In this paper, we develop a Hamilton–Jacobi theory for forced Hamiltonian and Lagrangian systems. We study the complete solutions, particularize for Rayleigh systems, and present some examples. Additionally, we present a method for the reduction and reconstruction of the Hamilton–Jacobi problem for forced Hamiltonian systems with symmetry. Furthermore, we consider the reduction of the Hamilton–Jacobi problem for a Čaplygin system to the Hamilton–Jacobi problem for a forced Lagrangian system.

The classical formulation1–3 of the Hamilton–Jacobi problem for a Hamiltonian system on T*Q consists in looking for a function S on Q×R, called the principal function (also known as the generating function), such that

(1)

where H:T*QR is the Hamiltonian function. With the ansatzS(qi, t) = W(qi) − tE, where E is a constant, the equation above reduces to

(2)

where W:QR is the so-called characteristic function. Both Eqs. (1) and (2) are known as the Hamilton–Jacobi equation.

Despite the difficulties to solve a partial differential equation instead of a system of ordinary differential equations, i.e., to solve the Hamilton–Jacobi equation instead of Hamilton equations, the Hamilton–Jacobi theory provides a remarkably powerful method to integrate the dynamics of many Hamiltonian systems. In particular, for a completely integrable system, if one knows as much constants of the motion in involution as degrees of freedom of the system, one can obtain a complete solution of the Hamilton–Jacobi problem and completely solve the Hamiltonian system or, in other words, reduce it to quadratures.4–6 

Geometrically, the Hamilton–Jacobi equation (2) can be written as

where dW is a 1-form on Q. This 1-form transforms the integral curves of a vector field XHdW on Q into integral curves of the dynamical vector field XH on TQ (the latter satisfying Hamilton equations). This geometric procedure2,7 has been extended to many other different contexts, such as nonholonomic systems,7–11 singular Lagrangian systems,12–14 higher-order systems,15 field theories,16–22 or contact systems.23 An unifying Hamilton–Jacobi theory for almost-Poisson manifolds was developed in Ref. 24. The Hamilton–Jacobi theory has also been generalized to Hamiltonian systems with non-canonical symplectic structures,25 non-Hamiltonian systems,26 locally conformally symplectic manifolds,27 Nambu–Poisson and Nambu–Jacobi manifolds,28,29 Lie algebroids,30 and implicit differential systems.31 The applications of Hamilton–Jacobi theory include the relation between classical and quantum mechanics,32–34 information geometry,35,36 control theory,37 and the study of phase transitions.38 

In the same fashion, in this paper, we develop a Hamilton–Jacobi theory for systems with external forces. This paper is the natural continuation of our previous paper about symmetries and constants of the motion of systems with external forces39 (see also Ref. 40). Mechanical systems with external forces (so-called forced systems) appear commonly in engineering and describe certain physical systems with dissipation.39–41 Moreover, they emerge after a process of reduction in a nonholonomic Čaplygin system.9,10,42–45 A particular type of external forces is the so-called Rayleigh forces,39,40,46 i.e., forces that can be written as the derivative of a “potential” with respect to the velocities. Forced systems on a Lie group have been studied in Ref. 46.

This paper is organized as follows: In Sec. II, we recall the geometric concepts we will make use of. In Sec. III, we develop a Hamilton–Jacobi theory for Hamiltonian systems with external forces. We consider the complete solutions, particularized for Rayleigh forces, and discuss some examples. The analogous theory for Lagrangian systems with external forces is described in Sec. IV. In Sec. V, we present a method for the reduction and reconstruction of solutions of the Hamilton–Jacobi problem for forced Hamiltonian systems with symmetry. Finally, Sec. VI is devoted to the reduction of Čaplygin systems to forced Lagrangian systems in order to obtain solutions of the forced Hamilton–Jacobi problem and reconstruct solutions of the nonholonomic Hamilton–Jacobi problem.

This paper is a natural continuation of our previous paper.39 Let us briefly recall the notations and results we will make use of. See also Ref. 40.

Consider a fiber bundle π: EM. Let us recall2,47,48 that a 1-form β on E is called semibasic if

for all vertical vector fields Z on E. If (xi, ya) are fibered (bundle) coordinates, then the vertical vector fields are locally generated by {/∂ya}. Hence, β is a semibasic 1-form if it is locally written as

We shall particularize this definition for the cases of tangent and cotangent bundles. In what follows, let Q be an n-dimensional differentiable manifold. Given a morphism of fiber bundles,

one can define an associated semibasic 1-form47,48βD on TQ by

where vqTqQ,uvqTvq(TQ).

If locally D is given by

then

Conversely, given a semibasic 1-form β on TQ, one can define the following morphism of fiber bundles:

for every vq,wqTqQ,uwqTwq(TQ), with TτQ(uwq)=wq. In local coordinates, if

then

Here, (qi,q̇i) are bundle coordinates in TQ.

Hence, there exists a one-to-one correspondence between semibasic 1-forms on TQ and fibered morphisms from TQ to T*Q.

An external force is geometrically interpreted as a semibasic 1-form on T*Q. A Hamiltonian system with external forces (the so-called forced Hamiltonian system) (H, β) is given by a Hamiltonian function H:T*QR and a semibasic 1-form β on T*Q. Let ωQ = −dθQ be the canonical symplectic form of T*Q. Locally these objects can be written as

where (qi, pi) are bundle coordinates in T*Q.

The dynamics of the system is given by the vector field XH,β, defined by

(3)

If XH is the Hamiltonian vector field for H, that is,

and Zβ is the vector field defined by

then we have

Locally, the above equations can be written as

(4)

Then, a curve qi(t), pi(t) in T*Q is an integral curve of XH,β if and only if it satisfies the forced Hamilton equations

(5)

Let us recall that the Poisson bracket is the bilinear operation,

(6)

with Xf, Xg being the Hamiltonian vector fields associated with f and g, respectively.

Definition 1.
Let (H, β) be a forced Hamiltonian system on T*Q. A function f on T*Q is called a constant of the motion (or a conserved quantity) if
(7)
or, equivalently, f is constant along the solutions of the forced Hamilton equations (5).

The PoincaréCartan 1-form on TQ associated with the Lagrangian function L:TQR is

(8)

where S* is the adjoint operator of the vertical endomorphism on TQ, which is locally

(9)

The Poincaré–Cartan 2-form is ωL = −dθL, so locally

(10)

One can easily verify that ωL is symplectic if and only if L is regular, that is, if the Hessian matrix

is invertible. The energy of the system is given by

where Δ is the Liouville vector field,

Similarly to the Hamiltonian framework, an external force is represented by a semibasic 1-form α on TQ. Locally,

The dynamics of the forced Lagrangian system (L, α) is given by

(11)

that is, the integral curves of the forced EulerLagrange vector fieldξL,α satisfy the forced Euler–Lagrange equations,

(12)

We can write ξL,α = ξL + ξα, where

This vector field is a second order differential equation (SODE), that is,

Definition 2.
Let (L, α) be a forced Lagrangian system on TQ. A function f on TQ is called a constant of the motion (or a conserved quantity) if
(13)
or, equivalently, f is constant along the solutions of the forced Euler–Lagrange equations (12).

