A new canonical affine BRACKET formulation of Hamiltonian Classical Field theories of first order
Abstract.
It has been a long standing question how to extend, in the finite-dimensional setting, the canonical Poisson bracket formulation from classical mechanics to classical field theories, in a completely general, intrinsic, and canonical way. In this paper, we provide an answer to this question by presenting a new completely canonical bracket formulation of Hamiltonian Classical Field Theories of first order on an arbitrary configuration bundle. It is obtained via the construction of the appropriate field-theoretic analogues of the Hamiltonian vector field and of the space of observables, via the introduction of a suitable canonical Lie algebra structure on the space of currents (the observables in field theories). This Lie algebra structure is shown to have a representation on the affine space of Hamiltonian sections, which yields an affine analogue to the Jacobi identity for our bracket. The construction is analogous to the canonical Poisson formulation of Hamiltonian systems although the nature of our formulation is linear-affine and not bilinear as the standard Poisson bracket. This is consistent with the fact that the space of currents and Hamiltonian sections are respectively, linear and affine. Our setting is illustrated with some examples including Continuum Mechanics and Yang-Mills theory.
1. Introduction
It is well-known that the Hamilton equations of classical mechanics are naturally formulated in terms of canonical geometric structures, namely, canonical symplectic forms and canonical Poisson brackets. Given the configuration manifold of the mechanical system, its momentum phase space carries a canonical symplectic form which induces a vector bundle isomorphism . Given the Hamiltonian of the system, the associated Hamilton equations of motion are determined by the Hamiltonian vector field intrinsically defined in terms of the canonical symplectic form as , i.e., we have the commutative diagram
(1.1) |
The time evolution of an observable along a solution of Hamilton’s equations, is given by
(1.2) |
where is the canonical Poisson bracket defined by
(1.3) |
Conversely, if a curve satisfies (1.2) for any observables , then is a solution of Hamilton’s equations (see, for instance, [1, 23, 43]). Recall that (1.3) defines a Lie algebra structure on , thereby satisfying the Jacobi identity
(1.4) |
The canonical Poisson formulation (1.2) of Hamilton’s equation has been at the origin of many developments in the understanding of the geometry and dynamics of Hamiltonian systems and their quantization, as well as of many generalizations. In particular, the Poisson formulation is very appropriate for developing the geometric description of the Poisson reduction of a Hamiltonian system which is invariant under the action of a symmetry Lie group (see, for instance, [49]).
When extending the geometric setting from classical mechanics to classical field theories, it is crucial to identify the geometric structures playing the role of these canonical structures. While the field-theoretic analogue of the canonical symplectic structure is well-known to be given by the canonical multisymplectic form on a space of finite dimension, the extended momentum phase space, it has been a long standing question how to extend the canonical Poisson bracket formulation from classical mechanics to classical field theories, in a completely general, intrinsic, and canonical way. Several contributions have been made in this direction as we will review below. In this paper will shall provide an answer to this question by constructing explicitly such a canonical bracket, giving its algebraic properties, and showing that it allows a canonical formulation of the Hamilton equations for field theories that naturally extends the formulation (1.2) of classical mechanics. Such canonical bracket structures could be used as a starting point of a covariant canonical quantization. One key difference between the canonical bracket that we propose and the canonical Poisson bracket of classical mechanics is its linear-affine nature. This is compatible with the fact that the set of currents and Hamiltonian sections are, respectively, linear and affine spaces for field theories. This linear-affine nature already arises in the case of time-dependent Hamiltonian Mechanics.
While in the present paper we focus on the extension to field theory of the canonical geometric setting (1.2)–(1.4) governing the evolution equations, which is of finite dimensional nature, one may also focus on the geometric structures related to the space of solutions of these equations. For a large class of field theories, this infinite dimensional space admits a Poisson (or a presymplectic) structure. This is not a new theory and we will quote an old paper by Peierls [50] (for a large list of contributions, see the recent papers [13, 14] and the references therein). In this setting, the boundary conditions of the theory play an important role [44]. In fact, in [44], Margalef-Bentabol and Villaseñor introduce the so-called “relative bicomplex framework” and develop a geometric formulation of the covariant phase space methods with boundaries, which is used to endow the space of solutions with (pre)symplectic structures. These ideas are used to discuss formulations of Palatini gravity, General Relativity and Holst theories in the presence of boundaries [4, 5, 6]. On the other hand, a classic research topic has been the relationship between finite and infinite dimensional approximation to Classical Field Theories (see Section 6).
Before explaining the main difficulties that emerge in the process of finding a canonical bracket formulation that extends (1.2)–(1.4) to classical field theories, we quickly review below the previous contributions to the geometric formulation of Hamiltonian Classical Field Theories.
1.1. Previous contributions on the geometric formulation of Hamiltonian classical field theories of first order.
The geometric description starts with the choice of the configuration bundle , whose sections are the fields of the theory. The case of time-dependent mechanics corresponds to the special situation . For classical field theories, there are two useful generalisations of the notion of momentum phase space: the restricted multi-momentum bundle and the extended multi-momentum bundle . Both are vector bundles on with vector bundle projections denoted and , and there is a canonical line bundle projection . The main property of the extended multi-momentum bundle is that it admits a canonical multisymplectic structure of degree ( being the dimension of ). The restricted multi-momentum bundle , however, does not admit a canonical multisymplectic structure.
An important difference with Hamiltonian Mechanics is that for Hamiltonian Classical Field Theories we don’t have a Hamiltonian function, but a Hamiltonian section of the canonical projection , see [11]. The corresponding evolution equations are the Hamilton-deDonder-Weyl equations. They form a system of partial differential equations of first order on with space of parameters and whose solutions are sections of the projection . More precisely, if we denote by
the local expression of the Hamiltonian section , then the Hamilton-deDonder-Weyl equations are locally
(1.5) | ||||
These equations go back, at least, to work of Volterra [52, 53]. In the literature, we can find the following geometric descriptions of the Hamilton-deDonder-Weyl equations (1.5) associated to :
-
•
From and the canonical multisymplectic structure one can produce a non-canonical multisymplectic structure on . Then, using , a special type of Ehresmann connections can be introduced on the fibration (which, in the present paper, will be called Hamiltonian connections for ) and the solutions of the evolution equations are the integral sections of these connections (see [19, 21]; see also [25, 26]). The multisymplectic structure can also be used to directly characterize the sections of the projection which are solutions of the Hamilton-deDonder-Weyl equations for (see [11, 19, 21, 25, 26]). From , one can also define the ”multisymplectic pseudo-brackets” and ”multisymplectic brackets” of -Hamiltonian forms which may be considered the field version of the Poisson bracket for functions in Classical Mechanics (see [28]).
-
•
Using an auxiliary Ehresmann connection on the configuration bundle and the Hamiltonian section , one may produce a Hamiltonian energy associated to and and a non-canonical multisymplectic structure on which allow us to describe the solutions of the evolution equations in a geometric form (see [11]; see also [25, 26]). In addition, it is possible to consider a suitable space of currents (a vector subspace of -forms on which are horizontal with respect to the projection ) and one may introduce a “Poisson bracket” of a current and an Hamiltonian energy associated to . This bracket allows the description of the Hamilton-deDonder-Weyl equations as in Hamiltonian Mechanics (see [12]).
-
•
The Hamiltonian section induces a canonical extended Hamiltonian density , which is a smooth -valued function defined on see [10, 32]; see also [24] for the particular case when a volume form on is fixed. Then, using and the canonical multisymplectic structure one write intrinsically a system of partial differential equations on whose solutions are sections of the fibration . The projection, via , of these sections are the solutions of the Hamilton-deDonder-Weyl equations for (see [24]).
-
•
From the configuration bundle one can construct the phase bundle , an affine bundle over , and the differential of the Hamiltonian section , as a section of . In addition, an affine bundle epimorphism from the -jet bundle of the fibration onto may be also introduced. Then, the solutions of the Hamilton-deDonder-Weyl equations are the sections whose first prolongation is contained in the submanifold (see [32, 33]; see also [35, 47] for the particular case of time-dependent Hamiltonian Mechanics).
1.2. The problem
The previous comments lead naturally to the following question:
Does there exist a completely canonical geometric formulation of the Hamilton-deDonder-Weyl equations which is analogous to the standard Poisson bracket formulation of time-independent Hamiltonian Mechanics?
A possible answer to this question could be the geometric formulation developed in [12] (see Section 1.1). However, this formulation is not canonical since most of the constructions in [12] depend on the chosen auxiliary connection in the configuration bundle. In fact, in a previous paper [46] Marsden and Shkoller justify the use of this connection in the geometric formulation of the theory and one may find, in that paper (see [46], page 554), the following cite:
It is interesting that the structure of connection is not necessary to intrinsically define the Lagrangian formalism (as shown in the preceding references), while for the intrinsic definition of a covariant Hamiltonian the introduction of such a structure is essential. Of course, one can avoid a connection if one is willing to confine ones attention to local coordinates.
However, in our paper, we will construct a bracket that does not use any auxiliary objects such as a connection in the configuration bundle and which is completely canonical, thereby giving an affirmative answer to the question above.
1.3. Answer to the problem and contributions of the paper
In order to give an affirmative answer to the question in Section 1.2, we will use the following previous contributions and results:
-
•
The construction of the phase space associated with the configuration bundle and the differential of a Hamiltonian section as a section of (see [32]).
-
•
The affine bundle epimorphism which was also introduced in [32].
-
•
The notion of a Hamiltonian connection associated with a Hamiltonian section . This type of objects were already considered in [19, 21] in order to characterize the solutions of the Hamilton-deDonder-Weyl equations for (although the authors of these papers did not use the terminology of a Hamiltonian connection).
We will combine the previous constructions as follows.
