1. Introduction
Polyharmonic curves of order
k in Riemannian manifolds are the critical points of the
k-energy functional
and are described by the Euler–Lagrange equation
Notice that the functional (
1) is considered a higher-order version of the energy functional
and, in this sense,
k-harmonic curves, also referred to as polyharmonic curves, higher-order geodesics, or Riemannian polynomials, are seen as a natural generalization of geodesic curves, the extremal curves of the functional (
3).
The study of polyharmonic curves fits into the more general theory of polyharmonic maps between Riemannian manifolds, just as the theory of geodesics falls under that of harmonic applications. Polyharmonic maps have only recently became a subject of interest (see [
1] and references therein), but biharmonic maps and, in particular, biharmonic submanifolds and curves have been extensively studied in the last decades (see, for instance, [
2,
3,
4,
5,
6]).
There is a strong relationship between optimal control problems and variational problems, particularly concerning the variational problem associated with the
k-energy functional (
1). The main topic of this subject is the study of the dynamic interpolation problem, where the goal is to find the curves that minimize the 2-energy functional and satisfy some interpolation conditions. Applications to motion planning and tracking problems for nonlinear systems were the special motivation for the analysis of this second-order problem. The first steps in this direction were given by L. Noakes, G. Heinzinger, and B. Paden in [
7] and by P. Crouch and F. Silva Leite in [
8], where the authors obtained the necessary optimality conditions for the problem and called Riemannian cubic splines to the curves under these conditions. In the context of robotic motion planning, a natural extension of the dynamic interpolation problem to higher orders has also been developed, giving rise to the notion of higher-order splines in Riemannian manifolds [
9,
10].
Applications of polyharmonic curves to trajectory planning problems in robotics and computational anatomy, especially when the configuration space is a Lie group, also brought the subject to the field of geometric mechanics. The Hamiltonian structure and symmetry reductions of the polyharmonic equation have deserved special attention and have also been extended to the study of optimal control problems for mechanical control systems [
11,
12,
13,
14,
15,
16,
17].
Polyharmonic curves depend on the choice of the parametrization, as happens with geodesics. From the point of view of differential geometry, these curves are studied by considering arclength parametrization. When they are seen as motion trajectories, arclength parametrization is not always possible (when motion reaches zero velocity) and is thought of as a constraint.
In this work, we present an intrinsic version of the
k-harmonic equation based on the symplectic formalism for higher-order regular Lagrangians given in [
18]. More specifically, we consider a geometric formulation of the
kth-order variational problem on a Riemannian manifold using the framework of symplectic geomety and define a generalized Legendre transformation involving higher-order tangent and cotangent bundles. The corresponding Hamiltonian equation obtained via this Legendre transformation is also explained. This study covers some research topics of interest, such as the interpolation theories involving geodesics and cubic splines. In fact, these cases are explored in the present work as being free harmonic and biharmonic curves (without any constraints on the parameter). The relationship of the variational problem with the optimal control problem is also an interesting field of research and is presented for the case of biharmonic curves, always with emphasis on the intrinsic approach.
The structure of the paper is as follows. In
Section 2, we recall some important notions from the geometry of higher-order tangent bundles. The variational problem associated with the
k-harmonic curves is studied in
Section 3. We begin by showing that the
kth-order Lagrangian is regular and then adapt the Lagrangian formalism of higher order to the problem being studied. A higher-order Legendre transformation that allows relating the Lagrangian and the Hamiltonian formalisms is described.
Section 4 is devoted to the first-order case, which corresponds to the classical geodesic problem. In
Section 5, the formalism for biharmonic curves is explored in more depth and, in this case, the associated optimal control problem is also exposed.
2. Higher-Order Tangent Bundles
Let M be a differentiable manifold of finite dimension n. Consider a local coordinate system on M, simply denoted by . Throughout this paper, we use similar abbreviations for the coordinate notations. Let k be an integer greater than or equal to 1.
In this work, we are interested in the formalism of higher-order tangent bundles. In order to introduce the geometry of those bundles (see [
18] for further details), we consider the well-defined equivalence relationship on the set of smooth curves in
M, as follows:
We say that two smooth curves in M, and , defined on an interval with , have a contact of order k at 0 if , and for a local coordinate system on M around x, the derivatives of and up to order k, included, coincide at 0.
The equivalence class determined by a curve is represented by and is called k-jet or k-velocity.
Definition 1. The tangent bundle of order k of M is the set of all equivalence classes of curves in M that have contact of order k and is denoted by .
The following characteristics of the tangent bundle should be emphasized:
is a -dimensional manifold and a fibered manifold over M with projection .
has natural local coordinates
induced by
, where
for
and
If , is identified with the manifold M and for , is just the tangent bundle of M, .
There are canonical projections
, which define several different fibered structures on
. Locally,
Note that
. The tangent applications
are defined by
for each
, with
.
