1. Introduction
The Hamilton–Jacobi theory is a topic of interest in mathematical physics since it is a way to integrate systems of first-order ordinary differential equations (Hamilton equations in the standard case). The classical method in Hamiltonian mechanics consists of obtaining a suitable canonical transformation, which leads the system to equilibrium [
1,
2,
3,
4], and is given by its generating function. This function is the solution to the so-called
Hamilton–Jacobi equation, which is a partial differential equation. The
“classical” Hamilton–Jacobi problem consists of finding this canonical transformation. Because of its interest, the method was generalized in other kinds of physical systems, such as, for instance, singular Lagrangian systems [
5] or higher-order dynamics [
6], and different types of solutions have been proposed and studied [
7,
8].
Nevertheless, in recent times, a lot of research has been done to understand the Hamilton–Jacobi equation from a more general geometric approach, and some geometric descriptions to the theory were done in Reference [
9,
10,
11,
12,
13]. From a geometric way, the above mentioned canonical transformation is associated with a foliation in the the phase space of the system which is represented by the cotangent bundle
of a manifold (the configuration manifold
Q). This foliation has some characteristic geometric properties: it is invariant by the dynamics, transversal to the fibers of the cotangent bundle, and Lagrangian with respect to the canonical symplectic structure of
(although this last property could be ignored in some particular situations). The restriction of the dynamical vector field in
to each leaf
of this foliation projects onto another vector field
on
Q, and the integral curves of these vector fields are one-to-one related. Hence, the complete set of dynamical trajectories are recovered from the integral curves of the complete family
of all these vector fields in the base. These geometric considerations can be done in an analogous way in the Lagrangian formalism; hence, this geometrical picture of the Hamilton–Jacobi theory can be also stated for this formalism. The
geometric Hamilton–Jacobi problem consists of finding this foliation and these vector fields
.
Following these ideas, the Lagrangian and Hamiltonian versions of the Hamilton–Jacobi theory, for autonomous and non-autonomous mechanical systems, was formulated in another geometrical way in Reference [
14]. The foundations of this geometric generalization are similar to those given in Reference [
9,
15]. Later, this framework was used to develop the Hamilton–Jacobi theory for many other kinds of systems in physics. For instance, other applications of the theory are to the case of singular Lagrangian and Hamiltonian systems [
16,
17,
18,
19], higher-order dynamical systems [
20,
21], holonomic and non-holonomic mechanics [
19,
22,
23,
24,
25,
26,
27], and control theory [
28,
29]. The theory is also been extended for dynamical systems described using other geometric structures, such as Poisson manifolds [
18,
30], Lie algebroids [
31,
32], contact manifolds (which model dissipative systems) [
33,
34], and other geometric applications and generalizations: [
35,
36,
37]. Furthermore, in Reference [
38,
39], the geometric discretization of the Hamilton–Jacobi equation was analyzed. Finally, the Hamilton–Jacobi theory is developed for the usual covariant formulations of first-order classical field theories, the
k-symplectic and
k-cosymplectic [
40,
41] and the multisymplectic ones [
42,
43], for higher-order field theories [
44,
45], for the formulation in the
Cauchy data space [
46], and for partial differential equations in general [
47,
48].
This review paper is devoted, first of all, to present, in
Section 2, the foundations of this modern geometric formulation of the Hamilton–Jacobi theory, starting from the most general problem and explaining how to derive the standard Hamilton–Jacobi equation for the Hamiltonian and the Lagrangian formalisms of autonomous mechanics. After this, the notion of complete solution allows us to establish the relation with the “classical” Hamilton–Jacobi theory based on canonical transformations, which is summarized in
Section 3, where this relation is also analyzed (this topic had been already discussed in Reference [
48]). We also present briefly a more general geometric framework for the Hamilton–Jacobi theory which was stated in Reference [
36], from which we can derive the majority of the applications of the theory to other kinds of physical systems, including the case of autonomous dynamical systems. This is done in
Section 4. Finally, among all the extensions of the theory, we have selected two of them for reviewing: the case of higher-order autonomous dynamical systems (Lagrangian and Hamiltonian formalisms), which is a direct application of the above general framework, and the generalization to the Lagrangian and Hamiltonian multisymplectic formalisms of first-order classical field theories, which can also be interpreted as a special case of the general framework. Both of them are treated in
Section 5.
