983).
69. D.Grigore and O.Popp, The complete classification of generalized homoge-
neous symplectic manifolds, J . Math. Phys. 30 (1989) 2476.
70. D.Grigore, A generalized Lagrangian formalism in particle mechanics and clas-
sical field theory, Fortschr. d. Phys. 41 (1993) 569.
71. F.Guedira and A.Lichnerowicz, Geometrie des algebres de Lie locales de Kir-
illov, J . Math. Pures et Appl. 63 (1984) 407.
72. V.Guillemin, S.Sternberg, Symplectic Techniques in Physics (Cambridge
Univ. Press., Cambridge, 1990).
73. C. Gunther, The polysimplectic Hamiltonian formalism in field theory and
calculus of variations, J. Diff. Geom. 25 (1987) 23.
74. A.Hamoui and A.Lichnerowicz, Geometry of dynamical systems with time-
dependent constraints and time-dependent Hamiltonians: An approach to-
wards quantization, J. Math. Phys. 25 (1984) 923.
75. J .Harlet, Time and time functions in parametrized non-relativistic quantum
mechanics, Class. Quant. Grav. 13 (1996) 361.
76. P.Havas, The connection between conservation laws and invariance groups:
Folklore, fiction and fact, Acta Phys. Austr. 38 (1973) 145.
338
BIBLIOGRAPHY
77. S.Hawking and G.Ellis, The Large Scale Structure of a Space-Time (Cambr.
Univ. Press, Cambridge, 1973).
78. M.Henneaux, Equation of motion, commutation relations and ambiguities in
the Lagrangian formalism, Ann. Phys. 140 (1982) 1.
79. J .Hietarinta, Nambu tensors and commuting vector fields, J . Phys. A 30
(1997) L27.
80. F.Hirzebruch, Topological Methods in Algebraic Geometry (Springer, Berlin,
1966).
81. S.Hojman and L.Urrutia, On the inverse problem of the calculus of variations,
J. Math. Phys. 22 (1981) 1896.
82. S.Hojman, The construction of a Poisson structure out of a symmetry and a
conservation law of a dynamical system, J . Phys. A 29 (1996) 667.
83. R.Ibaiiez, M.de Leon, J .Marrero and D.Martin de Diego, Co-isotropic and
Legendre-Lagrangian submanifolds and conformal J acobi morphisms, J. Phys.
A 30 (1997) 5427.
84. R.Ibaiiez, M.de Leon, J .Marrero and D.Martm de Diego, Dynamics of Poisson
and Nambu-Poisson brackets, J. Math. Phys. 38 (1997) 2332.
85. R.Ibaiiez, M.de Leon, J .Marrero and E.Padron, Nambu-J acobi and general-
ized J acobi manifolds, J. Phys. A 31 (1998) 1267.
86. L.Ibort and J .Marin-Solano, A geometric classification of Lagrangian functions
and the reduction of evolution space, J . Phys. A 25 (1992) 3353.
87. L.Ibort, M. de Leon and G.Marmo, Reduction of J acobi manifolds, J . Phys.
A 30 (1997) 2783.
88. A.Ibort, M. de Leon, E.Lacomba, J .Marrero, D.Martin de Diego and
P.Pitanga, Geometric formulation of mechanical systems subjected to time-
dependent one-sided constraints, J . Phys. A 31 (1998) 2655.
89. A.J adczyk and K.Pilch, Superspaces and Supersymmetries, Commun. Math.
Phys. 78 (1981) 391.
BIBLIOGRAPHY 339
90. F.Kamber and P.Tondeur, Foliated Bundles and Characteristic Classes, Lect.
Notes in Mathematics 493 (Springer, Berlin, 1975).
91. I.Kanatchikov, On field theoretical generalization of a Poisson algebra, Rep.
Math. Phys. 40 (1997) 225.
92. I.Kanatchikov, Towards to Born-Weyl quantization of fields, Int. J. Theor.
Phys. 37 (1998) 333.
93. G.Katzin and J .Levine, Characteristic functional structures of infinitesimal
symmetry mappings of dynamical systems. I. Velocity-dependent mappings of
second-order differential equations, J . Math. Phys. 26 (1985) 3080.