An external force R̄ is said to be of Rayleigh type (or simply Rayleigh for short)39,40 if there exists a function R on TQ such that

which can be locally written as

This function R is called the Rayleigh dissipation function (or the Rayleigh potential). In other words, the fibered morphism DR̄:TQT*Q associated with R̄ is given by the fiber derivative of R (see Ref. 46), namely,

A Rayleigh system(L,R) is a forced Lagrangian system with Lagrangian function L and with external force R̄ generated by the Rayleigh potential R. For a Rayleigh system (L,R) with Rayleigh force R̄, the forced Euler–Lagrange vector field is denoted by ξL,R̄, given by

(14)

This vector field can be written as ξL,R̄=ξL+ξR̄, where

(15)

and the forced Euler–Lagrange equations (12) can be written as

(16)

Remark 1.

If a Rayleigh potential R on TQ defines a Rayleigh force R̄ on TQ, R+f also defines R̄ for any function f on Q. In other words, given a Rayleigh force R̄, its associated Rayleigh potential R is defined up to the addition of a basic function on TQ.

The vertical differentiation47 dS on T*Q is given by

where ιS denotes the vertical derivation, given by

for any function f, any p-form ω, and any vector fields X1, …, Xp on TQ. In particular,

for any function f on TQ. We can then write a Rayleigh force as

A linear Rayleigh forceR̄ is a Rayleigh force for which R is a quadratic form in the velocities, namely,

where Rij is symmetric and non-degenerate, and hence, the Rayleigh force is

In such a case, one can define an associated Rayleigh tensorRT*Q × T*Q, given by

A linear Rayleigh system (L, R) is a Rayleigh system such that R̄ is a linear Rayleigh force with Rayleigh tensor R.

The Legendre transformation is a mapping Leg: TQT*Q such that the diagram

commutes. Here, τq and πQ are the canonical projections on Q. Locally,

with pi=L/q̇i. In what follows, let us assume that the Lagrangian L is hyper-regular, i.e., that Leg is a (global) diffeomorphism.

Definition 3.
The dissipative bracket is a bilinear map [·, ·]: C(TQ) × C(TQ) → C(TQ) given by
(17)
where S is the vertical endomorphism and Xf is the Hamiltonian vector field associated with f on (TQ, ωL), namely,

Proposition 1.

The dissipative bracket [·, ·] on (TQ, ωL) verifies the following properties:

  • [f, g] = [g, f] (it is symmetric) and

  • [f, gh] = [f, h]g + [f, g]h(“Leibniz rule”)

for all functionsf, gonTQ.

Proof.
In local coordinates,
as one can derive from Eqs. (9) and (10). Here, (Wij) is the inverse matrix of the Hessian matrix (Wij) of the Lagrangian L. From this expression, both assertions can be easily proven.□

Since the dissipative bracket is bilinear and verifies the Leibniz rule, it is a derivation or a so-called Leibniz bracket.49 

Proposition 2.
Consider a Rayleigh system(L,R)on (TQ, ωL). A functionfonTQis a constant of the motion of(L,R)if and only if
(18)
where,is the Poisson bracket(6)defined byωL.

Proof.
As a matter of fact,
and
so
(19)
Here, ξL,R̄ is the dynamical vector field given by Eq. (14) and ξR̄ is given by Eq. (15). In particular, by Eq. (13), the right-hand side of Eq. (19) vanishes if and only if f is a constant of the motion.□

Remark 2.

Other types of dissipative systems, particularly thermodynamical systems, exhibit a “double bracket” dissipation, i.e., their dynamics are described in terms of two brackets [in our case, the Poisson bracket (6) and the dissipative bracket (17)]. As a matter of fact, the dissipative bracket we defined above has certain similarities with other types of brackets.

The dissipative bracket [·, ·] defined above resembles the dissipative bracket (·, ·) appearing in the metriplectic framework.50,51 Both brackets are symmetric and bilinear. However, the latter requires the additional assumption that (EL, f) vanishes identically for every function f on TQ. Clearly, this requirement does not hold for our dissipative bracket.

On the other hand, our dissipative bracket [·, ·] can also be related with the so-called Ginzburg–Landau bracket [·, ·]GL.52 This bracket, together with symmetry and bilinearity, satisfies the positivity condition, i.e., [f, f]GL ≥ 0 holds pointwisely for all f on TQ. As a matter of fact, this holds for our bracket in the case of many relevant Lagrangians. For instance, consider a Lagrangian of the form
with positive masses mi. Then,
See also Ref. 46 for various types of systems with double bracket dissipation. A further research on our dissipative bracket and its applications on thermodynamics will be done elsewhere.

Consider a natural LagrangianL on TQ, i.e., a Lagrangian function of the form

where

is a (pseudo)Riemannian metric on Q. Clearly, L is regular if and only if g is non-degenerate. As it is well-known, these are the usual Lagrangians in classical mechanics. The Legendre transformation is now linear,

where (gij)=(gij)1. In other words, the Legendre transformation consists simply in the “raising and lowering of indices” defined by the metric g.

Consider a linear Rayleigh system (L,R), where L is natural. The associated Hamiltonian Rayleigh potentialR̃ on T*Q is given by

Similarly, the Hamiltonian Rayleigh forceR̃ on T*Q is given by

When the Lagrangian is regular, the Legendre transformation is well-defined, so we can define a tensor field S̃T(T*Q)T*Q given by

In particular, if the Lagrangian is natural, then we have

Hence, the Hamiltonian Rayleigh force R̃ can be expressed in terms of the Rayleigh potential R̃ as

We shall omit the adjective Hamiltonian and refer to R̃ and R̃ as the Rayleigh potential and the Rayleigh force, respectively, if there is no danger of confusion.

Let us introduce the vertical differentiation47dS̃ on T*Q as

where ιS̃ is defined analogously to ιS by replacing S with S̃. In particular,

for any function f on T*Q. We can then write a Hamiltonian Rayleigh force as

Consider a linear Rayleigh system (L, R), where L is natural. The associated Hamiltonian Rayleigh potential R̃ on T*Q is given by

where

Similarly, the Hamiltonian Rayleigh force R̃ on T*Q is given by

where Rji=Rkjgik. The associated Hamiltonian Rayleigh tensor R̂TQT*Q is given by

The linear Hamiltonian Rayleigh force R̃ can thus be written as

This motivates the next definition of linear Rayleigh forces in the Hamiltonian framework, without the need of considering natural Lagrangians, as follows:

Definition 4.
An external force is called a linear Hamiltonian Rayleigh force if it can be written as
for some tensor R̂TQT*Q. This tensor is called the Hamiltonian Rayleigh tensor. A linear Hamiltonian Rayleigh system(H,R̂) is a forced Hamiltonian system whose external force is a linear Hamiltonian Rayleigh force. When there is no ambiguity, the adjective Hamiltonian will be omitted.

Remark 3.

Since T*Q, unlike TQ, has not a canonical vertical endomorphism, there is not a natural way to define Hamiltonian non-linear Rayleigh forces, besides Legendre-transforming Lagrangian Rayleigh forces.

Let (H, β) be a forced Hamiltonian system on T*Q. Its dynamical vector field XH,β is given by Eq. (3). Given a 1-form γ on Q (i.e., a section of πQ: T*QQ), it is possible to project XH,β along γ(Q), obtaining the vector field

(20)

on Q, so that the following diagram commutes:

Lemma 3.

The vector fieldsXH,βandXH,βγareγ-related if and only ifXH,βis tangent toγ(Q).