Throughout the work it is considered that the manifolds are real, smooth, and second countable. In the same way, all the maps are assumed to be smooth. The summation convention for repeated cross indices is also assumed.
2. The Geometric Hamilton–Jacobi Theory
We summarize the main features of the geometric Hamilton–Jacobi theory for the Hamiltonian and Lagrangian formalisms of autonomous dynamical systems, as it is stated in Reference [
14] (also see Reference [
9,
15]).
2.1. Hamiltonian Hamilton–Jacobi Problem
Typically, a
(regular autonomous) Hamiltonian system is a triad
, where the bundle
represents the
phase space of a dynamical system (
Q is the
configuration space),
is the natural symplectic form in
, and
is the
Hamiltonian function. The dynamical trajectories are the integral curves
of the
Hamiltonian vector field associated with
H, which is the solution to the
Hamiltonian equation(Here,
and
are the sets of differentiable
k-forms and vector fields in
, and
denotes the inner contraction of
and
). In natural coordinates
of
, we have that
, and the curves
are the solution to the
Hamilton equationsDefinition 1. The generalized Hamiltonian Hamilton–Jacobi problem for a Hamiltonian systemis to find a vector fieldand a 1-formsuch that, ifis an integral curve of X, thenis an integral curve of, i.e., if, then. Then, the coupleis a solution to the generalized Hamiltonian Hamilton–Jacobi problem.
Theorem 1. The following statements are equivalent:
- 1.
The coupleis a solution to the generalized Hamiltonian Hamilton–Jacobi problem.
- 2.
The vector fields X andare α-related, i.e.,. As a consequence,, and it is called the vector field associated with the form.
- 3.
The submanifoldofis invariant by the Hamiltonian vector field(or, which means the same thing,is tangent to).
- 4.
The integral curves ofwhich have their initial conditions inproject onto the integral curves of X.
- 5.
The equationholds for the 1-form α.
Proof. (Guidelines for the proof):
The equivalence between 1 and 2 is a consequence of the Definition 1 and the definition of integral curves. Then, the expression is obtained by composing both members of the equality with and taking into account that .
Items 3 and 4 follow from 2.
Item 5 is obtained from Definition 1 and using the dynamical Equation (1). □
In order to solve the generalized Hamilton–Jacobi problem, it is usual to state a less general version of it, which constitutes the standard Hamilton–Jacobi problem.
Definition 2. The Hamiltonian Hamilton–Jacobi problem for a Hamiltonian systemis to find a 1-formsuch that it is a solution to the generalized Hamiltonian Hamilton–Jacobi problem and is closed. Then, the form α is a solution to the Hamiltonian Hamilton–Jacobi problem.
As is closed, for every point in Q, there is a function S in a neighborhood such that . It is called a local generating function of the solution .
Theorem 2. The following statements are equivalent:
- 1.
The formis a solution to the Hamiltonian Hamilton–Jacobi problem.
- 2.
is a Lagrangian submanifold ofwhich is invariant by, and S is a local generating function of this Lagrangian submanifold.
- 3.
The equationholds for α, or, which is equivalent, the functionis locally constant.
Proof. (Guidelines for the proof): They are consequences of Theorem 1 and Definition 2. □
The last condition, written in natural coordinates, gives the classical form of the Hamiltonian Hamilton–Jacobi equation, which is
These forms are particular solutions to the (generalized) Hamilton–Jacobi problem, but we are also interested in finding complete solutions to the problem. Then,
Definition 3. Let. A family of solutions, depending on n parameters, is a complete solution to the Hamiltonian Hamilton–Jacobi problem if the mapis a local diffeomorphism. Remark 1. Given a complete solution, as,, there is a family of functionsdefined in neighborhoodsof every point such that. Therefore, we have a locally defined functionwhich is called alocal generating functionof the complete solution.
A complete solution defines a Lagrangian foliation inwhich is transverse to the fibers, and such thatis tangent to the leaves. The functions that locally define this foliation are the components of a mapand give a family of constants of motion of. Conversely, if we havenfirst integralsofin involution, such that, then, with, define this transversal Lagrangian foliation, hence having a local complete solution. Thus, we can locally isolate, replace them in, and project to the basis, then obtaining the family of vector fieldsassociated with the local complete solution. Ifis a complete solution, then all the integral curves ofare obtained from the integral curves of the associated vector fields.