94. G.Katzin and J .Levine, Characteristic functional structures of infinites-
imal symmetry mappings of dynamical systems. IV. Classical (velocity-
independent) mappings of second-order differential equations, J. Math. Phys.
30 (1988) 2039.
95. O.Khudaverdian and A.Nersessian, Canonical Poisson brackets of different
gradings and strange superalgebras, J. Math. Phys. 32 (1991) 1938.
96. J .Kijowski and W.Tulczyjew, A Symplectic Framework for Field Theories
(Springer, Berlin, 1979).
97. T.Kimura, Generalized classical BRST cohomology and reduction of Poisson
manifolds, Commun. Math. Phys. 151 (1993) 155.
98. A.Kirillov, Local Lie algebras, Russian Math. Surveys 31 (1976) 55.
99. A.Kirillov, Geometric quantization, in Dynamical Systems IV, eds V.I.Arnol'd
and S.PNovikov (Springer, Berlin, 1990) p.137.
100. S.Kobayashi and K.Nomizu, Foundations of Differential Geometry, Vol.1.
(J ohn Wiley, N.Y. - Singapore, 1963).
101. I.Kolaf, P.Michor and J .Slovak, Natural Operations in Differential Geometry
(Springer, Berlin, 1993).
102. Y.Kosmann-Schwarzbach and F.Magri, Poisson-Nijenhuis structures, Ann.
l'lnst. Henri Poincare 53 (1990) 35.
340
BIBLIOGRAPHY
103. B.Kostant, Quantization and unitary representation, in Lectures in Modern
Analysis and Applications III, Lect. Notes in Mathematics, 170 (Springer,
Berlin, 1970), p.87.
104. I.Krasil'shchik, V.Lychagin and A.Vinogradov, Geometry of Jet Spaces and
Nonlinear Partial Differential Equations (Gordon and Breach, Glasgow, 1985).
105. D.Krupka, Some geometric aspects of variations! problems in fibred manifolds,
Folia Fac. Sci. Nat. UJEP Brunensis 14 (1973) 1.
106. O.Krupkova, Mechanical systems with nonholonomic constraints, J . Math.
Phys. 38 (1997) 5098.
107. B.Kupershmidt, The Variational Principles of Dynamics (World Scientific,
Singapore, 1992).
108. A.Lahiri, P.K.Roy and B.Bagchi, Supersymmetry in quantum mechanics, Int.
J. Mod. Phys. A 5 (1990) 1383.
109. S.Lang, Differential Manifolds (Addison-Weslay, Reading, Massachusetts,
1972).
110. M.de Leon and P.Rodrigues, Methods of Differential Geometry in Analytical
Mechanics (North-Holland, Amsterdam, 1989).
111. M.de Leon and J . Marrero, Constrained time-dependent Lagrangian systems
and Lagrangian submanifolds, J. Math. Phys. 34 (1993) 622.
112. M.de Leon and D. Martin de Diego, On the geometry of non-holonomic La-
grangian systems, J . Math. Phys. 37 (1996) 3389.
113. M.de Leon, J . Marrero and D. Martin de Diego, Mechanical systems with
nonlinear constraints, Int. J . Theor. Phys. 36 (1997) 979.
114. M.de Leon, J . Marrero and D. Martin de Diego, Non-holonomic Lagrangian
systems in jet manifolds, J. Phys. A 30 (1997) 1167.
115. M.de Leon, J . Marrero and E.Padron, Lichnerowicz-J acobi cohomology, J .
Phys. A 30 (1997) 6029.
BIBLIOGRAPHY 341
116. P.Libermann and C-M.Marie, Symplectic Geometry and Analitical Mechanics
(D.Reidel Publishing Company, Dordrecht, 1987).
117. A.Lichnerowicz, Les varietes de Poisson et leurs algebres de Lie associees, J .
Diff. Geom. 12 (1977) 253.
118. A.Lichnerowicz, Les varietes de J acobi et leurs algebres de Lie associees, J .
Math. Pures et Appl. 57 (1978) 453.