Proof.
By definition, XH,β and XH,βγ are γ-related if
(21)

Therefore, the integral curves of XH,βγ are mapped to integral curves of XH,β [which satisfy the forced Hamilton Eqs. (5)] via γ. Indeed, if σ is an integral curve of XH,βγ, then

so γσ is an integral curve of XH,β. Conversely, if γσ is an integral curve of XH,β for every integral curve σ of Y = QXH,βγ, then

for every integral curve σ, and hence, XH,β and Y are γ-related.

From Eq. (4), locally we have that

Then, Eq. (21) yields

In other words,

that is,

If γ is closed, we have

that is,

(22)

Let us recall that a Lagrangian submanifoldLT*Q is a maximal isotropic submanifold, i.e., a submanifold such that ωQ|S = 0 and dim S = 1/2 dim T*Q = dim Q. Clearly, a 1-form γ on Q is closed if and only if Im γ is a Lagrangian submanifold.

Definition 5.

A 1-form γ on Q is called a solution of the HamiltonJacobi problem for (H, β) if

  • it is closed and

  • it satisfies Eq. (22).

This equation is known as the HamiltonJacobi equation.

The results above can be summarized in the following theorem:

Theorem 4.

Letγbe a closed 1-form onQ. Then, the following conditions are equivalent:

  • γis a solution of the Hamilton–Jacobi problem for (H, β).

  • For every curveσ:RQsuch that
    for allt, thenγσis an integral curve ofXH,β.
  • Im γis a Lagrangian submanifold ofT*QandXH,βis tangent to it.

Remark 4.
By the Hamilton–Jacobi equation (22), dH + β vanishes on Im γ, and hence,
If dβ is non-degenerate, it is a symplectic form and Im γ is a Lagrangian submanifold on (T*Q, dβ). In general, it is not easy to see whether dβ is non-degenerate or not. However, there is a type of external forces for which this is simple: the Rayleigh forces linear in the momenta.

Lemma 5.

Let R̂ be a Hamiltonian Rayleigh tensor on T*Q, and let R̃ be the associated Hamiltonian Rayleigh force on T*Q. If R̂ is non-degenerate, then dR̃ is a symplectic form on T*Q.

Proof.
We have that
so
for a vector field X = Xi/∂qi + Yi/∂pi on T*Q. Then, X ∈ ker ΩQ,γ if and only if
If R̂ is non-degenerate (in particular, if R is positive-definite), then XkerdR̃ if and only if X = 0, so dR̃ is symplectic.□

When Rji does not depend on (qi), we can make the change of bundle coordinates,

so that dR̃=ωQ and (qi,p̃i) are Darboux coordinates.

Proposition 6.

Consider a linear Rayleigh system (H,R̂). Suppose that R̂ is non-degenerate. Then, a closed 1-form γ on Q is a solution of the Hamilton–Jacobi problem for (H,R̃) if and only if Im γ is a Lagrangian submanifold of (T*Q,dR̃).

For a Rayleigh system (H,R̃), the Hamilton–Jacobi equation can also be written as

(23)

We will also refer to the Hamilton–Jacobi problem for (H,R̄) as the Hamilton–Jacobi problem for (H,R). In the case of a linear Rayleigh system (H,R̂), we have

so the Hamilton–Jacobi equation (23) can be written as

(24)

We will also refer to the Hamilton–Jacobi problem for (H,R̄) as the Hamilton–Jacobi problem for (H,R̂).

One can consider a more general problem by relaxing the hypothesis of γ being closed.

Definition 6.

A weak solution of the HamiltonJacobi problem for (H, β) is a 1-form γ on Q such that XH,β and XH,βγ are γ-related. Here, XH,βγ is the vector field defined by (20).

Proposition 7.

Consider a 1-formγonQ. Then, the following statements are equivalent:

  • γis a weak solution of the Hamilton–Jacobi problem for (H, β).

  • γsatisfies the equation
    (25)
  • XH,βis tangent to the submanifold Im γT*Q.

  • Ifσ:RQsatisfies
    thenγσis an integral curve ofXH,β.

Let

Observe that Eq. (25) holds if and only if

Remark 5

(local expressions). Let γ be a 1-form γ on Q. Let H be a Hamiltonian function on T*Q, and let β,R̃, and R̂ be an external force, a Rayleigh potential, and a Rayleigh tensor on T*Q, respectively. If γ is closed, then

  • γ is a solution of the Hamilton–Jacobi problem for (H, β) if and only if
  • γ is a solution of the Hamilton–Jacobi problem for (H,R̃) if and only if
  • γ is a solution of the Hamilton–Jacobi problem for (H,R̂) if and only if
    Moreover, if γ is not necessarily closed, it is a weak solution of the Hamilton–Jacobi problem for (H, β) if and only if
    Similar expressions can be easily found for Rayleigh forces or linear Rayleigh forces.

The main interest in the standard Hamilton–Jacobi theory lies in finding a complete family of solutions to the problem.7,24 As it is explained below, knowing a complete solution of the Hamilton–Jacobi problem for a forced Hamiltonian system is tantamount to completely integrating the system, namely, there is a constant of the motion for each degree of freedom of the system, and these constants of the motion are in mutual involution.

Consider a forced Hamiltonian system (H, β) on T*Q and assume that dim Q = n.

Definition 7.

Let URn be an open set. A map Φ: Q × UT*Q is called a complete solution of the Hamilton–Jacobi problem for (H, β) if

  • Φ is a local diffeomorphism and

  • for any λ = (λ1, …, λn) ∈ U, the map
    is a solution of the Hamilton–Jacobi problem for (H, β).

For the sake of simplicity, we shall assume Φ to be a global diffeomorphism. Consider the functions given by

where πa denotes the projection over the ath component of Rn.

Proposition 8.

The functionsfaare constants of the motion. Moreover, they are in involution, i.e.,fa,fb=0, where,is the Poisson bracket defined byωQ.

Proof.
Given pT*Q, suppose that fa(p) = λa for each a = 1, …, n. Observe that
or, in other words,
By Theorem 4, XH,β is tangent to Im Φλ, and hence,
for every a = 1, …, n, that is, f1, …, fn are constants of the motion. In addition,

Example 1.
Consider a forced Hamiltonian system (H, β), with
Consider the functions
The dynamics of the system is given by
so that
and, thus, the functions are constants of the motion. Their Hamiltonian vector fields are given by
Clearly,
for every a, b = 1, …, n, so the constants of the motion are in involution. Consider the 1-form γ on Q given by
Clearly, γ is closed. In fact, it is exact,
Moreover,
and
Hence, γ is a complete solution of the Hamilton–Jacobi problem. When κi = 0 for every i = 1, …, n,
which is a complete solution for the (conservative) Hamilton–Jacobi problem for H. See Example 4 for the Lagrangian counterpart of this example.

Example 2
(free particle with a homogeneous linear Rayleigh force). Consider a linear Rayleigh system (H,R̂), with
where g does not depend on q (for instance, gij = miδij, with mi being the mass of the ith particle), and
where Rji does not depend on q. Then, the Hamilton–Jacobi equation (24) can be locally written as
so a complete solution is given by
where Rij=gjkRik. Clearly, γλ is closed; in fact, it is exact,

As it has been seen in Sec. III, the natural framework for the Hamilton–Jacobi theory is the Hamiltonian formalism on the cotangent bundle. Following the work of Cariñena, Gràcia, Marmo, Martínez, Muñoz-Lecanda, and Román-Roy,7,8,32,33,53 we introduce an analogous problem in the Lagrangian formalism on the tangent bundle as follows:

Definition 8.