As a first step, we consider the affine bundle isomorphism
where is the kernel of the affine bundle epimorphism and , with the affine bundle isomorphism induced by . Then, we introduce the section
of , given by
and we prove the following result: the section is canonically identified to the equivalence class of Ehresmann connections on the fibration that are Hamiltonian connections for , see Theorem 3.8.
So, the section associated to the Hamiltonian section is the field-theoretic analogue to the Hamiltonian vector field associated to a Hamiltonian function in Classical Mechanics. The following commutative diagram illustrates the situation
The analogy with the corresponding diagram given in 1.1 for classical mechanics is evident.
The next step is to introduce a suitable space of currents (a vector subspace of -forms on which are horizontal with respect to the fibration ), in such a way that the restriction of the standard exterior differential to takes values in the space of sections of the vector bundle , that is, we have the linear map
The dual vector bundle is chosen so that these differentials can be canonically paired with the Hamiltonian connections , thereby extending to the field-theoretic context the pairing between the differential of an observable and the Hamiltonian vector field , see (1.3). This is our motivation for introducing the space of currents and although it is different to the motivation in [12], just coincides with the space of currents in [12] (see Remark 3.14).
Now, if is the space of Hamiltonian sections, we can define the linear-affine canonical bracket
(1.6) |
given by
Then, one may prove that the evolution of any current along a solution of the Hamilton-deDonder-Weyl equations for is given by
(1.7) |
Conversely, if is such that (1.7) holds for all , then is a solution of Hamilton-deDonder-Weyl equations. The canonical bracket formulation (1.7) is the field-theoretic analogue to the canonical Poisson bracket formulation (1.2) of classical mechanics.
The previous tasks are performed in Section 3.4 (see Theorem 3.15). Here again, the analogy with the canonical Poisson formulation of classical mechanics (see (1.2) and (1.3)) is evident.
It is important to note the affine character of the canonical bracket in (1.6): the space of Hamiltonian sections is an affine space modelled over the vector space . Recalling that the canonical Poisson bracket on induces a Lie algebra structure on , a new question arises:
What are the algebraic properties of the bracket ?
Related to this question, we will prove that admits a canonical Lie algebra structure (see Theorem 4.2) and that the linear map
defined by
is a representation of the Lie algebra on the affine space (see Theorem 4.3 and the property (4.13)).
The previous results will be applied to the following examples: time-dependent Hamiltonian systems, Continuum Mechanics (including fluid dynamics and nonlinear elasticity) and Yang-Mills theories. Some of the constructions developed in the paper are illustrated in the Diagram in Appendix §C.
1.4. Structure of the paper
The paper is structured as follows. In Section 2, we review the geometric formulation of the Hamilton-deDonder-Weyl equations using the multisymplectic structure on the phase space induced by the Hamiltonian section. In Section 3, we introduce the canonical linear-affine bracket and we formulate the Hamilton-deDonder-Weyl equations using this bracket. In particular, we describe the evolution of a current along a solution of the Hamilton-deDonder-Weyl equations. In Section 4, we introduce a Lie algebra structure on and we prove that induces a representation of the Lie algebra on the affine space of Hamiltonian sections. In Section 5, we apply the previous results to several examples. The paper closes with three appendices. In the first one, we review the definition of the -jet bundle associated with a fibration, in the second one, we discuss the vertical lift of a section of a vector bundle as a vertical vector field on the total space and, in the third one, we present a Diagram which illustrates most of the relevant constructions in the paper.
2. Hamiltonian Classical Field Theories of first order
In this section, we review some basic constructions and results on Hamiltonian Classical Field Theories of first order (for more details, see [11]).
2.1. The restricted and extended multimomentum bundle associated with a fibration
The configuration bundle of a classical field theory is a fibration , that is, a surjective submersion from to . We assume and .
The extended multimomentum bundle associated with the configuration bundle is the vector bundle over whose fiber at the point is
Here, is the -jet bundle of the fibration (see Appendix A).
It is well-known that may be identified with the vector bundle over , whose fiber at is
In fact, if and then
(2.1) |
If are local coordinates on which are adapted with the fibration , then reads locally
where
So, are local coordinates on .
On we can define a canonical -form as follows
(2.2) |
for and , with the vector bundle projection.
From (2.2), has the local expression
The canonical multisymplectic structure on is the -form given by
Locally, we have
(2.3) |
It is clear that is closed and non-degenerate, that is, the vector bundle morphism
is a linear monomorphism.
The restricted multimomentum bundle is the vector bundle
Local coordinates on are .
There is a canonical projection given by
for , where is the linear map associated with . The local expression of is
Note that if then
(2.4) |
Remark 2.1.
Note that this last statement implies that . This situation recalls a particular case in the Poisson realm, the quotient of a symplectic manifold by a proper and free action of a symmetry Lie group inherits a Poisson structure. In this formalism, the quotient of the extended multimomentum bundle also has a new version of a multi-Poisson structure (see for example [8]) which is defined via a Lie algebroid structure on a subbundle of , when the base manifold is orientable. Note that, in such a case, if we fix a volume form on , we have an action of the real line on , which preserves the multisymplectic structure, and is the space of orbits of this action. For the definition and details on the construction of the multi-Poisson structure, we refer to [9].
2.2. Hamilton-deDonder-Weyl equations
Given a configuration bundle , a Hamiltonian section is a smooth section of the canonical projection
The local expression of is
where is a local real -function on .
Using the Hamiltonian section, we can define the -form on given by
(2.5) |
The local expression of is
(2.6) |
Note that if , is degenerate and the rank of its kernel is . On the other hand, if , is non-degenerate which implies that it is multisymplectic.
Proposition 2.2.
A (local) section of the projection is a solution of the Hamilton-deDonder-Weyl equations iff
where is the space of sections of the vertical bundle to .
Proof.
Using the local expression , it is routine to verify that if and only if
∎
3. A new canonical bracket formulation of Hamiltonian Classical Field Theories of first order
As we reviewed in the Introduction, the phase space of momenta for classical mechanics is the cotangent bundle of the configuration space , a smooth manifold of dimension . The cotangent bundle carries a canonical symplectic structure which induces a vector bundle isomorphism over the identity with inverse denoted . The Hamiltonian is a real -function on and the Hamiltonian vector field is given in terms of the differential and the vector bundle isomorphism as , see Diagram (1.1). For field theories, we don’t have a Hamiltonian function, but a Hamiltonian section
of the canonical projection . So, the following questions arise when extending the previous construction to field theories:
Question 1: What is the differential of ?
Question 2: Where does the differential of take values?
We will answer these questions in §3.1 by showing that the differential of is a section of the phase bundle associated with the fibration . The bundle was introduced in [32] and was used there to discuss a Tulczyjew triple for Classical Field Theories of first order. This will allow us to define the field-theoretic analogue to the vector bundle isomorphism in §3.2 and the field-theoretic analogue to the Hamiltonian vector field in §3.3. In particular, we will show that can be identified with the equivalence class of Hamiltonian Ehresmann connections associated to .
Going back to Classical Hamiltonian Mechanics, we recall that the set of observables is the space and that the Hamilton equations can be equivalently formulated in the Poisson bracket form (1.2) with respect to the canonical Poisson bracket giving by the formulas (1.3). In view of this formulation, we need to find a suitable space of currents for field theories (the observables in field theories) such that their differentials take values in a bundle dual to the target bundle of , this is the goal of §3.4. From this a canonical bracket can be obtained between currents and Hamiltonian sections. This construction is carried out in §3.5.
3.1. The phase bundle associated with a fibration and the differential of a Hamiltonian section
Let be the configuration bundle of the field theory and be the Hamiltonian section. Then, although
cannot be identified, in general, with a real -function on . However, to we can associate an extended Hamiltonian density
defined as follows. If we have and hence using (2.4), we conclude that there exists a unique such that . We thus define
3.1.1. The differential of and the extended phase bundle
Note that may be considered, in a natural way, as a -form on . Thus, we can take its exterior differential and we obtain a -form on which is a section of the vector bundle
Now, it is easy to prove that the vector bundles and are isomorphic. In fact, an isomorphism
is given by
for and . Note that , therefore, it induces an element of .
We denote by the section of the vector bundle induced by the differential of . In local coordinates, if
then
(3.1) |
and
(3.2) |
Note that if is the canonical projection, is a -form on and is the vertical lift to (see Appendix B) then, using (3.2) and (B.1), we deduce that
(3.3) |
This property of motivates the definition of the following affine subbundle of .
Definition 3.1.
The extended phase bundle of the configuration bundle is the affine subbundle of whose fiber at the point is
(3.4) |
From (3.3) we have
Note that is modelled over the vector bundle whose fiber at the point is
We remark that an element of has the following local form
and a generic element of has the local form
Therefore, the local coordinates on and are . In addition,
(3.5) |
3.1.2. The differential of a Hamiltonian section and the phase bundle
Now, given a point , we can consider an action of the abelian group on the fiber defined as follows. If then we define by
(3.6) |
In local coordinates, we get
and thus the quotient space may be identified with the reduced multimomentum bundle .
The tangent and cotangent lift of the previous action induces a fibred action of the vector bundle on the vector bundles and . In fact, if , and then the tangent lift is
(3.7) |
Note that, using (3.6), (3.7) and (B.1), it follows that
(3.8) |
for . If are local coordinates on , we have
(3.9) |
In a similar way, if then the cotangent lift is
(3.10) |
for . From (3.4) and (3.8), we deduce that this action restricts to the extended phase bundle , and to the vector bundle . In local coordinates, we have
Taking the quotient with respect to the action, we can introduce the following definition.
Definition 3.2.
The phase bundle of the configuration bundle is defined by
We note that is an affine bundle over modelled over the vector bundle . This bundle is isomorphic to the vector bundle
an isomorphism being given by
where is defined by
with and . Local coordinates on and are
(3.11) |
It is clear that there exists a one-to-one correspondence between the space of sections of the affine bundle and the set of sections of the extended phase bundle associated with , which are -equivariant. So, if is a Hamiltonian section then, using (3.5), it is easy to see that the vertical differential is -equivariant and, therefore, it induces a section
of the phase bundle . We can thus write the following definition.