119. R.Littlejohn and M.Reinsch, Internal or shape coordinates in the n-body prob-
lem, Phys. Rev. A 52 (1995) 2035.
120. R.Littlejohn and M.Reinsch, Gauge fields in the separation of rotations and
internal motions in the n-body problem, Rev. Mod. Phys. 69 (1997) 213.
121. L.Lusanna, An enlarged phase space for finite-dimensional constrained sys-
tems, unifying their Lagrangian, phase- and velocity-space descriptions, Phys.
Rep. 185 (1990) 1.
122. S.Mac Lane, Homology (Springer, Berlin, 1967).
123. C.-M.Marie, On J acobi manifolds and J acobi bundles, in Symplectic Ge-
ometry, Groupoids, and Integrable Systems, ed. PDazord and A.Weinstein
(Springer, Berlin, 1989), p. 227.
124. C.-M.Marie, Reduction of constrained mechanical systems and stability of
relative equilibria, Commun. Math. Phys. 174 (1995) 295.
125. C.-M. Marie, The Schouten-Nijenhuis bracket and interior products, J . Geom.
Phys. 23 (1997) 350.
126. G.Marmo, E.Saletan, A.Simoni and B.Vitale, Dynamical Systems. A Differ-
ential Geometric Approach to Symmetry and Reduction (J ohn Wiley, N.Y.,
1985).
127. G.Marmo, G.Mendella and W.Tulczyjew, Constrained Hamiltonian systems
as implicit differential equations, J. Phys. A 30 (1997) 277.
128. G.Marmo, G.Vilasi and A.Vinogradov, The local structure of n-Poisson and
n-J acobi manifolds and some applications, J. Geom. Phys (1998) (appear).
342
BIBLIOGRAPHY
129. J .Marsden and T.Ratiu, Reduction of Poisson manifolds, Lett. Math. Phys.
11 (1986) 161.
130. J .Marsden and T.Ratiu, Introduction to Mechanics and Symmetries (Springer,
Berlin, 1994).
131. G.Martin, A Darboux theorem for multi-symplectic manifolds, Lett. Math.
Phys. 16 (1988) 133.
132. E.Massa and E.Pagani, J et bundle geometry, dynamical connections and the
inverse problem of Lagrangian mechanics, Ann. Inst. Henri Poincare 61 (1994)
17.
133. P.Michor, A generalization of Hamiltonian mechanics, J. Geom. Phys. 2 (1985)
N2, 67.
134. M.Modugno, A.Vinogradov, Some variations on the notion of connections,
Ann. Matem. Pura ed Appl. CLXVI I (1994) 33.
135. R. Montgomery, The connection whose holonomy is the classical adiabatic
angles of Hannay and Berry and its generalization to the non-integrable case,
Commun. Math. Phys. 120 (1988) 269.
136. G.Morandi, C.Ferrario, G.Lo Vecchio, G.Marmo and C.Rubano, The inverse
problem in the calculus of variations and the geometry of the tangent bundle,
Phys. Rep. 188 (1990) 147.
137. M.Munoz-Lecanda, Hamiltonian systems with constraints: A geometric ap-
proach, Int. J. Theor. Phys. 28 (1989) 1405.
138. M.Munoz-Lecanda and N.Roman-Roy, Lagrangian theory for presymplectic
systems, Ann. Inst. Henri Poincare 57 (1992) 27.
139. M.Munoz-Lecanda and N.Roman-Roy, Gauge systems: Presymplectic and
group action theory, Int. J. Theor. Phys. 32 (1993) 2077.
140. Kh.Nirov and A.Razumov, Equivalence between Lagrangian and Hamiltonian
BRST formalisms, J. Math. Phys. 34 (1993) 3933.
BIBLIOGRAPHY 343
141. Kh.Nirov and A.Razumov, Generalized Schrodinger representation in BRST
quantization, Nucl. Phys. B429 (1994) 389.
142. L.Norris, Schouten-Nijenhuis brackets, J. Math. Phys. 38 (1997) 2694.
143. F.Pardo, The Helmholtz conditions in terms of constants of motion in classical
mechanics, J . Math. Phys. 30 (1989) 2054.