A vector field X on Q is called a solution of the Lagrangian Hamilton–Jacobi problem for (L, α) if

  • Leg◦X is a closed 1-form and

  • X satisfies the equation
    (26)

    This equation is known as the Lagrangian HamiltonJacobi equation. When there is no risk of ambiguity, we shall refer to the Lagrangian Hamilton–Jacobi problem (respectively, equation) as simply the Hamilton–Jacobi problem (respectively, equation).

If γ = Leg◦X is a closed 1-form, then Im γ is a Lagrangian submanifold of (T*Q, ωQ). Therefore, Im X is a Lagrangian submanifold of (TQ, ωL). In other words, Leg◦X is closed if and only if X*ωL = 0. Moreover, it is easy to see that X and ξL,α are X-related, that is,

Analogously to the Hamiltonian case, one can show the following result (see also Refs. 7 and 9):

Proposition 9.

LetXbe a vector field onQthat satisfiesX*ωL = 0. Then, the following assertions are equivalent:

  • X is a solution of the Hamilton–Jacobi problem for (L, α).

  • Im Xis a Lagrangian submanifold ofTQinvariant byξL,α.

  • For every curveσ:RQsuch thatσis an integral curve ofX, thenXσ:RTQis an integral curve ofξL,α.

Remark 6.
For a Rayleigh system (L,R), the Hamilton–Jacobi equation (26) can be written as
If σ is an integral curve of X, then Xσ is an integral curve of ξL,R̄, which satisfies the forced Euler–Lagrange equations (16).

As in the Hamiltonian case, one can consider a more general problem by relaxing the hypothesis of Leg◦X being closed.

Definition 9.

A weak solution of the Hamilton–Jacobi problem for (L, α) is a vector field X on Q such that X and ξL,α are X-related.

Clearly, a weak solution of the Hamilton–Jacobi problem for (L, α) is a solution of the Hamilton–Jacobi problem for (L, α) if and only if X*ωL = 0.

Proposition 10.

LetXbe a vector field onQ. Then, the following statements are equivalent:

  • X is a solution of the generalized Hamilton–Jacobi problem for (L, α).

  • Xsatisfies the equation
    (27)
  • The submanifold Im XTQis invariant byξL,α.

  • Ifσ:RQis an integral curve ofX, thenXσis an integral curve ofξL,α.

Proof.
The last two assertions are trivial. Let us now prove the equivalence between the first and the second statements. From the dynamical equation (11), we have
Since X and ξL,α are X-related, we can write
which yields Eq. (27).

The proof of the converse is completely analogous to the one of Theorem 1 in Ref. 7.□

Given a forced Lagrangian system (L, α) on TQ (with L hyper-regular), one can obtain an associated forced Hamiltonian system (H, β) on T*Q, where

Moreover, the dynamical vector fields ξL,α and XH,β [given by Eqs. (3) and (11), respectively] are Leg-related, i.e.,

Theorem 11.

Consider a hyper-regular forced Lagrangian system (L, α) onTQ, with the associated forced Hamiltonian system (H, β) onT*Q. Then,Xis a (weak) solution of the Hamilton–Jacobi problem for (L, α) if and only ifγ = Leg◦Xis a (weak) solution of the Hamilton–Jacobi problem for (H, β).

Proof.
Let X be a weak solution of the Hamilton–Jacobi problem for (L, α). Then,
since X and ξL,α are X-related. Composing the left- and right-hand sides with Q from the left, we obtain
Then, X=XH,βγ, and γ is a weak solution of the Hamilton–Jacobi problem for (H, β).
Conversely, if γ is a solution of the Hamilton–Jacobi problem for (H, β), X is γ-related to XH,β. Moreover,
and hence,
so X is a weak solution of the Lagrangian Hamilton–Jacobi problem.

Obviously, the Lagrangian weak solution is a solution if and only if the associated Hamiltonian solution is a closed 1-form.□

This result could be extended for regular but not hyper-regular Lagrangians (i.e., Leg is a local diffeomorphism), where it only holds in the open sets where Leg is a diffeomorphism.

Complete solutions for the Hamilton–Jacobi problem are defined analogously to the ones in T*Q (see Definition 7).

Definition 10.

Let URn be an open set. A map Φ: Q × UTQ is called a complete solution of the HamiltonJacobi problem for (L, α) if

  • Φ is a local diffeomorphism and

  • for any λ = (λ1, …, λn) ∈ U, the map
    is a solution of the Hamilton–Jacobi problem for (L, α).

Example 3
(fluid resistance). Consider the one-dimensional Rayleigh system (L,R),39,40 with
Then,
is a complete solution of the Hamilton–Jacobi problem. Clearly, Im X is a Lagrangian submanifold. As a matter of fact,
and
The forced Euler–Lagrange vector field is given by
whose solutions are
with initial conditions q(0) = q0 and q̇(0)=q̇0. Similarly, the integral curves of X are given by
Indeed, the integral curves of ξL,R̄ are recovered by taking λ=mq̇0ekq0/m.

Example 4.
Consider the generalization of the previous example to n dimensions, namely,
Consider the functions
Locally,
and
from where it is easy to see that
so, by Eq. (18), fa are constants of the motion. In addition,
for every a, b = 1, …, n, so they are in involution. Then,
is a complete solution of the Hamilton–Jacobi problem. See Example 1 for the Hamiltonian counterpart of this example.

Example 5
(free particle with a homogeneous linear Rayleigh force). Consider a linear Rayleigh system (L,R), with
where g does not depend on q (for instance, gij = miδij, with mi being the mass of the ith particle), and
where Rij does not depend on q. Then, the Hamilton–Jacobi equation can be locally written as
so a complete solution is given by
where Rji=gikRjk. See Example 2 for the Hamiltonian counterpart of this example.

Let G be a connected Lie group acting freely and properly on Q by a left action Φ, namely,

As usual, we denote by g the Lie algebra of G and denote the dual of g by g*. For each gG, we can define a diffeomorphism

Under these conditions, the quotient space Q/G is a differentiable manifold and πG: QQ/G is a G-principal bundle. The action Φ induces a lifted action ΦT* on T*Q given by

for every αqTq*Q. Since Φ is a diffeomorphism, its lift to T*Q leaves θQ invariant;2 in other words, θQ is G-invariant.

The natural momentum mapJ:T*Qg is given by

for each ξg. Here, ξQ is the infinitesimal generator of the action of ξg on Q, and ξQc is the generator of the lifted action on T*Q. The natural momentum map is G-equivariant for the lifted action on T*Q. For each ξg, we have a function Jξ:T*QR given by

that is,

A vector field Z on Q defines an associated function ιZ on T*Q. Locally, if

then

Given a vector field X on Q, its complete lift54 is a vector field Xc on T*Q such that

for any vector field Z on Q. Locally, if X has the form above, then

Definition 11.
A G-invariant forced Hamiltonian system (H, β) is a forced Hamiltonian system (H, β) such that H and β are both G-invariant, namely,
and
for each ξg.

If H is known to be G-invariant, the subgroup Gβ such that (H, β) is Gβ-invariant can be found through the following lemma:

Lemma 12.

