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.
I. INTRODUCTION
The classical formulation1–3 of the Hamilton–Jacobi problem for a Hamiltonian system on T*Q consists in looking for a function S on , called the principal function (also known as the generating function), such that
where is the Hamiltonian function. With the ansatzS(qi, t) = W(qi) − tE, where E is a constant, the equation above reduces to
where 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 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.
II. PRELIMINARIES
A. Semibasic forms and fibered morphims
Consider a fiber bundle π: E → M. 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 .
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 , with . In local coordinates, if
then
Here, 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.
B. Hamiltonian mechanics
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 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
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
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
Let us recall that the Poisson bracket is the bilinear operation,
with Xf, Xg being the Hamiltonian vector fields associated with f and g, respectively.
C. Lagrangian systems with external forces
The Poincaré–Cartan 1-form on TQ associated with the Lagrangian function is
where S* is the adjoint operator of the vertical endomorphism on TQ, which is locally
The Poincaré–Cartan 2-form is ωL = −dθL, so locally
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
that is, the integral curves of the forced Euler–Lagrange vector fieldξL,α satisfy the forced Euler–Lagrange equations,
We can write ξL,α = ξL + ξα, where
This vector field is a second order differential equation (SODE), that is,
D. Rayleigh forces
An external force is said to be of Rayleigh type (or simply Rayleigh for short)39,40 if there exists a function on TQ such that
which can be locally written as
This function is called the Rayleigh dissipation function (or the Rayleigh potential). In other words, the fibered morphism associated with is given by the fiber derivative of (see Ref. 46), namely,
A Rayleigh system is a forced Lagrangian system with Lagrangian function L and with external force generated by the Rayleigh potential . For a Rayleigh system with Rayleigh force , the forced Euler–Lagrange vector field is denoted by , given by
This vector field can be written as , where
and the forced Euler–Lagrange equations (12) can be written as
If a Rayleigh potential on TQ defines a Rayleigh force on TQ, also defines for any function f on Q. In other words, given a Rayleigh force , its associated Rayleigh potential 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 force is a Rayleigh force for which 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 tensorR ∈ T*Q × T*Q, given by
A linear Rayleigh system (L, R) is a Rayleigh system such that is a linear Rayleigh force with Rayleigh tensor R.
The Legendre transformation is a mapping Leg: TQ → T*Q such that the diagram
commutes. Here, τq and πQ are the canonical projections on Q. Locally,
with . In what follows, let us assume that the Lagrangian L is hyper-regular, i.e., that Leg is a (global) diffeomorphism.
E. Dissipative bracket
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 functions f, g on TQ.
Since the dissipative bracket is bilinear and verifies the Leibniz rule, it is a derivation or a so-called Leibniz bracket.49
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.
F. Natural Lagrangians and Hamiltonian Rayleigh forces
Consider a natural Lagrangian L 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 . 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 , where L is natural. The associated Hamiltonian Rayleigh potential on T*Q is given by
Similarly, the Hamiltonian Rayleigh force on T*Q is given by
When the Lagrangian is regular, the Legendre transformation is well-defined, so we can define a tensor field given by
In particular, if the Lagrangian is natural, then we have
Hence, the Hamiltonian Rayleigh force can be expressed in terms of the Rayleigh potential as
We shall omit the adjective Hamiltonian and refer to and as the Rayleigh potential and the Rayleigh force, respectively, if there is no danger of confusion.
Let us introduce the vertical differentiation47 on T*Q as
where is defined analogously to ιS by replacing S with . 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 on T*Q is given by
where
Similarly, the Hamiltonian Rayleigh force on T*Q is given by
where . The associated Hamiltonian Rayleigh tensor is given by
The linear Hamiltonian Rayleigh force 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:
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.
III. HAMILTON–JACOBI THEORY FOR SYSTEMS WITH EXTERNAL 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*Q → Q), it is possible to project XH,β along γ(Q), obtaining the vector field
on Q, so that the following diagram commutes:
The vector fields XH,β and are γ-related if and only if XH,β is tangent to γ(Q).
Therefore, the integral curves of 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 , then
so γ◦σ is an integral curve of XH,β. Conversely, if γ◦σ is an integral curve of XH,β for every integral curve σ of Y = TπQ◦XH,β◦γ, 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,
Let us recall that a Lagrangian submanifold 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.
A 1-form γ on Q is called a solution of the Hamilton–Jacobi problem for (H, β) if
it is closed and
it satisfies Eq. (22).
This equation is known as the Hamilton–Jacobi equation.
The results above can be summarized in the following theorem:
Let γ be a closed 1-form on Q. Then, the following conditions are equivalent:
γ is a solution of the Hamilton–Jacobi problem for (H, β).