Definition 3.3.
The differential of a Hamiltonian section is the section
defined by the following commutative diagram
where and are the canonical projections.
The local expression of is
(3.12) |
So, we have given an answer to Questions 1 and 2 stated above.
From the previous definition, we get the map
Note that and are affine spaces modelled over the vector spaces and , respectively, and is an affine map. Later in the paper, we shall use the corresponding linear map defined as follows. If is a -valued function on then it may be considered as a section of the vector bundle
So, we can take the standard differential and we obtain a section of the vector bundle
This vector bundle is isomorphic to , an isomorphism
is given by
for and . Note that and, therefore, it induces an element of . We denote by the section of the vector bundle induced by the differential via the isomorphism . If locally
the local expression of is
(3.13) |
3.1.3. Comments on the next steps
The differential of a Hamiltonian section is the field theoretic analogue to the differential of a Hamiltonian function in classical mechanics. In the next section, we will introduce a quotient affine bundle which is the field theoretic analogue to the tangent bundle of the phase space in classical mechanics. Recall that using the canonical symplectic structure of , one can define a canonical vector bundle isomorphism
and the Hamiltonian vector field on associated with a Hamiltonian function is given by . So, a natural question arises:
Question 3: Does there exist an affine bundle isomorphism which, in the presence of a Hamiltonian section , allows us to introduce a distinguished section of the affine bundle ?
In the next Section 3.2, we will give an affirmative answer to Question 3 and we will discuss the relation between and the solutions of the Hamilton-deDonder-Weyl equations for . The section will play the role of in Hamiltonian Mechanics.
3.2. The field-theoretic analogue to the canonical isomorphism
We will show that it is given by an affine bundle isomorphism .
Let be the -jet bundle associated with the fibration (see Appendix A). To define the quotient affine bundle, we shall use a construction in [32]. In this paper, the author introduced an affine bundle epimorphism
over the identity of . This epimorphism is constructed in several steps.
First, we consider the vector bundle monomorphism
induced by the canonical multisymplectic structure as follows
(3.14) |
Note that, from (2.3), there exists a unique -form at on , which we denote by , such that
(3.15) |
This defines a vector bundle morphism
over the canonical projection . Using local coordinates and on and , respectively, the local expression of is
(3.16) |
If , the following commutative diagram
(3.17) |
illustrates the relation between and .
We shall now use the vector bundle morphism to construct . For and one first defines
with such that , as follows:
(3.18) |
for . Then, if is the canonical projection, we set
(3.19) |
Note that is well-defined and its local expression is
(3.20) |
This proves that is an affine bundle epimorphism over the identity of (for more details, see [32]).
Recall that the affine bundle is modelled over the vector bundle . We denote by the standard coordinates on and (see Appendix A). From the local expression (3.20), it follows that the kernel of is a vector subbundle of which is locally characterized by
(3.21) |
We can thus consider the quotient affine bundle which is modelled over the quotient vector bundle . From (3.21), we have that a local basis of sections for this vector bundle is
for and . Note that in the quotient vector bundle
Local coordinates associated to this basis of sections on the quotient vector bundle (and also on the quotient affine bundle ) are denoted
The affine bundle epimorphism induces an affine bundle isomorphism
and, from (3.20), we deduce that the local expression of is
(3.22) |
By definition, the affine bundle isomorphism
is the inverse isomorphism to . If we consider the local coordinates on the phase bundle then, using (3.22), it follows that
(3.23) |
3.3. The field-theoretic analogue to the Hamiltonian vector field
Let be a Hamiltonian section. We have seen that the differential of
is a section of the phase bundle . So, we can define the section
of the quotient affine bundle by
(3.24) |
Using (3.12), (3.23) and the local basis of sections introduced above, we obtain that the local expression of is
(3.25) |
Now, we will show that plays the same role, in Hamiltonian Classical Field Theories of first order, that the Hamiltonian vector field associated with a Hamiltonian function in Classical Mechanics. This will give an affirmative answer to Question 3 in Section 3.1.
For this purpose, we will discuss the relation between and the solutions of the Hamilton-deDonder-Weyl equations for . This uses the notion of a Hamiltonian connection (see [19, 21]).
Let be an arbitrary fibration and an Ehresmann connection on . Denote by the horizontal lift induced by (see Appendix A). It is clear that if the previous map induces a vector bundle isomorphism between and which we also denote by
So, if , the image of by the previous map is called the horizontal lift of . If are local coordinates on which are adapted to the fibration , then the horizontal lift reads locally
and if
then its horizontal lift is
Now, suppose that is the configuration bundle of a Hamiltonian Classical Field Theory of first order, with , and consider the fibration . For a Hamiltonian section , we denote by the -form on given by (2.5).
Then, we may prove the following result.
Lemma 3.4.
Let be a Hamiltonian section, let be an Ehresmann connection on the fibration , and let be a section horizontal with respect to . Then, is a solution of the Hamilton-deDonder-Weyl equations for if and only if
(3.26) |
Proof.
Using that is horizontal with respect to we deduce that
if and only if
(3.27) |
So, it is clear that if (3.26) holds then is a solution of the Hamilton-deDonder-Weyl equations for by Proposition 2.2.
Conversely, assume that is a solution of the Hamilton-deDonder-Weyl equations for . Then (3.27) holds. Thus, using that
we deduce that
This proves the result. ∎
The previous result suggests the introduction of the following definition.
Definition 3.5.
Let be a Hamiltonian section. An Ehresmann connection on the fibration is said to be a Hamiltonian connection for if
As a direct consequence of the definition, if locally the Hamiltonian section is
and the Ehresmann connection is
then from (2.6), we have the equivalence
(3.28) |
So, our definition of a Hamiltonian connection is equivalent to that introduced in [19, 21]. Note that a Hamiltonian connection for may be identified with a section of the affine bundle (see Appendix A). Moreover, if is the section of the quotient affine bundle defined in (3.24) then, using (3.25) and (3.28), we obtain the following result.
Proposition 3.6.
Let be the section of the affine bundle induced by a Hamiltonian connection for and be the canonical projection. Then
Proposition 3.7.
If are Hamiltonian connections for the same Hamiltonian section , then they satisfy
Finally, from Propositions 3.6 and 3.7, it follows the following characterization of Hamiltonian connections for .
Theorem 3.8.
Let be a Hamiltonian section and let be an Ehresmann connection for the fibration . Then, is a Hamiltonian connection for if and only if
Theorem 3.8 suggests the introduction of the following definition.
Definition 3.9.
Let be a Hamiltonian section. Then, the section is called the equivalence class of the Hamiltonian connections for .
The following commutative diagram illustrates the results obtained in Sections 3.1, 3.2, 3.3
It is the field-theoretic analogue to Diagram 1.1 for Hamiltonian Mechanics.
The last step is to introduce a suitable space of currents for Hamiltonian Classical Field Theories of first order and a suitable canonical bracket formulation for the evolution of such currents along the solution of the Hamilton-deDonder-Weyl equations. This is the aim of the next two subsections.
3.4. A suitable space of currents for Hamiltonian Classical Field Theories
We shall define a space of currents for Hamiltonian Classical Field Theories of first order, which plays the same role that the space of observables in Hamiltonian Mechanics.
Recall that in Hamiltonian Mechanics, the Hamiltonian vector field is a section of the vector bundle and the space of observables is the set of real -functions on . Given an observable , we can consider a section (the differential of ) of the dual bundle to and the evolution of the observable along a solution of Hamilton’s equation is given as
When written for all observables , the previous equations are equivalent to the Hamilton equations.
Our goal is to carry out these construction for Hamiltonian Classical Field theories. As we have seen, given a Hamiltonian section , the object corresponding to the Hamiltonian vector field is the section of the quotient affine bundle . So, we need to overcome the following two steps:
First step: Describe the dual vector bundle to the affine bundle .
Second step: Introduce a space of currents and a differential operator
(3.29) |
on this space, such that the evolution of a current along a solution is given by
We will show that satisfies these equations for any if and only if is a solution of the Hamilton-deDonder-Weyl equations.
First step: Let (respectively, ) be the affine bundle epimorphism (respectively, isomorphism) considered in Section 3.2. Denote by and the vector bundles over defined by
It is clear that induces the vector bundle morphism
for and , with . Since is an epimorphism, we deduce that is a vector bundle monomorphism. In addition, the image of is the vector subbundle of whose fiber at the point is
Here, denotes the linear map associated with the affine map . So, we have a vector bundle isomorphism
over the identity of .
Now, denote by the vector bundle isomorphism induced by . Then, it is clear that the vector bundles and can be identified and, under this identification, is just the vector bundle isomorphism .
The following commutative diagram
illustrates the situation.
It is desirable to have an explicit realisation of dual vector bundle . As we know
(see Section 2.1). So it is possible to describe and, therefore, , as a certain vector subbundle of . We shall now give such a description.
First of all, using (3.4), it follows that the vector bundle
is isomorphic to the vertical bundle of the fibration . An isomorphism
is given by
(3.30) |
with .
Now, if we consider the standard fibred actions of on and then, it is clear that is -equivariant and, thus, it induces a vector bundle isomorphism
over the identity of . Then, from Definition 3.2, we deduce that the quotient vector bundle is isomorphic to . So, we have a vector bundle isomorphism
which is characterized by the following condition
(3.31) |
for and , with and the canonical projection.
We now consider the composition
of the two vector bundles isomorphisms and defined above and show that can be expressed in a simple way, which allows to describe its image explicitly.
Consider the vector bundle morphism defined in Section 3.2 which is characterized by Eq. (3.15) and has the local expression (3.16). Using (3.9), we deduce that induces the vector bundle morphism
over the identity of given by
(3.32) |
So, if and is the canonical isomorphism between the fibers by and of the vector bundles and , then the following diagram
(3.33) |
is commutative
Proposition 3.10.