Consider a forced Hamiltonian system (H, β). Suppose thatHisG-invariant. Letξg. Then, the following statements hold:

  • Jξis a constant of the motion if and only if
  • If the previous equation holds, thenξleavesβinvariant if and only if
    Moreover, the vector subspace
    (28)
    is a Lie subalgebra ofg.

Proof.
Let us prove the first statement. We have that
where we have used that θQ is g-invariant. Contracting with the dynamical vector field XH,β [given by Eq. (3)] yields
since H is g-invariant, so, by Eq. (7), Jξ is a constant of the motion if and only if β(ξQc) vanishes. The proofs of the other statements are completely analogous to the ones of the Lagrangian counterpart of this lemma (see Refs. 39 and 40 and also Ref. 46).□

In the following paragraphs, we shall briefly recall some results we will make use of (see Refs. 55 and 56 and references therein for more details).

A G-invariant Lagrangian submanifoldLT*Q is a Lagrangian submanifold in T*Q such that ΦgT*(L)=L for all gG. Given the momentum map J defined above and a Lagrangian submanifold LT*Q, it can be shown that J is constant along L if and only if L is G-invariant.

The quotient space T*Q/G has a Poisson structure induced by the canonical symplectic structure on T*Q such that π: T*QT*Q/G is a Poisson morphism.

Denote by GβG the Lie subgroup whose Lie algebra is gβ, defined by (28). Let μg*. Let us assume that (Gβ)μ=G. Then, J−1(μ)/G is a symplectic leaf of T*Q/G. Moreover, J−1(μ) is a coisotropic submanifold. Assume that LJ1(μ) is a Lagrangian submanifold. Then, by the coisotropic reduction theorem,2π(L) is a Lagrangian submanifold to J−1(μ)/G.

Furthermore, it can be shown that J−1(μ)/G is diffeomorphic to the cotangent bundle T*(Q/G). In addition, if J−1(μ)/G is endowed with the symplectic structure ωμ given by the Marsden–Weinstein reduction procedure, it is symplectomorphic to T*(Q/G) endowed with a modified symplectic structure ω̃Q/G. This modified symplectic form is given by the canonical symplectic form plus a magnetic term, namely,

where ωQ/G is the canonical symplectic form on T*(Q/G).

Combining the previous paragraphs, π(L) can be seen as a Lagrangian submanifold of a cotangent bundle with a modified symplectic structure. Let g̃* denote the adjoint bundle to πG: QQ/G via the coadjoint representation

A connection A on πG: QQ/G induces a splitting

This identification is given by

where horq:TπG(q)(Q/G)Hq denotes the horizontal lift of the connection A. Here, Hq denotes the horizontal space of A at qQ, namely,

If αqJ−1(μ), then J(αq) = μ and Ψ([αq]) = (αq◦horqμ), so

Consider a G-invariant forced Hamiltonian system (H, β) on T*Q. Then, H = HGπ, where HG:T*Q/GR is the reduced Hamiltonian on T*Q/G. Similarly, β = π*βG, where βG is the reduced external force on T*Q/G. Moreover, we can define the reduced Hamiltonian H̃μ on T*(Q/G) by

for each (α̃q̃)Tq̃*(Q/G), where q̃=[q]Q/G and H̃=HGΨ1. Similarly, let β̃=(ψ1)*βG and

for each (α̃q̃)Tq̃*(Q/G).

Let γ be a G-invariant solution of the Hamilton–Jacobi problem for (H, β). Then, the following diagram commutes:

Proposition 13

(reduction). Letγbe aG-invariant solution of the Hamilton–Jacobi problem for (H, β). LetL=ImγandL̃=Ψπ(L). Then,γinduces a mappingγ̃μsuch thatImγ̃μ=L̃andγ̃μis a solution the Hamilton–Jacobi problem for(H̃μ,β̃μ).

Proof.
Since dH + β vanishes along L, clearly dHG + βG vanishes along π(L). If α̃q̃L̃, then ψ1(α̃q̃)π(L),
and
so
Since γ is G-invariant, it induces a mapping γ̃:QT*(Q/G), which is also G-invariant. This mapping, in turn, induces a reduced solution γ̃μ:Q/GT*(Q/G) such that γ̃=πG*γ̃μ and Imγ̃μ=L̃.□

Proposition 14
(reconstruction). LetL̃be a Lagrangian submanifold of (T*(Q/G), ωQ/G + Bμ) for someμg*, which is a fixed point of the coadjoint action. Assume thatL̃=Imγ̃μ, whereγ̃μis a solution of the Hamilton–Jacobi problem for(H̃μ,β̃μ). Let
and take
Then,
  • Lis aG-invariant Lagrangian submanifold ofT*Qand

  • L=Imγ, whereγis a solution of the Hamilton–Jacobi problem for (H, β).

Proof.
The mapping γ̃μ:Q/GT*(Q/G) induces a G-invariant mapping
which, in turn, induces a G-invariant mapping γ: QT*Q. Let α̃q̃L̃, so (α̃q̃,[μ]q̃)T*(Q/G)×Q/Gg̃* and
and then

Example 6
(Calogero–Moser system with a linear Rayleigh force). Consider the linear Rayleigh system (H,R̂) on T*R2, with
and
so
Consider the action of R on R2 given by
Clearly, (H,R̃) is invariant under the corresponding lifted action on T*Q. The momentum map is
Then,
We can identify J1(μ)/R with R2, with coordinates (q, p) and the natural projection
We can introduce a reduced Hamiltonian
and a reduced external force
Let γ̃ be a closed 1-form on R. Then,
so γ̃ is a solution of the Hamilton–Jacobi problem for (h, r) if and only if
and hence,
is a complete solution depending on the parameter λR. The associated generating function is
where, without loss of generality, we have taken the integration constant as zero. We can reconstruct a complete solution γλ of the Hamilton–Jacobi for (H,R̃), given by
and the associated generating function is

A nonholonomic mechanical system is given by a Lagrangian function L=LqA,q̇A subject to a family of constraint functions,

For the sake of simplicity, we shall assume that the constraints Φi are linear in the velocities, i.e., ΦiqA,q̇A=ΦAi(q)q̇A. Then, the nonholonomic equations of motion are

where λi=λiqA,q̇A,1im, are Lagrange multipliers to be determined.

Geometrically, the constraints are given by a vector sub-bundle M of TQ locally defined by Φi = 0. The dynamical equations can then be rewritten intrinsically as

Under certain compatibility conditions, the vector field X is unique and it is denoted by Xnh.

A Čaplygin system (also spelled as Chaplygin) is a nonholonomic mechanical system such that we have the following:

  • The configuration manifold Q is a fibered manifold, say, ρ: QN, over a manifold N.

  • The constraints are provided by the horizontal distribution of an Ehresmann connection Γ in ρ.

  • The Lagrangian L:TQR is Γ-invariant.

A particular case is when ρ: QN = Q/G is a principal G-bundle and Γ is a principal connection. As a matter of fact, some Refs. 42, 44, 45, and 57 restrict their definition of the Čaplygin system to this particular case. Our more general definition is also considered in Refs. 9 and 58.

Let us recall that the connection Γ induces a Whitney decomposition TQ=HVρ, where H is the horizontal distribution and = ker  is the vertical distribution. Take fibered coordinates qA=qa,qi such that ρqa,qi=qa.