2.2. Lagrangian Hamilton–Jacobi Problem
The above framework for the Hamilton–Jacobi theory can be easily translated to the Lagrangian formalism of mechanics. Now, the phase space is the tangent bundle
of the configuration bundle
Q and the dynamics is given by the Lagrangian function of the system,
. Using the canonical structures in
, i.e., the
vertical endomorphism , and the
Liouville vector field , the Lagrangian forms
,
, and the Lagrangian energy
are constructed. Then, the
Lagrangian equation is
and
is a
Lagrangian dynamical system. Furthermore, the
Legendre transformation associated with
, denoted by
, is defined as the fiber derivative of the Lagrangian function. We assume that
is regular, i.e.,
is a local diffeomorphism, or, equivalently,
is a symplectic form (the Lagrangian is
hyper-regular if
is a global diffeomorphism). In that case, the Lagrangian Equation (
3) has a unique solution
, which is called the
Lagrangian vector field, in which integral curves are holonomic, and are the solutions to the Euler–Lagrange equations. (see Reference [
49] for details).
Definition 4. The generalized Lagrangian Hamilton–Jacobi problem for a Lagrangian systemis to find a vector fieldsuch that, ifis an integral curve of X, thenis an integral curve of, i.e., if, then. Then, the vector field X is a solution to the generalized Lagrangian Hamilton–Jacobi problem.
Theorem 3. The following statements are equivalent:
- 1.
The vector field X is a solution to the generalized Lagrangian Hamilton–Jacobi problem.
- 2.
The vector fields X andare X-related, i.e.,.
- 3.
The submanifold of is invariant by the Lagrangian vector field (or, which means the same thing, is tangent to ).
- 4.
The integral curves ofwhich have their initial conditions inproject onto the integral curves of X.
- 5.
The equationholds for the vector field X.
Proof. (Guidelines for the proof): The proof follows the same patterns as Theorem 1. □
As in the Hamiltonian formalism, we consider the following simpler case:
Definition 5. The Lagrangian Hamilton–Jacobi problem for a Lagrangian systemis to find a vector field X such that it is a solution to the generalized Lagrangian Hamilton–Jacobi problem and satisfies that. Then, this vector field X is a solution to the Lagrangian Hamilton–Jacobi problem.
Since , then, for every point of Q, there is a neighborhood and a function S such that (in U).
Theorem 4. The following statements are equivalent:
- 1.
The vector field X is a solution to the Lagrangian Hamilton–Jacobi problem.
- 2.
is a Lagrangian submanifold ofwhich is invariant by the Lagrangian vector field(and S is a local generating function of this Lagrangian submanifold).
- 3.
The equationholds for X, or, which is equivalent, the functionis locally constant.
Proof. (Guidelines for the proof): They are consequences of Theorem 3 and Definition 5. □
The last condition leads to the following expression which is the form of the Lagrangian Hamilton–Jacobi equation in natural coordinates,
As in the Hamiltonian Hamilton–Jacobi theory, we are interested in the complete solutions to the problem, which are defined as:
Definition 6. Let. A family of solutions, depending on n parameters, is a complete solution to the Lagrangian Hamilton–Jacobi problem if the mapis a local diffeomorphism. If we have a complete solution to the Lagrangian Hamilton–Jacobi problem, all the integral curves of the Lagrangian vector field are obtained from the integral curves of all the vector fields .
The equivalence between the Lagrangian and the Hamiltonian Hamilton–Jacobi problems is stated as follows:
Theorem 5. Letbe a (hyper)regular Lagrangian system, andits associated Hamiltonian system. Ifis a solution to the (generalized) Hamiltonian Hamilton–Jacobi problem, thenis a solution to the (generalized) Lagrangian Hamilton–Jacobi problem; conversely, ifis a solution to the (generalized) Lagrangian Hamilton–Jacobi problem, thenis a solution to the (generalized) Hamiltonian Hamilton–Jacobi problem.
Proof. (Guidelines for the proof): It can be proven that ; then, bearing in mind that , the proof follows using items 2 and 5 of Theorems 1 and 3 (or item 3 of Theorems 2 and 4). □
4. General Geometric Framework for the Hamilton–Jacobi Theory
The geometric Hamilton–Jacobi theory can be stated in a more general framework which allows us to extend the theory to a wide variety of systems and situations. Next, we present a summary of this general framework, as it is stated in Reference [
36] (also see Reference [
31] for another similar approach).