We have the equality
(3.34) |
This implies that the vector bundle is isomorphic to the vector subbundle of given by
(3.35) |
In particular, a local basis of sections of the vector subbundle is
(3.36) |
Proof.
Second step: The previous result together with (3.29) suggests the introduction of the following definition.
Definition 3.11.
The space of currents of a Hamiltonian Field Theory with configuration bundle is
Example 3.12.
i) Let be a -form on which is semi-basic with respect to the projection . Then,
Indeed, for , we have
which implies that , see (3.36).
ii) Let be a section of the vector bundle , that is, is a vector field on and
Define the -form on as follows
for and . If the local expression of is
it follows that . Thus,
which implies that and .
In the following theorem, we give the explicit description of the currents for the case when . Note that if then .
Theorem 3.13.
If then a section of the vector bundle is a current if and only if there exists a unique -semibasic -form on and a unique -vertical vector field on such that
Proof.
It is clear that if is a -vertical vector field and is a -semibasic -form on then is a current (see Examples 3.12).
Conversely, suppose that is a current. The local expression of is and
Thus, using (3.36), we deduce that
(3.37) |
This implies that
so if we conclude that
(3.38) |
Therefore, from (3.37) and (3.38), it follows that
Consequently, we have proved that there exists a local -vertical vector field and a local -semibasic -form on such that
Note that and are unique. Then, this last fact also proves the global result. ∎
Remark 3.14.
(i) Note that is a -module.
(ii)
In [12], the authors consider as a space of currents the set of horizontal Poisson -forms on . Moreover, they prove that a -form of this type may be described as
where is a vertical vector field on , is a -semibasic -form on and is a closed -semibasic -form on . Now, it is easy to prove that, under the previous conditions, there exists a unique closed -form on such that . So, if we take , we conclude that
The previous discussion shows that is just the space of currents which was considered in [12].
3.5. A suitable linear-affine bracket and the Hamilton-deDonder-Weyl equations
We consider the linear-affine bracket
defined by
(3.39) |
Assume that , that the local expression of the Hamiltonian section is
and that the local expression of the current is
with and local real -functions on . Then, using (3.25) and (3.39), we obtain the local expression of the linear-affine bracket (3.39) as
(3.40) |
Note that if we write the current as , with , the bracket takes the elegant form
As we know, is an affine space which is modelled over the vector space . Therefore, using (3.39), it follows that the bilinear bracket
associated with the linear-affine bracket is given by
Here, is the vertical differential of (see (3.13)),
is the vector bundle isomorphism associated with the affine bundle isomorphism , and
is the canonical projection. In local coordinates, we have
(3.41) |
Again, if we write the current as , with , the bilinear bracket takes the form
On the other hand, if then the space of currents is
and, using (3.25) and (3.39), we deduce that the linear-affine bracket
and the bilinear bracket
are locally given by
(3.42) |
and
(3.43) |
for and .
The following result extends to the field-theoretic context the canonical Poisson bracket formulation of Hamilton’s equations.
Theorem 3.15.
Let be a Hamiltonian section and a (local) section of the projection . Then, is a solution of the Hamilton-deDonder-Weyl equations for if and only if
(3.44) |
3.6. A remark on boundary conditions
If the boundary of the base space of the configuration bundle is not empty, then the Hamilton-deDonder-Weyl equations for a Hamiltonian section can be supplemented by boundary conditions.
The boundary of the configuration space is just
in such a way that
is again a fibration.
In a similar way, the restricted multimomentum bundle is a manifold with boundary,
and we have fibrations
A boundary condition for the Hamiltonian Classical Field theory is given by specifying a subbundle of , such that is a subbundle of . In such a case, we will consider only sections such that
(3.45) |
A standard assumption in the literature for the subbundle is
where is the canonical inclusion and is the Hamiltonian section (see, for instance, [2, 22, 39, 40]; see also [7] for boundary conditions in the Lagrangian formalism).
From (3.45), we deduce that among all the Hamiltonian connections
we should only consider those whose restriction to takes values in the tangent bundle , that is, H should induce a monomorphism of vector bundles
This remark is sufficient for the purposes in this paper.
A more detailed discussion of boundary conditions for a Hamiltonian Classical Field theory of first order and its relation with the section of the quotient affine bundle and with the theory of covariant Peierls brackets [50] in the space of the solutions will be postponed to a future publication (see the next Section 6).
4. The affine representation of the Lie algebra of currents on the affine space of Hamiltonian sections
In this section, we prove that the space of currents of a Hamiltonian Field theory of first order admits a Lie algebra structure and we show that the linear affine bracket introduced in Section 3.5 (see (3.39)) induces an affine representation of on the affine space of Hamiltonian sections.
We first review the notion of an affine representation of a Lie algebra on an affine space (for more details, see [38]).
Let be an affine space modelled over the vector space . The vector space of affine maps of on , , is a Lie algebra and the Lie bracket on is given by
for , where are the linear maps associated with , respectively.
An affine representation of a real Lie algebra on is a Lie algebra morphism
We first note that if is a configuration bundle with then it is easy to prove the following facts (see Section 5.1 for the particular case when is the real line and ):
-
•
is an affine subbundle of corank of the tangent bundle which is modelled over the vertical bundle to the fibration .
-
•
The restricted multimomentum bundle is just the dual bundle to .
-
•
The extended multimomentum bundle is the cotangent bundle of .
-
•
is an affine space which is modelled over the vector space of the currents.
So, in this case, we have a Lie algebra structure on . In fact, the Lie bracket on is just the Poisson bracket on given by (3.43). Moreover, using (3.42) and (3.43), we deduce that the linear-affine bracket
induces an affine representation of the Lie algebra on the affine space . More explicitly, we have
for and .
Therefore, in the rest of this section, we will assume the following hypothesis:
Assumption: In what follows, we will suppose that .
First we introduce a Lie algebra structure on , then we show that the linear affine bracket induces an affine representation of the Lie algebra on the affine space .
4.1. The Lie algebra structure on the space of currents
The construction of the Lie bracket is made in several steps which involve the definition of a vertical vector field on associated to a current.
4.1.1. Definition of the vertical vector field on associated to a current
Let be the vector bundle isomorphism over the identity of given by (3.32) and denote by the inverse morphism. If then, from Definition 3.11, we have and . So, we can consider the section of the vector bundle .
Remark 4.1.
The notation is justified by the following fact. The vector bundle is canonically isomorphic to (an isomorphism between these vector bundles is characterized by condition (3.31)). So, if is a Hamiltonian section then and
Moreover, as will be proved later (see the next Lemma 4.5), we can write the linear-affine bracket as
The reader can compare the previous expression with Eq. (1.3) for the definition of the canonical Poisson bracket on .
Note that a section of the vector bundle can be identified with a section of which is equivariant with respect to the fibred actions of on and on .
By applying this observation to , we denote by
the equivariant vector field on associated with the section . Since is equivariant, it follows that it is -projectable to a vertical vector field
on .
We now present a description of in terms of . Let be the vector bundle monomorphism given by (3.14). Denote by the image of by , so that
is a vector bundle isomorphism over the identity of . From (2.3), it follows that a local basis of is
Note that if then
In fact, if is a section of then
(4.1) |
Let be the vector bundle morphism over the canonical projection which is characterized by Eq. (3.15). If and is the canonical isomorphism between the fibers by and of the vector bundles and then, using (3.17) and (3.33), it follows that the following diagram
(4.2) |
is commutative. This implies that
or, in other words, satisfies the following condition
(4.3) |
Note that, since is non-degenerate, (4.3) may be considered as a definition of the equivariant vector field .
So, in conclusion, for a current we have the following objects:
-
•
An equivariant vertical vector field on , which is characterized by condition (4.3).
-
•
The induced section of the vector bundle
-
•
The vertical vector field on
which is the projection, via , of .
We now present the local expressions of the vector fields and . As we have seen (see Section 3.4), the local expression of an element is
with and local real -functions on . We have
(4.4) |
Thus, using (2.3) and (4.3), we deduce that
(4.5) |
Therefore, it follows that
(4.6) |
Following the terminology in [11] (see also [31]), Eq. (4.3) implies that is a Hamiltonian -form and that the vector field is a Hamiltonian vector field on the multisymplectic manifold .
4.1.2. Definition of the Lie bracket
As in [11], we consider the -form on the multisymplectic manifold defined by
Using the fact that is -projectable, it follows that such a -form is basic with respect to . In fact,
This equation suggests the introduction of the the -form on given by
(4.7) |
Using the local expressions
we deduce from (4.4) and (4.6) that
(4.8) |
So, it is clear that .
Moreover, we can prove the following result.
Theorem 4.2.
The bracket given by
defines a Lie algebra structure on the space of currents .
Proof.
Using Theorem 3.13, we deduce that there exists and isomorphism between the -modules and . In addition, from (4.8), it follows that under the previous isomorphism, the bracket is given by
for , where is the Lie bracket of vector fields in and is the Lie derivative operator. Using this definition of , it is easy to prove that induces a Lie algebra structure on . ∎
Note that if we write the observables locally as and with and , then the Lie bracket has the local expression
which is reminiscent of the local expression of the canonical Poisson bracket.
As a final remark on the definition of the morphism and the bracket on the currents, we can derive a version of the classical result that any Poisson structure on a manifold induces a Lie algebroid structure on the cotangent bundle of the manifold. In this case, we will obtain a Lie algebroid structure on the vector bundle over .
Indeed, it is clear that the vector bundle admits a Lie algebroid structure. The Lie bracket in the space of sections is just the restriction of the standard Lie bracket to -vertical vector fields and the anchor map is the inclusion . So, using the vector bundle isomorphism , we can induce a Lie algebroid structure on the vector bundle . In fact, a direct computation proves that the Lie bracket in the space of sections of , , is given by
(4.9) |
where is the anchor map.