With a slight abuse of notation, let hor:TQH denote the horizontal projector hereinafter and let Γai=ΓaiqA denote the Christoffel components of the connection Γ. Let us recall that Γ may be considered as a (1, 1)-type tensor field on Q with Γ2 = id, so hor = hor2 = (1/2)(id + Γ). The curvature of Γ is the (1, 2)-tensor field R=12[hor,hor], where [hor, hor] is the Nijenhuis tensor of hor,47 that is,

for each pair of vector fields X and Y on N. Locally,

where

The constraints are given by

that is, the solutions are horizontal curves with respect to Γ.

Since the Lagrangian is Γ-invariant,

for all YTyN, y = ρ(q1) = ρ(q2), where YH denotes the horizontal lift of Y to Q. We can then introduce a function on TN such that

so locally we have

Now let ρ(q) = y and xH with τQ(x) = q; let uTyN, UTu(TN) and XTx(TQ) such that X projects onto

We can then introduce a 1-form α on TN such that

where θL is the Poincaré-Cartan 1-form associated with L, given by Eq. (8). In other words, α is locally given by

It can be shown that is a regular Lagrangian and that the Čaplygin system is equivalent to the forced Lagrangian system (, α).

Assume that is hyper-regular. Then, the Čaplygin system has an associated forced Hamiltonian system (h, β), with

and

In particular, if L is natural, we have

The Hamilton–Jacobi equation for (h, β) is thus locally

In particular, if L is purely kinetical,

Let D denote a distribution on Q whose annihilator is

Then, we can form the algebraic ideal I(D0) in the algebra Λ*(Q), namely, if a k-form νI(D0), then

where βi ∈ Λk−1(q) and 1 ≤ im.

Theorem 15

[Ref. 9 (Theorem 4.3)]. LetHdenote the horizontal distribution defined by the connection Γ inρ: QN. LetXbe vector field onQsuch thatX(Q)Handd(LegX)I(H0). Then, the following conditions are equivalent:

  • For every curveσ:RQsuch that
    for allt, thenXσis an integral curve ofXnh.
  • dELXH0.

    A vector fieldXsatisfying these conditions is called a solution of the nonholonomic Hamilton–Jacobi problem for (L, Γ).

Theorem 16

[Ref. 9 (Theorem 4.5)]. Assume that a vector fieldXonQis a solution for the nonholonomic Hamilton–Jacobi problem for (L, Γ). IfXisρ-projectable to a vector fieldYonNandγ=Leg*Yis closed, thenYis a solution of the Lagrangian Hamilton–Jacobi problem for (, α) andγis a solution of the Hamilton–Jacobi problem for (h, β).

Conversely, letγbe a solution of the Hamilton–Jacobi problem for (h, β). Then,Y = Leg−1γis a solution of the Lagrangian Hamilton–Jacobi problem for (, α). If
then the horizontal liftYHis a solution for the nonholonomic Hamilton–Jacobi problem for (L, Γ).

Example 7

(mobile robot with fixed orientation). Consider the motion of a robot whose body maintains a fixed orientation with respect to the environment. The robot has three wheels with radius R, which turn simultaneously about independent axes and perform a rolling without sliding over a horizontal floor (see Refs. 9, 42, and 59 for more details).

Let (x, y) denote the position of the center of mass, and let θ and ψ denote the steering and rotation angles of the wheels, respectively. Hence, the configuration manifold is Q=S1×S1×R2, and the Lagrangian of the system is
Here, m is the mass, J is the moment of inertia, and Jω is the axial moment of inertia of the robot.
The constraints are induced by the conditions that the wheels roll without sliding, in the direction in which they point, and that the instantaneous contact point of the wheels with the floor have no velocity component orthogonal to that direction, so we have
The constraint distribution D is spanned by
The Abelian group G=R2 acts on Q by translations, namely,
Therefore, we have a principal G-bundle ρ: QN = Q/G with a principal connection given by the g-valued 1-form,
where g=R2 is the Lie algebra of G and e1,e2 is the canonical basis of R2 (identified with g).
One can show that the reduced forced mechanical system (, α) on TN is given by
and α is identically zero. A complete solution of the Hamilton–Jacobi problem for (, α) is given by
Its horizontal lift is
so
and thus,
Hence, YλH is a complete solution of the Hamilton–Jacobi problem for (L, Γ).

In this paper, we have obtained a Hamilton–Jacobi theory for Hamiltonian and Lagrangian systems with external forces. We have discussed the complete solutions of the Hamilton–Jacobi problem. Our results have been particularized for forces of Rayleigh type. We have presented a dissipative bracket for Rayleigh systems. Furthermore, we have studied the reduction and reconstruction of the Hamilton–Jacobi problem for forced Hamiltonian systems with symmetry. Additionally, we have shown how the Hamilton–Jacobi problem for a Čaplygin system can be reduced to the Hamilton–Jacobi problem for a forced Lagrangain system in order to obtain solutions of the latter and reconstruct solutions of the former.

In a previous paper, we studied the symmetries, conserved quantities, and reduction of forced mechanical systems (see Ref. 39, see also Ref. 40). Making use of results from this paper, one can obtain the constants of the motion in involution of a forced system and relate them with complete solutions of the Hamilton–Jacobi problem for that system (see Example 3). Furthermore, Lemma 15 from Ref. 39 has been translated to the Hamiltonian formalism (see Lemma 12) in order to extend the method of reduction of the Hamilton–Jacobi problem55 for forced Hamiltonian systems.

In another paper,60 we develop a Hamilton–Jacobi theory for forced discrete Hamiltonian systems. Our approach is based on the construction of a discrete flow on Q × Q (unlike the case without external forces,61 where the discrete flow is defined on Q). We define a discrete Rayleigh potential. Additionally, we present some simulations and analyze their numerical accuracy.

An additional open problem is the particularization of the results from this paper when the configuration space Q is a Lie group G with the Lie algebra g. If (L, α) is a G-invariant forced Lagrangian system on TG, the forced Euler–Lagrange equations for (L, α) are reduced to the Euler–Poincaré equations with forcing on g (see Ref. 46). We also plan to extend our results on forced systems to the Lie algebroid framework in order to use Atiyah algebroids when the system enjoys symmetries. Furthermore, we plan to extend the results from this paper for time-dependent forced Lagrangian systems in the framework of cosymplectic geometry (see Ref. 62). Additionally, the applications of the dissipative bracket (17) will be studied elsewhere.

We thank the referee for his/her constructive comments. The authors acknowledge financial support from the Spanish Ministry of Science and Innovation (MICINN) under Grant No. PID2019-106715GB-C21 and “Severo Ochoa Programme for Centres of Excellence in R&D” (Grant No. CEX2019-000904-S). Manuel Lainz wishes to thank MICINN and the Institute of Mathematical Sciences (ICMAT) for the FPI-Severo Ochoa predoctoral contract (No. PRE2018-083203). Asier López-Gordón would like to thank MICINN and ICMAT for the predoctoral contract (No. PRE2020-093814). He is also grateful for enlightening discussions on fiber bundles with his friend and colleague Alejandro Pérez-González.

The authors have no conflicts to disclose.