144. P.Pereshogin and P.Pronin, Geometrical treatment of nonholonomic phase in
quantum mechanics and applications, Int. J . Theor. Phys. 32 (1993) 219.
145. J .Perez Bueno, Generalized J acobi structures, J. Phys. A 30 (1997) 6509.
146. V.Perlick, The Hamiltonization problem from a global viewpoint, J. Math.
Phys. 33 (1992) 599.
147. G.Pidello and W.Tulczyjew, Derivations of differential forms on jet bundles,
Ann. Mat. Pura Appl. 147 (1987) 249.
148. R.De Pietri, L.Lusanna and M.Pauri, Standard and generalized Newtonian
gravities as "gauge" theories of the extended Galilei group: I. The standard
theory, Class. Quant. Grav. 12 (1995) 219.
149. J .Pommaret, Systems of Partial Differential Equations and Lie Pseudogroups
(Gordon and Breach, Glasgow, 1978).
150. B.Reinhart, Differential Geometry and Foliations (Springer, Berlin, 1983).
151. F.Riewe, Nonconservative Lagrangian and Hamiltonian mechanics, Phys. Rev.
E53 (1996) 1890.
152. M.Rothstein, The axioms of supermanifolds and a new structure arising from
them, Trans. Amer. Math. Society 297 (1986) 159.
153. C.Rovelli, Quantum mechanics without time: A model, Phys. Rev. D42
(1990) 2638.
154. C.Rovelli, Time in quantum gravity: An hypothesis, Phys. Rev. D43 (1991)
442.
344
BIBLIOGRAPHY
155. R.Sachs and H.Wu, General Relativity for Mathematicians (Springer, Berlin,
1977).
156. R.Santilli, Foundations of Theoretical Mechanics I (Springer, N.Y., 1978).
157. G.Sardanashvily, Gauge Theory in Jet Manifolds (Hadronic Press, Palm Har-
bor, 1993).
158. G.Sardanashvily, Constraint field systems in multimomenturn canonical vari-
ables, J. Math. Phys. 35 (1994) 6584.
159. G.Sardanashvily, Generalized Hamiltonian Formalism for Field Theory. Con-
straint Systems. (World Scientific, Singapore, 1995).
160. G.Sardanashvily, Stress-energy-momentum tensors in constraint field theories,
J. Math. Phys. 38 (1997) 847.
161. G.Sardanashvily, Hamiltonian time-dependent mechanics, J. Math. Phys. 39
(1998) 2714.
162. W.Sarlet, F.Cantrijn and D.Saunders, A geometric famework for the study of
non-holonomic Lagrangian systems, J. Phys. A 28 (1995) 3253.
163. W.Sarlet, F.Cantrijn and D.Saunders, A geometric famework for the study of
non-holonomic Lagrangian systems: II, J . Phys. A 29 (1996) 4265.
164. D.Saunders, The Geometry of Jet Bundles (Cambr. Univ. Press, Cambridge,
1989).
165. A.Schwarz, Geometry of Batalin-Vilkovisky quantization, Commun. Math.
Phys. 155 (1993) 249.
166. J .Slawianowski, Geometry of Phase Spaces (Wiley, N.Y.,1991).
167. J .Sniatycki, Geometric Quantization and Quantum Mechanics (Springer,
Berlin, 1980)
168. J .Souriau, Structures des Systemes Dynamiques (Dunod, Paris, 1970).
169. J .Stasheff, Homological reduction of constrained Poisson algebras, J . Diff.
Geom 45 (1997) 221.
BIBLIOGRAPHY 345
170. N.Steenrod, The Topology of Fibre Bundles (Princeton Univ. Press, Prince-
ton, 1972).
171. P.Stefan, Accessible sets, orbits and foliations with singularities, Proc. London
Math. Soc. 29 (1974) 699.
172. K.Sundermeyer, Constrained Dynamics (Springer, Berlin, 1982).
173. H.Sussmann, Orbits of families of vector fields and integrability of distribu-
tions, Thins. Amer. Math. Soc. 180 (1973) 171.
174. F.Takens, Symmetries, conservation laws and variational principles, in Geom-
etry and Topology, Lect. Notes in Mathematics 597 (Springer-Verlag, Berlin,
1977), p. 581.