- For every curve such thatfor all t, then γ◦σ is an integral curve of XH,β.
Im γ is a Lagrangian submanifold of T*Q and XH,β is tangent to it.
Let be a Hamiltonian Rayleigh tensor on T*Q, and let be the associated Hamiltonian Rayleigh force on T*Q. If is non-degenerate, then is a symplectic form on T*Q.
When does not depend on (qi), we can make the change of bundle coordinates,
so that and are Darboux coordinates.
Consider a linear Rayleigh system . Suppose that is non-degenerate. Then, a closed 1-form γ on Q is a solution of the Hamilton–Jacobi problem for if and only if Im γ is a Lagrangian submanifold of .
For a Rayleigh system , the Hamilton–Jacobi equation can also be written as
We will also refer to the Hamilton–Jacobi problem for as the Hamilton–Jacobi problem for . In the case of a linear Rayleigh system , we have
so the Hamilton–Jacobi equation (23) can be written as
We will also refer to the Hamilton–Jacobi problem for as the Hamilton–Jacobi problem for .
One can consider a more general problem by relaxing the hypothesis of γ being closed.
A weak solution of the Hamilton–Jacobi problem for (H, β) is a 1-form γ on Q such that XH,β and are γ-related. Here, is the vector field defined by (20).
Consider a 1-form γ on Q. Then, the following statements are equivalent:
γ is a weak solution of the Hamilton–Jacobi problem for (H, β).
- γ satisfies the equation(25)
XH,β is tangent to the submanifold Im γ ⊂ T*Q.
- If satisfiesthen γ◦σ is an integral curve of XH,β.
Let
Observe that Eq. (25) holds if and only if
(local expressions). Let γ be a 1-form γ on Q. Let H be a Hamiltonian function on T*Q, and let , and 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 if and only if
- γ is a solution of the Hamilton–Jacobi problem for if and only ifMoreover, if γ is not necessarily closed, it is a weak solution of the Hamilton–Jacobi problem for (H, β) if and only ifSimilar expressions can be easily found for Rayleigh forces or linear Rayleigh forces.
A. Complete solutions
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.
Let be an open set. A map Φ: Q × U → T*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 mapis 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 .
The functions fa are constants of the motion. Moreover, they are in involution, i.e., , where is the Poisson bracket defined by ωQ.
IV. HAMILTON–JACOBI THEORY FOR LAGRANGIAN SYSTEMS WITH EXTERNAL FORCES
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:
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 Hamilton–Jacobi 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,
Let X be a vector field on Q that satisfies X*ωL = 0. Then, the following assertions are equivalent:
X is a solution of the Hamilton–Jacobi problem for (L, α).
Im X is a Lagrangian submanifold of TQ invariant by ξL,α.
For every curve such that σ is an integral curve of X, then is an integral curve of ξL,α.
As in the Hamiltonian case, one can consider a more general problem by relaxing the hypothesis of Leg◦X being closed.
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.
Let X be a vector field on Q. Then, the following statements are equivalent:
X is a solution of the generalized Hamilton–Jacobi problem for (L, α).
- X satisfies the equation(27)
The submanifold Im X ⊂ TQ is invariant by ξL,α.
If is an integral curve of X, then X◦σ is an integral curve of ξL,α.
The proof of the converse is completely analogous to the one of Theorem 1 in Ref. 7.□
A. Equivalence between Lagrangian and Hamiltonian Hamilton–Jacobi problems
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.,
Consider a hyper-regular forced Lagrangian system (L, α) on TQ, with the associated forced Hamiltonian system (H, β) on T*Q. Then, X is a (weak) solution of the Hamilton–Jacobi problem for (L, α) if and only if γ = Leg◦X is a (weak) solution of the Hamilton–Jacobi problem for (H, β).
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.
B. Complete solutions
Complete solutions for the Hamilton–Jacobi problem are defined analogously to the ones in T*Q (see Definition 7).
Let be an open set. A map Φ: Q × U → TQ is called a complete solution of the Hamilton–Jacobi problem for (L, α) if
Φ is a local diffeomorphism and
- for any λ = (λ1, …, λn) ∈ U, the mapis a solution of the Hamilton–Jacobi problem for (L, α).
V. REDUCTION AND RECONSTRUCTION OF THE HAMILTON–JACOBI PROBLEM
Let G be a connected Lie group acting freely and properly on Q by a left action Φ, namely,
As usual, we denote by the Lie algebra of G and denote the dual of by . For each g ∈ G, we can define a diffeomorphism
Under these conditions, the quotient space Q/G is a differentiable manifold and πG: Q → Q/G is a G-principal bundle. The action Φ induces a lifted action on T*Q given by
for every . Since Φ is a diffeomorphism, its lift to T*Q leaves θQ invariant;2 in other words, θQ is G-invariant.