On the other hand, the space of sections of the vector bundle may be identified with the -vertical and -equivariant vector fields on . Thus, using that the standard Lie bracket of vector fields is closed for this subspace, we may induce a Lie algebroid structure on the vector bundle . Note that every -vertical and -equivariant vector field on is -projectable a vector field on and this fact determines the anchor map of the Lie algebroid .
Now, using the vector bundle isomorphism
we can translate the previous Lie algebroid structure to a Lie algebroid structure on . In fact, if then, from the commutativity of the diagram (4.2), it follows that vector field on is -projectable to a vector field . This fact determines the anchor map of the Lie algebroid . Moreover, from (4.9), we deduce that the Lie bracket on is given by
(4.10) |
Finally, if then, using (4.7), (4.10) and the fact that
we conclude that
(4.11) |
Eqs (4.10) and (4.11) are reminiscent of the properties of the Lie algebroid structure on the cotangent bundle of a Poisson manifold (in particular, the cotangent bundle of an arbitrary manifold).
4.2. The linear-affine bracket is an affine representation
In this subsection, we show that the linear-affine bracket induces an affine representation of the Lie algebra of currents on the affine space of Hamiltonian sections.
Theorem 4.3.
The map
(4.12) |
is an affine representation of the Lie algebra on the affine space .
More explicitly, we have
(4.13) |
In order to prove this theorem, we will use the following results.
Lemma 4.4.
The linear map
is a Lie algebra anti-morphism between the Lie algebras and , where is the standard Lie bracket of vector fields.
Lemma 4.5.
Let be an element of .
-
(i)
If is a Hamiltonian section then
(4.14) where is the extended Hamiltonian density associated with and is the vertical differential of . So, we have
(4.15) -
(ii)
If then
(4.16) where is the linear part of the differential of .
Remark 4.6.
Note that and, thus, . In fact,
Proof.
(of Theorem 4.3) It is clear that is a linear map. So, we must prove that is a Lie algebra morphism. For this purpose, we will use Lemmas 4.4 and 4.5.
If are currents and is a Hamiltonian section then, using (4.12) and (4.14), we obtain
Therefore, from Lemma 4.4, it follows that
Now, using (4.14) and the fact the vector fields and are -projectable on the vector fields and , respectively, we deduce
So, from (4.16), we obtain that
Finally, using (4.12) and the fact that
we conclude that
which proves the result. ∎
Proof.
5. Examples
5.1. Time-dependent Hamiltonian Mechanics
In this section, we will use the following terminology. Let be a vector bundle. Then, we can consider the vector bundle
The sections of this vector bundle are just the time-dependent sections of . For this reason, the vector bundle will be called the time-dependent vector bundle associated with .
For time-dependent Mechanics, the base space of the configuration bundle is the real line , that is, we have a fibration . This fibration is trivializable but not canonically trivializable. In fact, if one choses a reference frame one may trivialize the fibration. This means that may be identified with a global product and, under this identification, is the canonical projection on the first factor. For simplicity, in what follows, we will assume that this is our starting point although all the constructions in this section may be extended, in a natural way, if we don’t chose a reference frame (for an affine formulation of frame-independent Mechanics, we remit to [34, 35, 36, 37, 41, 47].)
So, if the configuration bundle is trivial, the -jet bundle may be identified with the affine subbundle of
Thus, is isomorphic to the vector bundle .
An Ehresmann connection on the fibration is completely determined by a vector field on satisfying
In fact, the horizontal subbundle associated with the connection is of rank and generated by the vector field .
The extended (resp. restricted) multimomentum bundle may be identified with the cotangent bundle (resp. the vector bundle ) and, under this identification, the multisymplectic structure on is just the canonical symplectic structure on .
We have a principal -action on given by
for and . The principal bundle projection is just the canonical projection
Note that admits a Poisson structure of corank which is induced by the canonical symplectic structure on . In fact, the -action on the extended multimomentum bundle preserves the symplectic form and the canonical projection is a Poisson map.
In this case, a Hamiltonian section may be identified with a global time-dependent Hamiltonian function . In addition, the couple
is a cosymplectic structure on and, thus, we can consider the Reeb vector field of .
Remark 5.1.
We recall that a cosymplectic structure on a manifold of odd dimension is a couple , where is a closed -form, is a closed -form and is a volume form on . The Reeb vector field associated with the structure is the vector field on which is completely characterized by the conditions
Note that the -form on is given by
where is the canonical symplectic structure on . So, the Reeb vector field of the cosymplectic structure on is
with the Hamiltonian vector field on associated with the function
Thus, if are canonical coordinates on then
and the integral curves of are just of the solution of the Hamilton equations for , that is,
Therefore, is the evolution vector field associated with the time-dependent Hamiltonian function (for more details see, for instance, [15, 20]).
In addition, the vector bundle may be identified with the time-dependent cotangent bundle to ,
where is the canonical projection.
Under the previous identification, the principal -action on is given by
for and .
Moreover, the extended phase bundle is
In other words, may be identified with the time-dependent vector bundle associated with the vector bundle , that is,
and the principal -action on is given by
for and .
Thus, the phase bundle may be identified with the time-dependent cotangent bundle to
Then, the differential of is just the vertical differential (with respect to the projection ) of the Hamiltonian function , that is,
On the other hand, the vector bundle , which was introduced in Proposition 3.10, is isomorphic to the cotangent bundle to .
Moreover, as we know, the -jet bundle to the projection
may be identified with the affine subbundle of given by
(5.1) |
or, equivalently, with the time-dependent tangent bundle to
Under all the previous identifications, the affine bundle isomorphism is given by
with the canonical symplectic structure of . Thus, in this case, the vector subbundle is trivial and this implies that there exists a unique Hamiltonian connection for the hamiltonian section . In fact, the horizontal subbundle of such a connection is generated by the evolution vector field .
In addition, if is the inverse morphism of it is clear that, under the identification of with the affine subbundle of given by (5.1), the image of the section of is just the evolution vector field .
On the other hand, the space of currents is the set of smooth real functions on (the space of observables in Classical Mechanics)
and it is clear that the space of sections of the projection
may be also identified with . Then, the linear-affine bracket
is given by
for .
In addition, the bracket on the space of observables is just the standard Poisson bracket induced by the canonical symplectic structure of . In other words,
for .
Finally, the affine representation of the Lie algebra on the affine space is given by
Concretely, we have
5.2. A particular case: the configuration bundle is trivial and the base space is orientable
In this section we will assume that , is the canonical projection on the first factor. In this case, the affine bundle
can be identified with the vector bundle
Let us further assume that is orientable, with , and fix a volume form on . We denote by the -vector on which is characterized by the condition
Using the volume form vol on , we have
and the vector bundle may be trivialized as the trivial line vector bundle . Using vol again, the reduced multimomentun bundle is isomorphic to the vector bundle.
We can also identify it with
As in the general case, we will denote by the vector bundle projection. We will see that this space admits a multisymplectic structure.
Proposition 5.2.
Let be the -form on given by
for and , where is the projection on the second factor. Then, is a multisymplectic structure on .
Proof.
A direct computation proves that the local expression of is
Thus, the local expression of is
(5.2) |
Therefore,
which implies that is a multisymplectic structure on . ∎
On the other hand, the extended multimomentum bundle may be identified with the Withney sum of the vector bundles and , that is,
So, using the volume form vol, we deduce that
Under this identification, the canonical multisymplectic structure on is
where is the canonical coordinate on . Here, we also denote by and vol the pullbacks to of and vol, respectively.
Moreover, a Hamiltonian section is just a global Hamiltonian function on and the -form on is
(5.3) |
On the other hand, under the identification between and and using (B.1) (see Appendix B), it follows that
(5.4) |
where is the standard coordinate on .
So, the extended phase bundle is isomorphic to the affine bundle over
where is the vertical bundle of the fibration .
Now, using the previous identifications, we have that the fibred action of on is just the standard action of on . Therefore, since
for , we deduce that the phase bundle is isomorphic to the vector bundle . An isomorphism
between these spaces is given by
for , with an arbitrary real number.
Thus, the image of the vertical differential of under the previous isomorphism is just the equivalence class induced by the vertical differential of the extended Hamiltonian density . This implies that, under the identification between and , the differential of (as a section of the affine bundle ) is just (as a section of the vector bundle ).
Next, following Section 3.2 (see (3.14) and (3.15)), we will define the vector bundle monomorphism
as follows
for and . We have
Now, we consider the restriction to of the dual morphism of , that is,
A direct computation proves that
for and . Thus, if are local coordinates on such that
then, using (5.2), we deduce that
This, from (3.20) and under the identification between and , implies that , with the affine bundle epimorphism given by (3.15), (3.18) and (3.19).
So, if is the affine bundle isomorphism induced by then, under the identification between and , we conclude that
We will denote by
the inverse morphism of .
Now, from (5.3) and Definition 3.5, we deduce that a connection on the fibration is Hamiltonian if and only if
where is the vertical differential of with respect to the projection .
Moreover, using that , it follows that the vector subbundle of introduced in Proposition 3.10 is
In addition, as we know (see first step in Section 3.4), we have that
On the other hand, under the identification between and , the projection is just the canonical projection on the second factor. Thus, the affine space is isomorphic to the vector space and the linear-affine bracket
given by (3.39) may be considered as a bracket
defined by
5.3. Continuum Mechanics
In this section we develop the formulation of Continuum Mechanics as a Canonical Hamiltonian Field Theory. This covers the case of fluid mechanics and nonlinear elasticity. We shall assume that the reference configuration of the continuum is described by a manifold of dimension , possibly with boundary, and we suppose that the continuum evolves in a dimensional manifold without boundary, the ambient manifold, typically . The elements denote the labels of the material points of the continuum, whereas the elements denote the current positions of these material points. The evolution of the continuum is described by a map , where is the interval of time. Hence, describes the position of the material point at time . We shall assume that for each fixed, the map is a smooth embedding. Boundary conditions will be described in §5.3.4
5.3.1. Lagrangian and Hamiltonian formulations in continuum mechanics.
Continuum mechanics is usually written either as a Lagrangian field theory or as an infinite dimensional classical Lagrangian or Hamiltonian system. While the infinite dimensional description is more classical, the field-theoretic description is especially useful for the derivation of multisymplectic integrators for fluid and elasticity ([16, 17, 18, 42, 45]).