4.1. Slicing Problems
In general, a dynamical system is just a couple , where P is a manifold and is a vector field which defines the dynamical equation on P. Then, in order to state the analogous to the Hamilton–Jacobi problem for this system in a more general context, consider a manifold , a vector field , and a map , as it is showed in the following diagram:
Proposition 3. The following statements are equivalent:
- 1.
If γ is an integral curve of X, thenis an integral curve of Z.
- 2.
The vector fields X and Z are α-related:Furthermore, if α is an injective immersion, (inducing a diffeomorphism), then these properties are equivalent to: - 3.
The vector field Z is tangent to, and, if, then.
Then, the mapis a bijection between the integral curves of X and the integral curves of Z in.
Proof. They are immediate, bearing in mind the commutativity of the above diagram. □
Definition 10. A slicing of a dynamical systemis a triplewhich is a solution to the slicing Equation (9).
If
and
are coordinates in
and
P, respectively, and
,
, and
, then
, and
is a solution to the slicing equation if, and only if,
We say that the vector field X gives a “partial dynamics” or a “slice” of the “whole dynamics” which is given by Z, and the whole dynamics can be recovered from these slices. In fact, the integral curves of Z contained in can be described by a solution to the slicing equation; but we need a complete solution to describe all the integral curves of Z, and it can be defined as a family of solutions depending on the parameters of a space .
Definition 11. A complete slicing of a dynamical systemis a mapand a vector fieldalong the projectionsuch that:
- 1.
The mapis surjective,
- 2.
for every, the mapandare a slicing of Z.
Thus, a complete slicing is a family of maps and vector fields satisfying the above conditions.
As for every point there exists such that , the integral curve of Z through p is described by the integral curve of through x by means of the map . In addition, if each is an immersion (for instance, when it is a diffeomorphism), then are determined by the .
The hypothesis of
being an embedding holds in many situations, for instance, for the sections of a fiber bundle
. Then, we can consider the slicing problem for sections
of
, as before. In this case, as
is an embedding, Equation (9) determines
X, and
X is given from
by the equation
In this case, Proposition 3 states that a section
of
is a solution to the slicing equation for
if, and only if,
4.2. Recovering the Hamilton–Jacobi Equation for Hamiltonian and Lagrangian Dynamical Systems
Consider the case of a Hamiltonian system , where is a symplectic manifold, is a Hamiltonian function, and is its Hamiltonian vector field, i.e., the solution to (1). Then,
Theorem 7. Ifis a solution to the slicing Equation (9) for, thenIn addition, ifis an embedding satisfying the condition, thenconversely, ifand α satisfies this equation and, then α is a solution to the slicing Equation (9). In the particular case where is a fiber bundle (for instance, and ), we can consider the slicing problem as before, but only for sections of ,
Being
an embedding, the Equation (9) determines
X, and, composing this equation with
, we obtain that
. Therefore, the slicing Equation (9) reads
In this way, Equation (9) can be considered as a generalization of the Hamilton–Jacobi equation in the Hamiltonian formalism, which is just the slicing equation for a closed 1-form in Q. Therefore, locally, and the slicing equation looks in the ordinary form .
The same applies to the Lagrangian formalism. In this case , and, if is a regular Lagrangian function, is the Lagrangian vector field solution to the Lagrangian Equation (3). Then, all proceeds as in the Hamiltonian case.
The Definitions 3 and 6 of complete solutions to the Hamiltonian and Lagrangian Hamilton–Jacobi problems, respectively, are particular cases of the Definition 11 of complete slicings.
5. The Hamilton–Jacobi Problem for Other Physical Systems
Using the general framework presented in the above section, the Hamilton–Jacobi problem can be stated for a wide kind of physical systems. Next, we review two of them. Other applications of the theory are listed in detail in the Introduction.
5.1. Higher-Order (Autonomous) Dynamical Systems
Let
Q be a
n-dimensional manifold and let
the
kth-order tangent bundle of
Q, which is endowed with natural coordinates
,
,
. If
is the Lagrangian function of an autonomous
kth-order Lagrangian system, using the canonical structures of the higher-order tangent bundles, we can construct the Poincaré-Cartan forms and the Lagrangian energy in which coordinate expressions are
where
, and
. Thus, we have the higher-order Lagrangian system
. Assuming that the Lagrangian function is regular, i.e.,
is a symplectic form, the Lagrangian equation
has a unique solution
(the Lagrangian vector field) in which the integral curves are holonomic (i.e., they are canonical liftings
of curves
) and are the solutions to the
Otrogradskii or
higher-order Euler–Lagrange equations (see Reference [
50,
51,
52] for details).