175. L.Takhtajan, On foundations of the generalized Nambu mechanics, Commun.
Math. Phys. 160 (1994) 295.
176. W.Tulczyjew, Lagrangian submanifolds and Hamiltonian dynamics, C. R.
Acad. Sci. Paris A 283 (1976) 15.
177. W.Tulczyjew, Lagrangian submanifolds and Lagrangian dynamics, C. R.
Acad. Sci. Paris A 283 (1976) 675.
178. W.Tulczyjew, The Legendre transformation, Ann. Inst. Henri Poincare A
XXVI I (1977) 101.
179. W.Tulczyjew, The Euler-Lagrange resolution, in Differential Geometric Meth-
ods in Mathematical Physics, Lect. Notes in Mathematics 836 (Springer-
Verlag, Berlin, 1980), p. 22.
180. W.Tulczyjew, Geometric Formulation of Physical Theories (Bibliopolis,
Naples, 1989).
181. I.Vaisman, Lectures on the Geometry of Poisson Manifolds (Birkhauser Ver-
lag, Basel, 1994).
182. I.Vaisman, Second order Hamiltonian vector fields on tangent bundles, Diff.
Geom. and Appl. 5 (1995) 153.
346
BIBLIOGRAPHY
183. A.Vershik, Classical and non-classical dynamics with constraints, in Global
Analysis - Studies and Applications 1, Lect. Notes in Mathematics 1108
(Springer, Berlin, 1984), p. 278.
184. A.Vershik, V.Gershkovich, Nonholonomic dynamical systems, in Dynamical
Systems VII, eds V.I.Arnol'd and S.P.Novikov (Springer, Berlin, 1994) p.l .
185. F.Warner, Foundations of Differential Manifolds and Lie Groups (Springer,
Berlin, 1983).
186. A.Weinstein, The local structure of Poisson manifolds, J. Diff. Geom. 18
(1983) 523.
187. A.Weinstein, Co-isotropic calculus and Poisson gruppoids, J . Math. Soc.
Japan 40 (1988) 705.
188. N.Woodhouse, Geometric Quantization (Clarendon Press, Oxford, 1980).
189. K.Yano and S.Ishihara, Tangent and Cotangent Bundles, Pure and Applied
Mathematics Ser. 16 (Marcel Dekker, N.Y., 1973).
190. O. Zakharov, Hamiltonian formalism for nonregular Lagrangian theories in
fibered manifolds, J. Math. Phys. 33 (1992) 607.
191. Q.Zhao, The representation of Lie groups and geometric quantizations, Com-
mun. Math. Phys. 194 (1998) 135.
192. R.Zulanke and P.Wintgen, Differentialgeometrie und Faserbiindel, Hochschul-
bucher fur Mathematik, Band 75 (VEB Deutscher Verlag der Wissenschaften,
Berlin, 1972).
I ndex
acceleration
absolute, 184
relative, 184
adjoint of a form, 91
adjoint representation, 46
admissible metric, 314
almost tangent structure, 32
annihilator of a distribution, 33
antighost, 317
anti-BRST charge, 321
anti-BRST transformations, 319
atlas, 10
holonomic, 15
of constant local trivializations, 45
automorphism of fibre bundles, 12
holonomic, 53
vertical, 12
base of a fibre bundle, 10
basic De Rham cohomology, 95
basic exterior form, 95
Berry connection, 281
Berry oscillator, 281
bivector field, 29
BRST charge, 321
BRST extension of the canonical -form,
320
BRST extension of a Hamiltonian form,
320
BRST transformations, 319
bundle, 9
affine, 14
associated, 51
of principal connections, 51
principal, 48
vector, 12
dual, 13
canonical automorphism, 242
canonical complex, 91
canonical form on a frame bundle, 53
canonical homology, 91
canonical horizontal splitting, 38
canonical Liouville form, 74
canonical symplectic form, 74
canonical vector field, 67,74
canonical 3-form, 230
jet prolongation, 283
Cartan equations, 116,191
Casimir function, 67
chain complex, 89
characteristic distribution, 69
of a presymplectic form, 81
characteristic foliation of a presymplec-
tic form, 81
coadjoint representation, 47
coboundary on a group, 136
cochain complex, 90
347
348
INDEX
cocycle associated with an action of a
group, 136
cocycle on a group, 136
codistribution, 33
cofiag of a codistribution, 34