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

1.
H.
Goldstein
,
Mecánica Clásica
(
Editorial Reverté
,
1987
), google-books-ID: vf2JiybeDc4C.
2.
R.
Abraham
and
J.
Marsden
,
Foundations of Mechanics
(
AMS Chelsea Publishing/American Mathematical Society
,
2008
).
3.
V. I.
Arnold
,
Mathematical Methods of Classical Mechanics
, Graduate Texts in Mathematics (
Springer-Verlag
,
1978
).
4.
S.
Grillo
, “
Non-commutative integrability, exact solvability and the Hamilton–Jacobi theory
,”
Anal. Math. Phys.
11
,
71
(
2021
).
5.
S.
Grillo
,
J. C.
Marrero
, and
E.
Padrón
, “
Extended Hamilton–Jacobi theory, symmetries and integrability by quadratures
,”
Mathematics
9
,
1357
(
2021
).
6.
S.
Grillo
and
E.
Padrón
, “
A Hamilton–Jacobi theory for general dynamical systems and integrability by quadratures in symplectic and Poisson manifolds
,”
J. Geom. Phys.
110
,
101
129
(
2016
).
7.
J. F.
Cariñena
,
X.
Gràcia
,
G.
Marmo
,
E.
Martínez
,
M. C.
Muñoz-Lecanda
, and
N.
Román-Roy
, “
Geometric Hamilton–Jacobi theory
,”
Int. J. Geom. Methods Mod. Phys.
03
,
1417
1458
(
2006
).
8.
J. F.
Cariñena
,
X.
Gràcia
,
G.
Marmo
,
E.
Martínez
,
M. C.
Muñoz-Lecanda
, and
N.
Román-Roy
, “
Geometric Hamilton–Jacobi theory for nonholonomic dynamical systems
,”
Int. J. Geom. Methods Mod. Phys.
07
,
431
454
(
2010
).
9.
D.
Iglesias-Ponte
,
M.
de León
, and
D. M.
de Diego
, “
Towards a Hamilton–Jacobi theory for nonholonomic mechanical systems
,”
J. Phys. A: Math. Theor.
41
,
015205
(
2008
).
10.
T.
Ohsawa
,
O. E.
Fernandez
,
A. M.
Bloch
, and
D. V.
Zenkov
, “
Nonholonomic Hamilton–Jacobi theory via Chaplygin Hamiltonization
,”
J. Geom. Phys.
61
,
1263
1291
(
2011
).
11.
M.
de León
,
J.
Carlos Marrero
, and
D.
Martín de Diego
, “
Linear almost Poisson structures and Hamilton–Jacobi equation. Applications to nonholonomic mechanics
,”
J. Geom. Mech.
2
,
159
(
2010
).
12.
M.
de León
,
J.
Carlos Marrero
,
D.
Martín de Diego
, and
M.
Vaquero
, “
On the Hamilton–Jacobi theory for singular Lagrangian systems
,”
J. Math. Phys.
54
,
032902
(
2013
).
13.
M.
de León
,
D.
Martín de Diego
, and
M.
Vaquero
, “
A Hamilton–Jacobi theory for singular Lagrangian systems in the Skinner and Rusk setting
,”
Int. J. Geom. Methods Mod. Phys.
09
,
1250074
(
2012
).
14.
M.
Leok
,
T.
Ohsawa
, and
D.
Sosa
, “
Hamilton–Jacobi theory for degenerate Lagrangian systems with holonomic and nonholonomic constraints
,”
J. Math. Phys.
53
,
072905
(
2012
).
15.
L.
Colombo
,
M.
de León
,
P. D.
Prieto-Martínez
, and
N.
Román-Roy
, “
Geometric Hamilton–Jacobi theory for higher-order autonomous systems
,”
J. Phys. A: Math. Theor.
47
,
235203
(
2014
).
16.
M.
de León
,
J. C.
Marrero
, and
D.
Martín de Diego
, “
A geometric Hamilton–Jacobi theory for classical field theories
,” in
Variations, Geometry and Physics
(
Nova Science Publishing
, New York,
2009
), pp.
129
140
.
17.
M.
de León
and
S.
Vilariño
, “
Hamilton–Jacobi theory in k-cosymplectic field theories
,”
Int. J. Geom. Methods Mod. Phys.
11
,
1450007
(
2014
).
18.
L.
Vitagliano
, “
Geometric Hamilton–Jacobi field theory
,”
Int. J. Geom. Methods Mod. Phys.
09
,
1260008
(
2012
).
19.
M.
de León
and
M.
Zając
, “
Hamilton–Jacobi theory for gauge field theories
,”
J. Geom. Phys.
152
,
103636
(
2020
).
20.
M.
de León
,
P. D.
Prieto-Martínez
,
N.
Román-Roy
, and
S.
Vilariño
, “
Hamilton–Jacobi theory in multisymplectic classical field theories
,”
J. Math. Phys.
58
,
092901
(
2017
).
21.
C. M.
Campos
,
M.
de León
,
D.
Martín de Diego
, and
M.
Vaquero
, “
Hamilton–Jacobi theory in Cauchy data space
,”
Rep. Math. Phys.
76
,
359
387
(
2015
).
22.
V.
Zatloukal
, “
Classical field theories from Hamiltonian constraint: Canonical equations of motion and local Hamilton–Jacobi theory
,”
Int. J. Geom. Methods Mod. Phys.
13
,
1650072
(
2016
).
23.
M.
de León
,
M.
Lainz
, and
Á.
Muñiz-Brea
, “
The Hamilton–Jacobi theory for contact Hamiltonian systems
,”
Mathematics
9
(
16
),
1993
(
2021
).
24.
M.
de León
,
D.
Martín de Diego
, and
M.
Vaquero
, “
A Hamilton–Jacobi theory on Poisson manifolds
,”
J. Geom. Mech.
6
,
121
140
(
2014
).
25.
A. A.
Martínez-Merino
and
M.
Montesinos
, “
Hamilton–Jacobi theory for Hamiltonian systems with non-canonical symplectic structures
,”
Ann. Phys.
321
,
318
330
(
2006
).
26.
S.
Rashkovskiy
, “
Hamilton–Jacobi theory for Hamiltonian and non-Hamiltonian systems
,”
J. Geom. Mech.
12
,
563
(
2020
).
27.
O.
Esen
,
M.
de León
,
C.
Sardón
, and
M.
Zajşc
, “
Hamilton–Jacobi formalism on locally conformally symplectic manifolds
,”
J. Math. Phys.
62
,
033506
(
2021
).
28.
M.
de León
and
C.
Sardón
, “
Geometric Hamilton–Jacobi theory on Nambu–Poisson manifolds
,”
J. Math. Phys.
58
,
033508
(
2017
).
29.
M.
de León
and
C.
Sardón
, “
A geometric Hamilton–Jacobi theory on a Nambu–Jacobi manifold
,”
Int. J. Geom. Methods Mod. Phys.
16
,
1940007
(
2019
).
30.
M.
Leok
and
D.
Sosa
, “
Dirac structures and Hamilton–Jacobi theory for Lagrangian mechanics on Lie algebroids
,”
J. Geom. Mech.
4
,
421
(
2012
).
31.
O.
Esen
,
M.
de León
, and
C.
Sardón
, “
A Hamilton–Jacobi theory for implicit differential systems
,”
J. Math. Phys.
59
,
022902
(
2018
).
32.
J. F.
Carinena
,
X.
Gracia
,
E.
Martinez
,
G.
Marmo
,
M. C.