The Hamilton–Jacobi problem for higher-order Lagrangian dynamical systems is just the slicing problem for the particular situation represented in the diagram
i.e., for sections of the natural projection
,
; thus, we have the following settings (see Reference [
20,
21] for the details and proofs):
Definition 12. The generalized kth-order Lagrangian Hamilton–Jacobi problem for the higher-order Lagrangian systemis to find a sectionand a vector fieldsuch that, ifis an integral curve of X, thenis an integral curve of; i.e., if, then. Then, the coupleis a solution to the generalized kth-order Lagrangian Hamilton–Jacobi problem.
Theorem 8. The following statements are equivalent:
- 1.
The coupleis a solution to the generalized kth-order Lagrangian Hamilton–Jacobi problem.
- 2.
The vector fields X andare s-related, i.e.,. As a consequence,, and X is said to be the vector field associated with the sections.
- 3.
The submanifoldofis invariant by the Lagrangian vector field(or, which means the same thing,is tangent to).
- 4.
The integral curves ofwhich have initial conditions inproject onto the integral curves of X.
- 5.
The equationholds for α.
Proof. (Guidelines for the proof): The proof follows a pattern similar to that of Theorem 1, but now using Definition 12. □
Definition 13. The kth-order Lagrangian Hamilton–Jacobi problem for the higher-order Lagrangian systemis to find a sectionsuch that it is a solution to the generalized kth-order Lagrangian Hamilton–Jacobi problem and satisfies that. Then, this section s is a solution to the kth-order Lagrangian Hamilton–Jacobi problem.
Observe that that ; i.e., is a closed 1-form and then there exists , , such that .
Theorem 9. The following statements are equivalent:
- 1.
The section s is a solution to the generalized kth-order Lagrangian Hamilton–Jacobi problem.
- 2.
is a Lagrangian submanifold of, which is invariant by the Lagrangian vector field(and S is a local generating function of this Lagrangian submanifold).
- 3.
The equationholds for s, or, which is equivalent, the functionis locally constant.
Proof. (Guidelines for the proof): They are consequences of Theorem 8 and Definition 13. □
In natural coordinates, from this last condition, we obtain that
This system of partial differential equations for S generalizes Equation (4) to higher-order systems.
Definition 14. Let. A family of solutions, depending on n parameters, is a complete solution to the kth-order Lagrangian Hamilton–Jacobi problem if the mapis a local diffeomorphism. For the Hamiltonian formalism, let be the Hamiltonian function of a (regular) higher-order dynamical system. Using the canonical Liouville forms of the cotangent bundle, and , where () are canonical coordinates in , the dynamical equation for the Hamiltonian system is , and it has a unique solution . As we are working in the cotangent bundle , the Hamiltonian Hamilton–Jacobi problems for higher-order systems is stated in the same way as in the first-order case; hence, it is the slicing problem for the particular situation represented in the diagram
Therefore, all the definitions and results are like in the first-order case, and the relation between both the Lagrangian and the Hamiltonian Hamilton–Jacobi problems is stated as in Theorem 5.
5.2. Multisymplectic Field Theories
The Hamilton–Jacobi theory for multisymplectic field theories has been studied in Reference [
42,
43,
45]. Next, we state the Lagrangian and the Hamiltonian problems for these systems. For details on multisymplectic field theories, see, for instance, Reference [
53,
54,
55] and the references therein.
5.2.1. Multisymplectic Lagrangian Hamilton–Jacobi Problem
Let
a bundle, where
M is an oriented manifold with
and
. The Lagrangian description of multisymplectic classical field theories is stated in the first-order jet bundle
, which is also a bundle
. Natural coordinates in
adapted to the bundle structure are
(
;
). Giving a Lagrangian density associated to a Lagrangian function
and using the canonical structures of
, we can define the
Poincaré–Cartan forms associated with
,
, and
, in which local expression is
where
and
. The Lagrangian function is regular if
is a multisymplectic
-form (i.e., 1-nondegenerate). Then, the couple
is a
multisymplectic Lagrangian system. The
Lagrangian problem consists of finding
m-dimensional,
-transverse, and holonomic distributions
in
such that their integral sections
are canonical liftings
of sections
that are solutions to the Lagrangian field equation
In coordinates, the components of
satisfy the
Euler–Lagrange equationsDefinition 15. The generalized Lagrangian Hamilton–Jacobi problem for the multisymplectic Lagrangian systemis to find a section(which is called a jet field ) and an m-dimensional integrable distributionin E such that, ifis an integral section of, thenis an integral section of; i.e., if, for every, then, for every. Then, the coupleis a solution to the generalized Hamiltonian Hamilton–Jacobi problem.