cohomology group, 90
cohomology of a group, 136
coisotropic ideal, 89
coisotropic imbedding, 83
coisotropic submanifold
of a J acobi manifold, 60
of a Poisson manifold, 70
of a symplectic manifold, 78
complete family of Hamiltonian forms,
256
composite connection, 55
composite fibre bundle, 53
configuration bundle, 6
configuration space, 3
connection, 42
affine, 44
dual, 44
dynamic, 164
symmetric, 166
complete, 157
curvature-free, 45
flat, 45
holonomic, 158
Lagrangian, 191
linear , 44
principal, 50
associated, 52
connection form, 51
local, 51
conservation law, 219
constraint non-holonomic, 216
constraint reaction acceleration, 209
constraint reaction force, 217
constraint reference frame, 214
constraint space, 124
final, 125,130
Lagrangian, 249
primary, 124
constraints
admissible, 212
complete, 125
holonomic, 206
ideal, 209
first-class, 126,132
primary, 124
second class, 126,132
secondary, 125
tertiary, 125
contact automorphism, 64
infinitesimal, 64
contact form, 37,61
contact manifold, 61
contact transformation, 63
contraction, 28
coordinates
affine, 14
adapted, 176
canonical, 70,140,243
canonically conjugate, 70
fibred, 10
holonomic , 16
induced, 16
copresymplectic structure, 193
Coriolis theorem, 186
cosymplectic structure, 193
INDEX
349
cotangent bundle, 16
covariant differential, 43
vertical, 55
current, 219
curvature, 44
D'Alembert principle, 209
generalized, 216
Darboux coordinates, 61
De Donder form, 194
democracy group, 202
De Rham cohomology group, 90
De Rham complex, 90
derivative
linear, 15
total, 36,40
variational, 187
vertical, 287
differential equation, 41
differential ideal, 34
differential operator, 41
differential system, 33
Dirac constraint system, 129
Dirac Hamiltonian system, 126
completely integrable, 128
direct product of Poisson structures,
68
distribution, 32
completely integrable, 33
involutive, 32
regular, 33
totally non-holonomic, 33
dynamic equation
autonomous, 106
conservative, 161
first order, 160,239
autonomous, 107
quadratic, 161
second order, 160
autonomous , 108
endomorphism of fibre bundles, 12
vertical, 155
energy function, 116,222
canonical, 218
Hamiltonian, 271
equilibrium equation, 244
Euler-Lagrange map, 203
Euler-Lagrange operator, 187
Euler-Lagrange-Cartan operator, 191
Euler-Lagrange-type operator, 202
even subspace, 322
evolution equation, 112,247
exterior algebra, 26
exterior differential
contravariant, 92
horizontal, 203
vertical, 155
exterior form, 25
exterior product of vector bundles, 14
exterior superform, 326
fibration, 10
fibre basis, 13
first integral of motion, 112
flag of a differential system, 33
flag of a distribution, 33
foliation, 34
of level surfaces, 35
simple, 35
singular, 35
force
350 INDEX
centrifugal, 181
Coriolis, 181
external, 198
inertial, 179
Lorentz, 198
universal, 199
frame, 13
frame bundle, 52
frame connection, 183
Hamiltonian, 237
Lagrangian, 264
free motion equation, 179
fundamental identity, 99
Galilei group, 182
gauge algebra bundle, 49
gauge conditions, 267
gauge convention, 281
gauge fields, 122
gauge freedom, 122
gauge potentials, 51
gauge transformations, 12
of a principal bundle, 271
Gauss principle, 216
general covariant transformations, 53
generalized almost J acobi manifold, 99
generalized almost Poisson manifold,
99
generalized invariant transformations,
223
generating function, 246
of a foliation, 35
of a Lagrangian submanifold, 79
generator, 21,47
geodesic equation, 109
geodesic vector field, 109
geometric quantization, 328
geometric prequantization, 328
germ of a submanifold, 79