In the field theoretic Lagrangian description, the map is interpreted as a section of the trivial fiber bundle , , by writing . The equations of motion are given by the Euler-Lagrange equations for a given Lagrangian density , . Since the bundle is trivial, we have . We denote by the local coordinates. Writing locally the Lagrangian density as , the Euler-Lagrange equations are given by
In the infinite dimensional classical Lagrangian description, the map is interpreted as a curve in the infinite dimensional manifold of smooth embeddings of into . The equations are given by the (classical) Euler-Lagrange equations for the Lagrangian function defined from as
where we assumed that the Lagrangian density does not depend explicitly on the time , and denotes the tangent map to , i.e. locally . When is hyperregular, to this classical Lagrangian description is formally associated a classical Hamiltonian description with respect to the Hamiltonian defined on the (regular) cotangent bundle of . The Hamiltonian is defined by
where is such that . In this case, the associated equations can formally be written with respect to the canonical Poisson bracket on .
The Hamiltonian formulation that we present below is different from this one, since it is associated to the field theoretic Lagrangian formulation. Roughly speaking, while the canonical Hamiltonian formulation recalled above is based on a Legendre transform with respect to the time direction only, the canonical Hamiltonian field theoretic description that we will describe below is based on a Legendre transform with respect to all the variables in the base manifold .
We warn the reader that the coordinates that were used in the previous sections for the base manifold are here given by for . The coordinates used earlier on the fiber of are here given by and represent the material velocity and the deformation gradient of the continuum.
5.3.2. Lagrangian density and Legendre transform
The Lagrangian density of continuum mechanics is defined with the help of given tensor fields on and . In order to treat both fluid dynamics and elasticity from a unified perspective, we shall consider here a Riemannian metric on , two volume forms and on , and a Riemannian metric on . Additional tensor fields can be introduced to describe electromagnetic effects or microstructures. The volume forms and are the mass density and the entropy density in the reference configuration and are locally written as and . The potential energy density is a bundle map
covering the projection . In local coordinates, it reads
This is a general form of potential energy density for continua, including fluid and elasticity, which may describe both internal and stored energies.
The associated Lagrangian density is given by the kinetic minus the potential energy, and reads
(5.5) | ||||
in local coordinates. Note that the Lagrangian is defined with the help of the given tensor fields , , , and . We chose to work with the cometric associated to , in order to directly get the Finger deformation (or left Cauchy-Green) tensor , rather than its inverse, later.
The restricted multimomentum bundle for continuum mechanics is given by
with coordinates . The restricted Legendre transform of the Lagrangian density is
(5.6) |
with and given by
(5.7) |
with the momentum density (in the Lagrangian description) and is the Piola-Kirchoff stress tensor density. Note that the coordinates on the fiber of correspond to the coordinates denoted earlier.
The Eulerian versions of these tensor densities are the Eulerian momentum density and the Cauchy stress tensor density given by the Piola transformation
(5.8) |
From the second relation, we have
(5.9) |
Note that the first relation in (5.7) is always invertible, but the invertibility of the second relation depends in the potential energy density . As we shall illustrate below, relation (5.9) is extremely useful to check if the restricted Legendre transform (5.6) is an isomorphism, in which case we say that the Lagrangian density is hyperregular.
5.3.3. The Hamiltonian density and the linear-affine bracket for Continuum Mechanics
By assuming that is hyperregular, we get the Hamiltonian
(5.10) |
where is expressed in terms of the variables in by inverting the second relation in (5.7).
A section of the restricted momentum bundle is locally given by
and, in the hyperregular case, the Euler-Lagrange equation are equivalent to the Hamilton-deDonder-Weyl equations given by
(5.11) |
These equations admit the canonical linear-affine bracket formulation, that is, a section is a solution of the previous equations if
(5.12) |
where the currents for Continuum Mechanics are of the form
with . The canonical linear-affine bracket is given by
This formulation assumes that the Legendre transform is invertible. Except in some simple situations, this invertibility is a priori difficult to check. We shall show below how to facilitate the approach by using two symmetries of the potential energy density . The first one, the material covariance, is related to the isotropy of the continuum, while the second, the material frame indifference, is a general covariance assumption of continuum theories, see [48] and [30].
We assume that the potential energy density is of the form
where is the potential energy density in the Eulerian description. This assumption is compatible with the assumption of material covariance. Here is a bundle map
covering the identity on . From this expression, we compute the momenta from the second equality in (5.7) as
(5.13) |
where we introduced the notations
These are the local expressions of the mass density and entropy density in Eulerian description, and of the Finger deformation (or left Cauchy-Green) tensor.
The associated Cauchy stress tensor density , see the second equation in (5.8), is
(5.14) |
If in addition satisfies the material frame indifference, then
(5.15) |
for all diffeomorphisms of and we have the Doyle-Ericksen formula
(5.16) |
By inserting these relations into (5.10), we get the following result which is a step towards a more explicit expression of the Hamiltonian density, because in practice , rather than , is given.
Proposition 5.3.
Assume that the Lagrangian is hyperregular and that satisfies the two invariance mentioned above, and consider the associated Eulerian potential energy density . Then, the Hamiltonian density of continuum mechanics is given by
(5.17) |
5.3.4. Boundary conditions
We briefly describe two mains boundary conditions used in Continuum Mechanics following §3.6. These conditions only arise at the spatial part of the boundary of the base manifold , hence the bundle is over only and the boundary condition reads .
For a continuum moving in a fixed domain diffeomorphic to , we have the boundary condition on the motion and, in addition, the boundary condition on the Piola-Kirchhoff stress tensor given by , for all , with the normal vector field to with respect to . This corresponds to zero tangential traction on the boundary, a condition that vanishes for fluids. In this case, the subbundle is given by
In particular, we have and is a vector bundle.
For a free boundary continuum we take
which corresponds to zero traction on the boundary. This reduces to zero pressure at the boundary for fluids. We have and again is a vector bundle.
5.3.5. Fluid dynamics
In this case the energy density only depends on the mass density and entropy density and , so the Cauchy stress density is given by
see (5.14), where is the pressure of the fluid. In this case (5.9) yields
This relation is of the form
(5.18) |
for some function . If the function is invertible on , with inverse , then relation (5.18) is invertible, with inverse
(5.19) |
In this case the Lagrangian density is hyperregular. Note that the function , and hence the hyperregularity, depends on the state function of the fluid, i.e., the relation .
Hyperregularity is satisfied for a large class of state equations, including the important case of a perfect gas for which , where is the adiabatic index and is the volume form associated to , i.e. . In this case, we compute the pressure as .
Note that, as it should, satisfies (5.14). Computing the derivative of with respect to the Riemannian metric, we get
so one directly checks that the Doyle-Ericksen formula (5.16) is verified.
For fluids, the Hamiltonian density is
where is found from (5.19). In particular, for the perfect gas, we have .
The fluid equations can thus be written in the canonical linear-affine bracket form (5.12).
5.3.6. Nonlinear elasticity
In general, the Hamiltonian density in nonlinear elasticity takes a complicate expression due to the dependence of on the Finger deformation tensor . For example, for the compressible neo-Hookean material (see [51], [3]), with , the energy density is
where is the bulk modulus, is the Lamé constant, and is the volume form associated to the Riemannian metric , obtained by lowering the indices of . One observes that (5.14) is satisfied. The Doyle-Ericksen formula yields the expression of the stress tensor density
We thus get which can then be inserted in (5.17) to yield the Hamiltonian density.
We shall illustrate the derivation of the Hamiltonian density by considering the simplified situation . In this case , so we get the momenta
Using this and , we get the Hamiltonian density
The nonlinear elasticity equations can thus be written in the canonical linear-affine bracket form (5.12).
5.4. Yang-Mills theory
Yang-Mills theory may be considered as a singular Lagrangian field theory of first order associated with a principal -bundle over an oriented Riemannian (or a Lorentzian manifold) space (possibly with boundary) of dimension and where is a compact Lie group of dimension (we will follow [39]).
We will denote by the metric on . For simplicity, we will assume that the principal bundle is trivial, is a Riemannian metric and .
Under the previous conditions, the configuration bundle of the theory is the vector bundle
where is the Lie algebra of .
Then, we will proceed as follows. We will introduce a Lagrangian density on the -jet bundle of the fibration . This Lagrangian density is singular. In fact, the image of the corresponding Legendre transformation is a proper submanifold of the restricted multimomentum bundle . Using the restricted and the extended Legendre transformation, we will construct a constrained Hamiltonian section of the fibration . Now, if we consider an (arbitrary) hamiltonian section , whose restriction to coincides with , we will obtain a Hamiltonian field theory in such a way that the solutions of the Hamilton-deDonder-Weyl equations for which are contained in are just the solutions of the corresponding Yang-Mills theory.
5.4.1. The Lagrangian formalism
Note that the sections of the vector bundle are the principal connections on the trivial principal bundle . As we know, the -jet bundle is an affine bundle over . The key point is that there is a canonical epimorphism off affine bundles (over the vector bundle projection ), , which is characterized by the condition
for all principal connections . In other words, the image by of the -jet bundle of a principal connection is just the curvature of the connection.
If are local coordinates on and is a basis of , we have the corresponding local coordinates on and on . Moreover,
with
(5.20) |
Here, are the structure constants of the Lie algebra with respect to the basis .