Munoz-Lecanda
, and
N.
Roman-Roy
, “
Hamilton–Jacobi theory and the evolution operator
,” in
Mathematical Physics and Field Theory. Julio Abad, in Memoriam
, 1 ed., edited by
M.
Asorey
,
J. V.
Garcia-Esteve
,
M. F.
Rañada
, and
J.
Sesma
(
Prensas Univ.
, Zaragoza,
2009
), p.
177
186
, http://hdl.handle.net/2117/3019.
33.
G.
Marmo
,
G.
Morandi
, and
N.
Mukunda
, “
The Hamilton–Jacobi theory and the analogy between classical and quantum mechanics
,”
J. Geom. Mech.
1
,
317
(
2009
).
34.
A.
Budiyono
, “
Quantization from Hamilton–Jacobi theory with a random constraint
,”
Physica A
391
,
4583
4589
(
2012
).
35.
F. M.
Ciaglia
,
F.
Di Cosmo
, and
G.
Marmo
, “
Hamilton–Jacobi theory and information geometry
,” in
Geometric Science of Information
, Lecture Notes in Computer Science, edited by
F.
Nielsen
and
F.
Barbaresco
(
Springer International Publishing
,
2017
), pp.
495
502
.
36.
F. M.
Ciaglia
,
F.
Di Cosmo
,
D.
Felice
,
S.
Mancini
,
G.
Marmo
, and
J. M.
Pérez-Pardo
, “
Hamilton–Jacobi approach to potential functions in information geometry
,”
J. Math. Phys.
58
,
063506
(
2017
).
37.
N.
Sakamoto
, “
Analysis of the Hamilton–Jacobi equation in nonlinear control theory by symplectic geometry
,”
SIAM J. Control Optim.
40
,
1924
1937
(
2002
).
38.
R. C.
Kraaij
,
F.
Redig
, and
W. B.
van Zuijlen
, “
A Hamilton–Jacobi point of view on mean-field Gibbs-non-Gibbs transitions
,”
Trans. Am. Math. Soc.
374
,
5287
5329
(
2021
).
39.
M.
de León
,
M.
Lainz
, and
A.
López-Gordón
, “
Symmetries, constants of the motion, and reduction of mechanical systems with external forces
,”
J. Math. Phys.
62
,
042901
(
2021
).
40.
A.
López-Gordón
, “
The geometry of Rayleigh dissipation
,” arXiv:2107.03780 (
2021
).
41.
O.
Esen
,
P.
Guha
, and
H.
Gümral
, “
On geometrical couplings of dissipation and curl forces
,” arXiv:2101.11884 (
2021
).
42.
F.
Cantrijn
,
J.
Cortés
,
M.
de León
, and
D. M.
de Diego
, “
On the geometry of generalized Chaplygin systems
,”
Math. Proc. Cambridge Philos. Soc.
132
,
323
351
(
2002
).
43.
L. C.
García-Naranjo
and
J. C.
Marrero
, “
The geometry of nonholonomic Chaplygin systems revisited
,”
Nonlinearity
33
,
1297
1341
(
2020
).
44.
A. M.
Bloch
,
P. S.
Krishnaprasad
,
J. E.
Marsden
, and
R. M.
Murray
, “
Nonholonomic mechanical systems with symmetry
,”
Arch. Ration. Mech. Anal.
136
,
21
99
(
1996
).
45.
J.
Koiller
, “
Reduction of some classical non-holonomic systems with symmetry
,”
Arch. Ration. Mech. Anal.
118
,
113
148
(
1992
).
46.
A.
Bloch
,
P. S.
Krishnaprasad
,
J. E.
Marsden
, and
T. S.
Ratiu
, “
The Euler-Poincaré equations and double bracket dissipation
,”
Commun. Math. Phys.
175
,
1
42
(
1996
).
47.
M.
de León
and
P. R.
Rodrigues
,
Methods of Differential Geometry in Analytical Mechanics
, North-Holland Mathematics Studies Vol. 158 (
North-Holland; Distributors for the USA and Canada, Elsevier Science Pub. Co.
,
1989
).
48.
C.
Godbillon
,
Géométrie Différentielle et Mécanique Analytique
, Collection Méthodes (
Hermann
,
1969
).
49.
J.-P.
Ortega
and
V.
Planas-Bielsa
, “
Dynamics on Leibniz manifolds
,”
J. Geom. Phys.
52
,
1
27
(
2004
).
50.
B.
Coquinot
and
P. J.
Morrison
, “
A general metriplectic framework with application to dissipative extended magnetohydrodynamics
,”
J. Plasma Phys.
86
,
835860302
(
2020
).
51.
P.
Birtea
,
M.
Boleantu
,
M.
Puta
, and
R. M.
Tudoran
, “
Asymptotic stability for a class of metriplectic systems
,”
J. Math. Phys.
48
,
082703
(
2007
).
52.
M.
Grmela
and
H. C.
Öttinger
, “
Dynamics and thermodynamics of complex fluids. I. Development of a general formalism
,”
Phys. Rev. E
56
,
6620
6632
(
1997
).
53.
J. F.
Cariñena
,
X.
Gràcia
,
G.
Marmo
,
E.
Martínez
,
M. C.
Muñoz-Lecanda
, and
N.
Román-Roy
, “
Structural aspects of Hamilton–Jacobi theory
,”
Int. J. Geom. Methods Mod. Phys.
13
,
1650017
(
2016
).
54.
K.
Yano
and
S.
Ishihara
,
Tangent and Cotangent Bundles: Differential Geometry
(
Marcel Dekker, Inc.
,
1973
).
55.
M.
de León
,
D.
Martín de Diego
, and
M.
Vaquero
, “
Hamilton–Jacobi theory, symmetries and coisotropic reduction
,”
J. Math. Pures Appl.
107
,
591
614
(
2017
).
56.
J. E.
Marsden
,
G.
Misiolek
,
J.-P.
Ortega
,
M.
Perlmutter
, and
T.
Ratiu
,
Hamiltonian Reduction by Stages
, Lecture Notes in Mathematics (
Springer-Verlag
,
2007
).
57.
F.
Cantrijn
,
M.
de León
,
J. C.
Marrero
, and
D.
Martín de Diego
, “
Reduction of constrained systems with symmetries
,”
J. Math. Phys.
40
,
795
820
(
1999
).
58.
M.
de León
and
D.
Martín de Diego
, “
On the geometry of non-holonomic Lagrangian systems
,”
J. Math. Phys.
37
,
3389
3414
(
1996
).
59.
S. D.
Kelly
and
R. M.
Murray
, “
Geometric phases and robotic locomotion
,”
J. Rob. Syst.
12
,
417
431
(
1995
).
60.
M.
de León
,
M.
Lainz
, and
A.
López-Gordón
, “
Discrete Hamilton–Jacobi theory for systems with external forces
,” arXiv:2110.14431 (
2021
).
61.
M.
de León
and
C.
Sardón
, “
Geometry of the discrete Hamilton–Jacobi equation: Applications in optimal control
,”
Rep. Math. Phys.
81
,
39
63
(
2018
).
62.
L. J.
Colombo
,
M.
de León
,
M. E.
Eyrea Irazú
, and
A.
López-Gordón
, “
Symplectic and cosymplectic reduction for simple hybrid forced mechanical systems with symmetries
,” arXiv:2112.02573 (
2021
).