ghost, 317
graded commutative algebra, 322
Banach, 322
graded differentiation, 325
graded Lie algebra, 25
graded manifold, 324
graded space, 322
graded S-envelope, 323
graded B-module, 323
graph of a morphism, 70
Grassman algebra, 323
Grassman bundle, 302
Hamilton equation
with respect to a symplectic struc-
ture, 114
Hamilton equations, 147,239
constrained, 259
Hamilton operator, 146,238
Hamilton-De Donder equation, 194
Hamilton-J acobi equation, 246
Hamiltonian, 234
admissible, 126
autonomous, 111
Hamiltonian action, 134
Hamiltonian connection, 147,238
locally, 144,232
Hamiltonian density, 142
Hamiltonian form, 141,234
associated with a Lagrangian, 249
constrained, 259
locally, 233
polysymplectic, 141
INDEX 351
Hamiltonian function, 235
Hamiltonian map, 115,142,235
Hamiltonian system
completely integrable, 115
gauge invariant, 122
generalized, 129
presymplectic, 119
pull-back, 121
Hamiltonian vector field
with respect to a J acobi structure,
59
with respect to a Poisson structure,
68
with respect to a presymplectic struc-
ture, 82
locally, 82
with respect to a symplectic struc-
ture, 76
locally, 76
Helmholtz-Sonin map, 203
homology group, 89
horizontal density, 27
horizontal distribution, 45
horizontal foliation, 45
horizontal form, 27
horizontal vector field, 156
infinitesimal symmetry, 112
integral curve, 20
admissible, 32
integral invariant of Poincare-Cartan,
234
integral manifold, 32
maximal, 33
of maximal dimension, 33
integral of motion, 219
integral section, 43
interior product, 28
left, 28
of vector bundles, 13
right, 28
inverse problem, 107,202
involution, 67
isomorphism of fibre bundles, 12
isotropic submanifold, 78
J acobi bracket, 57
generalized, 99
J acobi bundle, 72
J acobi field, 289
J acobi identity, 57
generalized, 99
J acobi manifold, 57
generalized, 99
J acobi morphism, 59
conformal, 60
infinitesimal, 59
J acobi structure, 57
conformaUy equivalent, 60
J acobi vectors, 202
mass-weighted, 202
jet bundle, 36
affine, 36
jet manifold, 36
higher order, 40
repeated, 38
second order, 39
sesquiholonomic, 39
jet of sections, 36
jet of submanifolds, 299
jet prolongation
of morphisms, 37
352
INDEX
of sections, 37
second order, 40
of vector fields, 37
kernel
of a differential operator, 41
of a fibred morphism, 11
of a form, 30
kinetic energy, 201
Koszul-Brylinski-Poisson homology, 91
Lagrange equations, 187
autonomous, 116
Lagrangian, 186
almost regular, 258
autonomous, 115
generalized, 81
hyperregular, 117,189
non-degenerate, 188
quadratic, 262
regular, 115,188
semiregular, 253
Lagrangian function, 186
Lagrangian submanifold, 78
of a J acobi manifold, 60
of a Poisson manifold, 70
Lagrangian system, 186
left-invariant form, 47
canonical, 47
Legendre bundle, 102,140
homogeneous, 101
Legendre map, 117,188
Legendre morphism, 193
Legendre submanifold, 63
Legendre vector bundle, 140
Leibniz rule, 66
Lichnerowicz-J acobi cohomology, 94
Lichnerowicz-Poisson cochain complex,
92
Lichnerowicz-Poisson cohomology, 92
Lie algebra
dual, 47
left, 46
right, 46
Lie bracket, 20
Lie coalgebra, 47
Lie derivative, 31
Lie-Poisson structure, 72
lift of a vector field
canonical, 22
horizontal, 43
vertical, 22
Liouville vector field, 22
mass metric, 196
mass tensor, 196
Maurer-Cartan equation, 47
momentum mapping, 134
equivariant, 135
momentum phase space, 4,108
morphism of fibre bundles, 11
affine, 15
linear, 13
motion, 160
multisymplectic form, 101,102
multivector field, 22
simple, 23
Nambu-J acobi bracket, 99
Nambu-J acobi manifold, 99
Nambu-Poisson bracket, 99
Nambu-Poisson manifold, 99
INDEX