Next, we will introduce the Lagrangian density
First of all, since the manifold is oriented, the vector bundle is the trivial line bundle . So, the Lagrangian density is, in fact, a real -function .
In addition, we will fix an -invariant scalar product on (which is possible, since is compact). Then, the scalar product on and the Riemannian metric on induce a bundle metric on the vector bundle
So, we can consider the real function given by
where the norm is taken with respect to the bundle metric on the vector bundle
The local expression of is
with
and the matrix of the coefficients of , the inverse matrix and
Thus, the Euler-Lagrange equations for
are, in this case, the well-known Yang-Mills equations
(5.21) |
5.4.2. The Legendre transformations and the constrained Hamiltonian formalism
First of all, we will consider the restricted Legendre transformation associated with
Note that, since is oriented, the restricted multimomentum bundle may be identified with the dual bundle of . So,
The transformation is given by
for , and . The local expression of is
This implies that the image of is the vector subbundle (over ) of
Thus, the map is a submersion with connected fibers and is almost regular.
On the other hand, we can consider the extended Legendre transformation associated with defined by
The local expression of is
where
Note that if is the canonical projection then the image of is a submanifold of which is diffeomorphic to , via the restriction of to . The following diagram illustrates the situation
The maps and are surjective submersions and, thus, we have a constrained Hamiltonian field theory. As a consequence (see, for instance, [21]), one may introduce a constrained Hamiltonian section
in such a way that
In fact, . In addition, if and are the standard local coordinates on and , respectively, we can take local coordinates
on and , respectively, with
(5.22) |
Then,
where
(5.23) |
Note that we are assuming
In addition,
It is clear that
Moreover, on
(5.24) |
So, using (5.20), (5.21) and (5.24), we obtain that
(5.25) |
(5.24) and (5.25) are just the Yang-Mills equations for the Yang-Mills theory in the Hamiltonian side. Thus, Yang-Mills theory may be considered as a constrained (singular) Hamiltonian field theory.
On the other hand, if is a Hamiltonian section which extends (that is, ), then we may consider the corresponding Hamiltonian field theory associated with . Furthermore, using the classical results on singular Lagrangian field theories (see [21]), if is a solution of the Hamilton-deDonder-Weyl equations for which is contained in then is just a solution of the Yang-Mills equations.
The following diagram illustrates the situation
5.4.3. The Lie algebra of currents and the linear-affine bracket for the extended Hamiltonian field theory
After the previous subsections, we could apply all the machinery in this paper for the extended Hamiltonian field theory and, as a consequence, we could deduce results on the Yang-Mills theory. This will be the subject of a future research. Anyway, we will remark a couple of general facts on the Lie algebra of currents, the linear-affine bracket (in Section 3.5) and the Yang-Mills equations as constrained Hamilton-deDonder-Weyl equations:
-
•
First of all, following the proof of Theorem 4.2, we have that the space of currents, as a -module, may be identified with the product . But, since the configuration bundle is a vector bundle, we have that is generated by vertical lifts of sections of the projection (see Appendix B). In fact, if
is a -form on and then the local expression of the vertical lift of the section is
So, if we chose a local basis of -forms and -forms on
respectively, we have a local basis
of the space . Moreover, following the proof of Theorem 4.2, we also deduce that the Lie brackets in between the previous sections are all zero. Note that if are sections of the vector bundle then
for .
-
•
Consider the linear-affine bracket
for the Hamiltonian field theory which extends Yang-Mills theory. We have that if is a -form on , , is a -form on and
is a section of then
In particular, if is an extension of the Yang-Mills Hamiltonian section then, is a solution of the Yang-Mills equations if and only if
and
for a -form on , and a -form on .
6. Conclusions and future work
In this paper, we have developed a completely canonical geometric formulation of Hamiltonian Classical Field Theories of first order which is analogous to the canonical Poisson formulation of time-independent Hamiltonian Mechanics. This formulation is valid for any configuration bundle and is independent of any external structures such as connections or volume forms. We have defined a space of currents and endowed it with a Lie algebra structure, and we have shown that the bracket induces an affine representation of the Lie algebra of currents on the affine space of Hamiltonian sections. An important difference with the case of time-independent Hamiltonian Mechanics is the linear-affine character of our bracket, which is consistent with the fact that the set of currents and the set of Hamiltonian sections are linear and affine spaces, respectively. We have applied our results to several examples and we have proved their effectiveness.
The results of this paper open some interesting future directions of research:
-
•
Develop appropriate processes of reduction by symmetry for Hamiltonian Classical Field Theories of first order by exploiting the canonical linear-affine bracket formulation proposed in this paper.
- •
-
•
Discuss a canonical affine formulation of Lagrangian Classical Field Theories of first oder and obtain the equivalence with the Hamiltonian formulation for the case when Lagrangian density is almost regular (this is, for instance, the case of Yang-Mills theories discussed in this paper).
Appendix A The -jet bundle associated with a fibration
Let be a fibration, that is, a surjective submersion from to . Assume that and .
The -jet bundle associated with is the affine bundle over whose fiber at the point is
is modelled over the vector bundle
where is the vertical bundle of .
In fact, if and then one may define, in a natural way, the element of as the linear map
If are local coordinates on which are adapted to the fibration then one may consider the corresponding local coordinates on . Indeed, if , has local coordinates and
then has local coordinates .
Sections of the affine bundle may be identified with Ehresmann connections on the fibration .
We recall that an Ehresmann connection on is a vector subbundle over of satisfying
Note that if is an Ehresmann connection on , then one may define the horizontal lift associated with as a vector bundle morphism
In fact, if then is the unique vector in whose projection over is , that is,
It is clear that the horizontal lift induces, in a natural way, a section of the -jet bundle . The previous correspondence is one-to-one. Indeed, if
then
A (local) section is said to be horizontal with respect to the Ehresmann connection if
So, if
then is a horizontal section of if and only if it satisfies the following system of partial differential equations
The Ehresmann connection is said to be integrable if the distribution is completely integrable. In such a case, for every point there exists a unique horizontal (local) section of such that and
Appendix B The vertical lift of a section of a vector bundle
In this appendix, we review the definition of the vertical lift of a section of a vector bundle.
Let be a vector bundle and a smooth section of . Then, one may define the vertical lift of as a vector field on given by
It is clear that is vertical with respect to the vector bundle projection .
Next, we will obtain a local expression of . For this purpose, we consider local coordinates in an open subset of and a local basis of in . Then, we will denote by the corresponding local coordinates on . Moreover, if
then
A particular case of the previous situation is the following one.
Let be a fibration, with , and the multimomentum bundle associated with the fibration . It is a vector bundle over (see Section 2.1).
Now, if is a -form on then is a -form on . Moreover, also is a section of the vector bundle . Thus, one may consider the vertical lift of as a -vertical vector field on . In fact, if are canonical coordinates on as in Section 2.1, then
(B.1) |
Appendix C Diagram
The diagram below illustrates some of the objects used in the paper and their relations. The arrows with label (resp. ) indicate that we associate the vector bundle dual to a given affine (resp. vector) bundle. The four dashed arrows pointing down indicate the associated model vector bundle to a given affine bundle. The diagram illustrates the isomorphism of affine bundles and three vector bundle isomorphisms naturally associated to it:
-
(1)
the vector bundle isomorphism obtained by taking (minus) the affine dual to ;
-
(2)
the vector bundle isomorphism naturally induced by on the associated model vector bundles;
-
(3)
the vector bundle isomorphism obtained by taking (minus) the dual to .
Given a Hamiltonian section , an observable and a section (the space of sections of the vector bundle associated to ), the diagram also illustrates the sections , , , and the corresponding objects obtained by applying sharp operators, namely , , , . These objects enter into the definition of the three structures , , as
References
-
[1]
Abraham, R.; Marsden, J.E.
Foundations of mechanics. Second edition, revised and enlarged. With the assistance of Tudor Ratiu and Richard Cushman. Benjamin/Cummings Publishing Co., Inc., Advanced Book Program, Reading, Mass., 1978. -
[2]
Asorey, M.; Ibort, A.; Spivak, A.
Admissible boundary conditions for Hamiltonian field theories. Int. J. Geom. Methods Mod. Phys. 14 (2017), no. 8, 1740006, 17 pp. -
[3]
Başar, Y.; Weichert, D.
Nonlinear Continuum Mechanics of Solids: Fundamental Mathematical and Physical Concepts. Springer-Verlag, 2000. -
[4]
Barbero G., J. Fernando; Margalef-Bentabol, Juan; Varo, Valle; Villaseñor, Eduardo J. S.
Palatini gravity with nonmetricity, torsion, and boundaries in metric and connection variables. Phys. Rev. D 104 (2021), no. 4, Paper No. 044046, 6 pp. -
[5]
Barbero G., J. Fernando; Margalef-Bentabol, Juan; Varo, Valle; Villaseñor, Eduardo J. S.
Covariant phase space for gravity with boundaries: metric versus tetrad formulations. Phys. Rev. D 104 (2021), no. 4, Paper No. 044048, 16 pp. -
[6]
Barbero G., J. Fernando; Margalef-Bentabol, Juan; Varo, Valle; Villaseñor, Eduardo J. S.
On-shell equivalence of general relativity and Holst theories with nonmetricity, torsion, and boundaries. Phys. Rev. D 105 (2022), no. 6, Paper No. 064066, 15 pp. -
[7]
Binz, E.; Sniatycki, J.; Fischer, H.
Geometry of classical fields. North-Holland Mathematics Studies, 154. Notas de Matemática (Mathematical Notes), 123. North-Holland Publishing Co., Amsterdam, 1988. -
[8]
Bursztyn, H.; Cabrera, A.; Iglesias, D.