Remark 2. The Hamilton–Jacobi problem can also be stated associating the distributionsandwithmultivector fields. Anm-multivector field, on a manifoldis a section of the bundle, where(i.e, a skew-symmetric contravariant tensor field). Ifis anm-multivector field in, then, for every, there is a neighborhoodand local vector fieldssuch that. Then, ifis anm-dimensional distribution in, sections ofarem-multivector fields in, and a multivector field isintegrableif its associated distribution is also.
Now, if, letandbe them-multivector fields associated with the distributionsand, respectively, then, the Lagrangian Hamilton–Jacobi problem can be represented by the diagram
whereanddenote the natural extensions of the mapsandto the multitangent bundles; thus, this problem can be considered as a special case of aslicing problem.
Theorem 10. The following statements are equivalent:
- 1.
The coupleis a solution to the generalized Lagrangian Hamilton–Jacobi problem.
- 2.
The distributionsandare Ψ-related. As a consequence,and is called the distribution associated with.
- 3.
The distributionis tangent to the submanifoldof.
- 4.
Integral sections ofwhich have boundary conditions inproject onto the integral sections of.
- 5.
If γ is an integral section of the distributionassociated with the jet field Ψ, then, for every, the equationholds for Ψ.
Proof. (Guidelines for the proof):
The equivalence between 1 and 2 is a consequence of the Definition 15, the equivalence between distributions and multivector fields, and the definition of integral sections.
Items 3 and 4 follow from 2.
Item 5 is obtained from Definition 15 and using field Equation (10). □
Definition 16. The Lagrangian Hamilton–Jacobi problem for the multisymplectic Lagrangian systemis to find a jet fieldsuch that it is solution to the generalized Lagrangian Hamilton–Jacobi problem and satisfies that. Then, the jet field Ψ is a solution to the Lagrangian Hamilton–Jacobi problem.
The condition is equivalent to asking that the form is closed and then there exists a -form , with , such that . Furthermore, is -semibasic, since , and, hence, , are also.
Theorem 11. The following statements are equivalent:
- 1.
The jet field Ψ is a solution to the Lagrangian Hamilton–Jacobi problem.
- 2.
is an m-Lagrangian submanifold ofand the distributionis tangent to it.
- 3.
The formis closed.
In coordinates,
, and the Lagrangian Hamilton–Jacobi equation has the form
Definition 17. Let. A family of solutions, depending on n parameters, is a complete solution to the Lagrangian Hamilton–Jacobi problem if the mapis a local diffeomorphism. A complete solution defines an -dimensional foliation in which is transverse to the fibers and such that the distribution is tangent to it. Then, all the sections which are solutions to the Euler–Lagrange equations (i.e., all the integral sections of the distribution ) are recovered from a complete solution.
5.2.2. Multisymplectic Hamiltonian Hamilton–Jacobi Problem
The Hamiltonian formalism for a regular first-order multisymplectic field theory is developed in the so-called
reduced dual jet bundle of ,
, where
is the bundle of
m-forms over
E vanishing when they act on
-vertical bivectors. It is endowed with the canonical projections
and
, and natural coordinates in
are denoted
. The physical information is given by a
Hamiltonian section h of the natural projection
, which is associated with a
local Hamiltonian function such that
. Then, from the canonical form
, we construct the Hamilton-Cartan multisymplectic form
in which coordinate expression is
and the couple
is a
multisymplectic Hamiltonian system. Then, the
Hamiltonian problem consists of finding integrable
m-dimensional
-transverse distributions
in
such that their integral sections
are solutions to the Hamiltonian field equation
The existence of such distributions
is assured. In coordinates, this equation gives the
Hamilton–De Donder–Weyl equationsDefinition 18. The generalized Hamiltonian Hamilton–Jacobi problem for the multisymplectic Hamiltonian systemis to find a sectionand an integrable m-dimensional distributionin E such that, ifis an integral section of, thenis an integral section of, i.e., if, for every, then, for every. Then, the coupleis a solution to the generalized Hamiltonian Hamilton–Jacobi problem.