The natural momentum map is given by
for each . Here, ξQ is the infinitesimal generator of the action of on Q, and 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 , we have a function 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
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:
Consider a forced Hamiltonian system (H, β). Suppose that H is G-invariant. Let . Then, the following statements hold:
- Jξ is a constant of the motion if and only if
- If the previous equation holds, then ξ leaves β invariant if and only ifMoreover, the vector subspaceis a Lie subalgebra of .(28)
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 submanifold is a Lagrangian submanifold in T*Q such that for all g ∈ G. Given the momentum map J defined above and a Lagrangian submanifold , it can be shown that J is constant along if and only if 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*Q → T*Q/G is a Poisson morphism.
Denote by Gβ ⊂ G the Lie subgroup whose Lie algebra is , defined by (28). Let . Let us assume that . Then, J−1(μ)/G is a symplectic leaf of T*Q/G. Moreover, J−1(μ) is a coisotropic submanifold. Assume that is a Lagrangian submanifold. Then, by the coisotropic reduction theorem,2 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 . 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, can be seen as a Lagrangian submanifold of a cotangent bundle with a modified symplectic structure. Let denote the adjoint bundle to πG: Q → Q/G via the coadjoint representation
A connection A on πG: Q → Q/G induces a splitting
This identification is given by
where denotes the horizontal lift of the connection A. Here, denotes the horizontal space of A at q ∈ Q, namely,
If αq ∈ J−1(μ), then J(αq) = μ and Ψ([αq]) = (αq◦horq, μ), so
Consider a G-invariant forced Hamiltonian system (H, β) on T*Q. Then, H = HG◦π, where 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 on T*(Q/G) by
for each , where and . Similarly, let and
for each .
Let γ be a G-invariant solution of the Hamilton–Jacobi problem for (H, β). Then, the following diagram commutes:
(reduction). Let γ be a G-invariant solution of the Hamilton–Jacobi problem for (H, β). Let and . Then, γ induces a mapping such that and is a solution the Hamilton–Jacobi problem for .
is a G-invariant Lagrangian submanifold of T*Q and
, where γ is a solution of the Hamilton–Jacobi problem for (H, β).
VI. ČAPLYGIN SYSTEMS
A nonholonomic mechanical system is given by a Lagrangian function 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., . Then, the nonholonomic equations of motion are
where , 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: Q → N, over a manifold N.
The constraints are provided by the horizontal distribution of an Ehresmann connection Γ in ρ.
The Lagrangian is Γ-invariant.
A particular case is when ρ: Q → N = 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 , where is the horizontal distribution and Vρ = ker Tρ is the vertical distribution. Take fibered coordinates such that .
With a slight abuse of notation, let denote the horizontal projector hereinafter and let 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 , 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 Y ∈ TyN, y = ρ(q1) = ρ(q2), where denotes the horizontal lift of Y to Q. We can then introduce a function ℓ on TN such that
so locally we have
Now let ρ(q) = y and with τQ(x) = q; let u ∈ TyN, U ∈ Tu(TN) and X ∈ Tx(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 ℓ is a regular Lagrangian and that the Čaplygin system is equivalent to the forced Lagrangian system (ℓ, α).
Assume that ℓ 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 in the algebra Λ*(Q), namely, if a k-form , then
where βi ∈ Λk−1(q) and 1 ≤ i ≤ m.
[Ref. 9 (Theorem 4.3)]. Let denote the horizontal distribution defined by the connection Γ in ρ: Q → N. Let X be vector field on Q such that and . Then, the following conditions are equivalent:
- For every curve such thatfor all t, then X◦σ is an integral curve of Xnh.
.
A vector field X satisfying these conditions is called a solution of the nonholonomic Hamilton–Jacobi problem for (L, Γ).
[Ref. 9 (Theorem 4.5)]. Assume that a vector field X on Q is a solution for the nonholonomic Hamilton–Jacobi problem for (L, Γ). If X is ρ-projectable to a vector field Y on N and is closed, then Y is a solution of the Lagrangian Hamilton–Jacobi problem for (ℓ, α) and γ is a solution of the Hamilton–Jacobi problem for (h, β).
(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).
VII. CONCLUSIONS AND OUTLOOK
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 . 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 (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.
ACKNOWLEDGMENTS
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.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
DATA AVAILABILITY
Data sharing is not applicable to this article as no new data were created or analyzed in this study.