Multisymplectic geometry and Lie groupoids. Geometry, mechanics, and dynamics, 57–73, Fields Inst. Commun., 73, Springer, New York, 2015. -
[9]
Bursztyn, H; Martínez Alba N.; Rubio, R.
On higher Dirac structures. International Mathematics Research Notices, Volume 2019, Issue 5, March 2019, Pages 1503–1542. -
[10]
Campos, Cédric M.; Guzmán, Elisa; Marrero, Juan Carlos
Classical field theories of first order and Lagrangian submanifolds of premultisymplectic manifolds. J. Geom. Mech. 4 (2012), no. 1, 1–26. -
[11]
Cariñena, J.F.; Crampin, M.; Ibort, A.
On the multisymplectic formalism for first order field theories. Differential Geom. Appl. 1 (1991), no. 4, 345–374. -
[12]
Castrillón López, M.; Marsden, J.E.
Some remarks on Lagrangian and Poisson reduction for field theories. J. Geom. Phys. 48 (2003), no. 1, 52–83. -
[13]
Ciaglia, Florio M.; Di Cosmo, Fabio; Ibort, Alberto; Marmo, Giuseppe; Schiavone Luca; Zampini, Alessandro
The Geometry of the solution space of first order Hamiltonian field theories I: from particle dynamics to free Electrodynamics. Preprint, arXiv:2208.14136, 2022. -
[14]
Ciaglia, Florio M.; Di Cosmo, Fabio; Ibort, Alberto; Marmo, Giuseppe; Schiavone Luca; Zampini, Alessandro
The Geometry of the solution space of first order Hamiltonian field theories II: non-Abelian gauge theories Preprint, arXiv:2208.14155, 2022. -
[15]
Chinea, D.; de León, M.; Marrero, J.C.
The constraint algorithm for time-dependent Lagrangians. J. Math. Phys. 35 (1994), no. 7, 3410–3447. -
[16]
Demoures, F; Gay-Balmaz F.
Multisymplectic variational integrators for barotropic and incompressible fluid models with constraints, Preprint arXiv:2102.10907 -
[17]
Demoures, F; Gay-Balmaz F.
Unified discrete multisymplectic Lagrangian formulation for hyperelastic solids and barotropic fluids, J. Nonlin. Sc., accepted. -
[18]
Demoures, F; Gay-Balmaz F.; Ratiu, T.S
Multisymplectic variational integrator and space/time symplecticity, Anal. Appl., 14 (2014), no. 3, 341–391. -
[19]
de León, M.; Marín-Solano, J.; Marrero, J.C.
Ehresmann connections in classical field theories. Differential geometry and its applications (Granada, 1994), 73–89, An. Fís. Monogr., 2, CIEMAT, Madrid, 1995. -
[20]
de León, M.; Marín-Solano, J.; Marrero, J.C.
The constraint algorithm in the jet formalism. Differential Geom. Appl. 6 (1996), no. 3, 275–300. -
[21]
de León, M.; Marín-Solano, J.; Marrero, J.C.
A geometrical approach to classical field theories: a constraint algorithm for singular theories. New developments in differential geometry (Debrecen, 1994), 291–312, Math. Appl., 350, Kluwer Acad. Publ., Dordrecht, 1996. -
[22]
de León, M.; Martín de Diego, D.; Santamaría-Merino, A.
Symmetries in classical field theory. Int. J. Geom. Methods Mod. Phys. 1 (2004), no. 5, 651–710. -
[23]
de León, M.; Rodrigues, P.R.
Methods of differential geometry in analytical mechanics. North-Holland Mathematics Studies, 158. North-Holland Publishing Co., Amsterdam, 1989. -
[24]
Echeverría-Enríquez, A.; de León, M.; Muñoz-Lecanda, M.C.; Román-Roy, N.
Extended Hamiltonian systems in multisymplectic field theories. J. Math. Phys. 48 (2007), no. 11, 112901, 30 pp. -
[25]
Echeverría-Enríquez, A.; Muñoz-Lecanda, M.C.; Román-Roy, N.
Multivector field formulation of Hamiltonian field theories: equations and symmetries. J. Phys. A 32 (1999), no. 48, 8461–8484. -
[26]
Echeverría-Enríquez, A.; Muñoz-Lecanda, M.C.; Román-Roy, N.
Geometry of multisymplectic Hamiltonian first-order field theories. J. Math. Phys. 41 (2000), no. 11, 7402–7444. -
[27]
Forger, M.; Romero, Sandro V.
Covariant Poisson brackets in geometric field theory. Comm. Math. Phys. 256 (2005), no. 2, 375–410. -
[28]
Forger, M.; Salles, Mário O.
On covariant Poisson brackets in classical field theory. J. Math. Phys. 56 (2015), no. 10, 102901, 26 pp. -
[29]
García, P. L.
The Poincaré-Cartan invariant in the calculus of variations. Symposia Mathematica, Vol. XIV (Convegno di Geometria Simplettica e Fisica Matematica, INDAM, Rome, 1973), pp. 219–246. Academic Press, London, 1974. -
[30]
Gay-Balmaz, F.; Marsden, J.E.; Ratiu T.S.
Reduced variational formulations in free boundary continuum mechanics. J. Nonlinear Sci. 22 (2012), no. 2, 553–597. -
[31]
Goldschmidt, H.; Sternberg, S.
The Hamilton-Cartan formalism in the calculus of variations. Ann. Inst. Fourier, 23 (1973), 203–267. -
[32]
Grabowska, K.
A Tulczyjew triple for classical fields. J. Phys. A 45 (2012), no. 14, 145207, 35 pp. -
[33]
Grabowska, K.; Grabowski, J.
Tulczyjew triples: from statics to field theory. J. Geom. Mech. 5 (2013), no. 4, 445–472. -
[34]
Grabowska, K.; Grabowski, J.; Urbański, P.
Lie brackets on affine bundles. Ann. Global Anal. Geom. 24 (2003), no. 2, 101–130. -
[35]
Grabowska, K.; Grabowski, J.; Urbański, P.
AV-differential geometry: Poisson and Jacobi structures. J. Geom. Phys. 52 (2004), no. 4, 398–446. -
[36]
Grabowska, Katarzyna; Grabowski, Janusz; Urbański, P.
Frame-independent mechanics: geometry on affine bundles. Travaux mathématiques. Fasc. XVI, 107–120, Trav. Math., 16, Univ. Luxemb., Luxembourg, 2005. -
[37]
Grabowska, Katarzyna; Grabowski, Janusz; Urbański, P.
AV-differential geometry: Euler-Lagrange equations. J. Geom. Phys. 57 (2007), no. 10, 1984–1998. -
[38]
Hansoul, S.; Lecomte, P.
Affine representations of Lie algebras and geometric interpretation in the case of smooth manifolds. Int. Math. Res. Not. 2005, no. 16, 981–1003. -
[39]
Ibort, A.; Spivak, A.
Covariant Hamiltonian field theories on manifolds with boundary: Yang-Mills theories. J. Geom. Mech. 9 (2017), no. 1, 47–82. -
[40]
Ibort, A.: Spivak, A.
On a covariant Hamiltonian description of Palatini’s gravity on manifolds with boundary. Preprint arXiv:1605.03492 -
[41]
Iglesias, D.; Marrero, J. C.; Padrón, E.; Sosa, D.
Lagrangian submanifolds and dynamics on Lie affgebroids. Rep. Math. Phys. 57 (2006), no. 3, 385–436. -
[42]
Lew A.; Marsden, J. E.; Ortiz, M.; West, M.
Asynchronous variational integrators. Arch. Rational Mech. Anal., 167 (2003), no.2, 85–146. -
[43]
Libermann, P.; Marle, C.-M.
Symplectic geometry and analytical mechanics. Translated from the French by Bertram Eugene Schwarzbach. Mathematics and its Applications, 35. D. Reidel Publishing Co., Dordrecht, 1987. -
[44]
Margalef-Bentabol, Juan; Villaseñor, Eduardo J. S.
Geometric formulation of the covariant phase space methods with boundaries. Phys. Rev. D 103 (2021), no. 2, Paper No. 025011, 37 pp. -
[45]
Marsden, J.E.; Patrick, G.W.; Shkoller, S.
Multisymplectic geometry, variational integrators and nonlinear PDEs, Comm. Math. Phys., 199, 1998, 351–395. -
[46]
Marsden, J.E.; Shkoller, S.
Multisymplectic geometry, covariant Hamiltonians, and water waves. Math. Proc. Cambridge Philos. Soc. 125 (1999), no. 3, 553–575. -
[47]
Martínez, E.; Mestdag, T.; Sarlet, W.
Lie algebroid structures and Lagrangian systems on affine bundles. J. Geom. Phys. 44 (2002), no. 1, 70–95. -
[48]
Marsden, J. E.; Hughes, T.J.R.
Mathematical Foundations of Elasticity. (Prentice Hall, New York, 1983) (reprinted by Dover, New York, 1994). -
[49]
Marsden, J.E.; Ratiu, T.S.
Introduction to Mechanics and Symmetry. A basic exposition of classical mechanical systems. Texts in Applied Mathematics, 17. Springer-Verlag, New York, 1994. -
[50]
Peierls, R.E.
The commutation laws of relativistic field theory. Proc. Roy. Soc. London. Ser. A. 214, (1952). 143–157. -
[51]
Simo J.C.; Taylor, R.L., and Pister, K.S.
Variational and projection methods for the volume constraint in finite deformation elasto-plasticity, Comp. Meth. Appl. Mech. Engng. 51, (1985), 177–208. -
[52]
Volterra, V.
Sulle equazioni differenziali che provengono da questiono di calcolo delle variazioni, Rend. Cont. Acad. Lincei, ser. IV, vol. VI, 1890, 42–54. -
[53]
Volterra, V.
Sopra una estensione della teoria Jacobi-Hamilton del calcolo delle varizioni, Rend. Cont. Acad. Lincei, ser. IV, vol. VI, 1890, 127–138