Remark 3. As in the Lagrangian case, the Hamiltonian Hamilton–Jacobi problem can be considered as a special case of the followingslicing problem:
where and are m-multivector fields associated with the distributions and , respectively.
The following Theorems and Definitions are analogous to those of the Lagrangian case.
Theorem 12. The following conditions are equivalent.
- 1.
The coupleis a solution to the generalized Hamiltonian Hamilton–Jacobi problem.
- 2.
The distributionsandare s-related. As a consequence, the distributionis given by, and it is called the distribution associated withs.
- 3.
The distributionis tangent to the submanifoldof.
- 4.
Integral sections ofwhich have boundary conditions inproject onto the integral sections of.
- 5.
If γ is an integral section of the distributionassociated with s, then, for every, the equationholds for s.
Definition 19. The Hamiltonian Hamilton–Jacobi problem for the multisymplectic Hamiltonian systemis to find a sectionsuch that it is a solution to the generalized Hamilton–Jacobi problem and satisfies that. The section s is a solution to the Hamiltonianian Hamilton–Jacobi problem.
Theorem 13. The following conditions are equivalent.
- 1.
The coupleis a solution to the generalized Hamiltonian Hamilton–Jacobi problem.
- 2.
is an m-Lagrangian submanifold ofand the distributionis tangent to it.
- 3.
The formis closed.
As the
-semibasic
m-form
is closed, there exists a local
-semibasic
-form
, such that
. In coordinates, if
, where
are local functions, we obtain that
from which we obtain the classical Hamiltonian Hamilton–Jacobi equation
The definition and the characteristics of complete solution are like in the Lagrangian case.
5.2.3. Relation between the Multisymplectic Hamilton–Jacobi Problems
Let
be the Legendre map defined by the Lagrangian
, which is locally given by
If
is a regular or a hyperregular Lagrangian (i.e.,
is a local or global diffeomorphism), then
and
. In addition, the integral sections of the distributions
and
, which are the solution to the Lagrangian and the Hamiltonian problems, respectively, are in one-to-one correspondence through
. (see Reference [
43] for definitions and details). Then, we have:
Theorem 14. Letbe a regular or a hyperregular Lagrangian. Then, ifis a jet field solution to the (generalized) Lagrangian Hamilton–Jacobi problem, then the sectionis a solution to the (generalized) Hamiltonian Hamilton–Jacobi problem. Conversely, ifis a solution to the (generalized) Hamiltonian Hamilton–Jacobi problem, then the jet fieldis a solution to the (generalized) Lagrangian Hamilton–Jacobi problem.
Proof. (Guidelines for the proof): The proof is like in Theorem 5, but using multivector fields. □
Remark 4. As a final remark, notice that the Hamilton–Jacobi theory fornon-autonomous(i.e.,time-dependent)
dynamical systemscan be recovered from the multisymplectic Hamilton–Jacobi theory as a particular case takingand identifying the distributions,
,
and their associated multivector fields,
,
with time-dependent vector fields (see Reference [43]). 6. Discussion
In this work, the Lagrangian and the Hamiltonian versions of the Hamilton–Jacobi theory are reviewed from a modern geometric perspective.
First, this formulation is done for autonomous dynamical systems, and, in particular, the Hamiltonian case is compared with the “classical” Hamiltonian Hamilton–Jacobi theory, which is based in using canonical transformations.
There is also a general framework for the theory, which is also reviewed in the work. It contains the above standard theory for autonomous dynamical systems as a particular case and allows us to extend the Hamilton–Jacobi theory to a wide range of physical systems. In particular, two of these extensions have been analyzed here: the higher-order (autonomous) dynamical systems and the (first-order) classical Lagrangian and Hamiltonian field theories, using their multisymplectic formulation.
This geometric model has been extended and applied to many kinds of physical systems (as it is mentioned in the Introduction and cited in the bibliography). As a future line of research that has not been explored yet, the application of this geometric framework to state the Hamilton–Jacoby equation for dissipative systems in classical field theories should be explored, using an extension of the contact formalism which has been recently introduced to describe geometrically these kinds of dissipative field theories [
56,
57].