353
Newtonian system, 197
standard, 201
Noether conservation law, 221
Noether current, 221
normalizer, 88
odd subspace, 322
parameter bundle, 274
Pfaffian system, 34
physical phase space, 122
Poincare-Cartan form, 115,187
Poisson action, 137
Poisson algebra, 66
Poisson automorphism, 67
Poisson bivector field, 58
Poisson bracket, 66
generalized, 99
Poisson bundle, 72
Poisson codifferential, 90
Poisson manifold, 66
generalized, 99
reduced, 85
Poisson morphism, 67
Poisson reduction, 85
Poisson structure, 66
coinduced, 68
generating, 69
non-degenerate, 66
regular, 66
zero, 66
polysymplectic form, 103
preconnection, 303
prequantization bundle, 329
presheaf, 18
canonical, 19
presymplectic form, 81
presymplectic manifold, 81
projection, 10
pull-back fibre bundle, 12
pull-back form, 26
rank of a bi vector field, 29
rank of a fibred morphism, 11
reduction leafwise, 87
reduction of a Poisson map, 86
reductive structure, 85
of a Poisson manifold, 85
Reeb vector field, 62
reference frame, 175
complete, 176
relativistic configuration space, 303
relativistic dynamic equation, 312
relativistic geodesic equation, 313
relativistic Hamiltonian, 309
relativistic momentum phase space, 307
relativistic principle, 307
relativistic symplectic form, 307
relativistic system, 306
relativistic velocity, 305
restriction of an exterior form, 26
restriction of a fibre bundle, 12
right-invariant form, 47
canonical, 48
Schouten-Nijenhuis bracket, 24
of exterior forms, 97
section, 10
holonomic, 37
shape coordinates, 281
shape space, 281
sheaf, 18
354 INDEX
of continuous functions, 19
soldering form, 43
solution
of a differential equation, 41
of a first order dynamic equation,
107
of a Hamiltonian system, 111
of a second order dynamic equa-
tion, 108
spray, 110
stalk, 18
standard vector field, 3
standard 1-form, 3
star operation, 91
subbundle, 11
superfunction, 324
symmetry transformations, 225
symplectic action, 133
symplectic automorphism, 74
symplectic bundle, 72
symplectic foliation, 69
symplectic form, 73
symplectic manifold, 73
reduced, 84
symplectic morphism, 73
symplectic realization, 75
symplectic reduction, 84
symplectic submanifold, 78
symplectomorphism, 74
tangent bundle, 15
tangent lift
of a function, 27
of an exterior form, 27
of a multivector field, 23
tangent map, 16
tangent-valued form, 31
canonical, 32
tensor product of vector bundles, 14
torsion of a dynamic connection, 166
total space, 10
transition functions, 10
translation-reduced configuration space,
202
trivialization, 10
typical fibre, 10
unified phase space, 283
universal unit system, 8
variational operator, 203
variational sequence, 203
vector field, 20
complete, 21
holonomic, 108
left-invariant, 46
principal, 272
projectable, 21
right-invariant, 46
subordinate, 32
vertical, 22
vector superfield, 325
vector superspace, 324
velocity
absolute, 176
relative, 176
velocity hyperboloids, 306
velocity phase space, 3,108
vertical configuration space, 285
vertical cotangent bundle, 17
vertical extension of a Hamiltonian form,
288
INDEX
355
vertical extension of a Lagrangian, 291
vertical momentum phase space, 286
vertical splitting, 17
vertical tangent bundle, 17
vertical tangent map, 17
vertical velocity phase space, 286
Whitney sum of affine bundles, 15
Whitney sum of vector bundles, 14
n-body, 281
n-symplectic form, 102
B-supermanifold, 324
1-parameter group of diffeomorphisms,
21
1-paxameter group of local diffeomor-
phisms, 20
3-velocity, 305
3-velocity phase space, 305
4-velocity, 305
4-velocity phase space, 305

Potrebbero piacerti anche