Definition 2. Let γ be a smooth curve in M. The lift to of γ is a smooth curve in denoted by and defined by , where .
If is locally given by , then locally represents .
2.1. The Liouville Vector Field of Higher Order
In order to introduce the notion of a Liouville vector field of higher order, we begin by defining
k vertical bundles of
determined by foliations of type (
4) of
. Let
.
Definition 3. The vertical bundle of over , denoted by , is the set of all tangent vectors to that are projected onto zero by .
According to (
4) and (
5), if
and
X is an element of
at
, then
X is locally written as
Remark 1. In the particular case when and , the projection is just the canonical projection of the tangent bundle , . The only vertical bundle is over M, usually denoted by , whose elements are tangent vectors of and which are projected onto zero for .
Now consider:
The canonical applications
The vector bundle isomorphisms over
locally defined by
where
is the induced bundle of
via
.
Remark 2. If , we have just one vectorial bundle isomorphism over ,which is locally given by .
If , we can define two vectorial bundle isomorphisms over
,
which is locally defined by Definition 4. The canonical vector field of order r on is the vector field defined by the composition where is the identity map in .
The Liouville vector field of order k is the canonical vector field of order 1 on ,
.
For , .
Remark 3. If , then is the identity map, and we have just the Liouville vector field on , : If , we have two canonical vector fields on , the Liouville vector fields and , which are locally given by 2.2. The Canonical Almost-Tangent Structure of Higher Order
We now generalize to higher order the notion of canonical almost-tangent structures. For , consider the following:
Definition 5. The endomorphism defined byis called the vertical endomorphism of order r of ,
.
The vertical endomorphism is called a canonical almost-tangent structure of order k on .
Proposition 1. The vertical endomorphism of order r of has a constant rank equal to and satisfies According to the above proposition, is an almost-tangent structure on since and .
Remark 4. If ,
we have two vertical endomorphisms of ,
and ,
whose matrix representations are, respectively, given bywhere and 0 are the identity matrix and the null matrix of order n, respectively. In this case,
and ,
so determines an almost-tangent structure of order 2 on ,
the so-called canonical almost-tangent structure of order 2 on .
Notice that .
Proposition 2. Let be the vertical endomorphism of order i of and let be the canonical vector field of order i on (i = r, s). The following relationships are satisfied:with .
Notice that, on , we have .
Definition 6. The vertical differentiation of order r on the exterior algebra of , denoted by , is given by the commutatorwhere d is the exterior differentiation and is the inner product of .
Proposition 3. The vertical differentiation of order r on the exterior algebra on satisfies, for each function f on ,
the following relationship: Remark 5. Notice that on , we have Moreover, on ,
we obtainwhere f is a function on .
2.3. The Tulczyjew Differential Operator
Definition 7. The Tulczyjew differential operator or total time derivative operator on is the operator that maps each function f on to a function on such thatfor each , with
defined in (
6).
In local coordinates, we obtain
We shall mention that the total time derivative may be naturally extended to an operator that acts on differentiable forms. This operator maps p-forms on into p-forms on . Moreover, we have , where d is the exterior differentiation defined on the exterior algebra on .
Definition 8. Let X be a vector field along a curve γ in M. The kth-order lift of X is a vector field along the lifted curve , , satisfying .
Note that
is obtained by applying repeated lifts to
X. Its local coordinate expression is given by
where
.
3. Higher-Order Variational Problem
From now on, we take
M to be a Riemannian manifold with the Riemannian metric
. The Levi–Civita connection on
M is denoted by ∇. Let
represent the covariant derivative
along the curve
in
M, with
X being a vector field along
. Set
as the
jth-order covariant derivative of
, where
and
. Consider the following sign convention for the curvature tensor field
R:
See [
19] for more details about Riemannian geometry.
Remember that if a curve
in
M is locally represented by
, then
is locally represented by
. Thus, the velocity vector field along the curve
is
. Moreover, given a vector field
, the covariant derivative of
X along
is given by
In particular, the covariant acceleration of
can be written as
where we simplify the notations of the derivatives, using
for the first derivative
and similar notations for the higher-order derivatives. Here,
are the Christoffel symbols defining the Riemannian connection, which can be obtained using the identity
where
are the components of the Riemannian metric and
is the inverse matrix of the matrix
.
Using the Riemannian structure of
M, we can also define the bundle morphism
from
to
given by
. The morphism
can be expressed as follows.
where
and
3.1. The k-Energy Functional
Let
be the class of smooth curves
satisfying the boundary conditions
where
,
are fixed
n-vectors (
;
) and
. Consider the
kth-order variational problem described by the action functional
defined by (
1). From the point of view of intrinsic variational calculus,
can be written as
where
L is the Lagrangian of order
k associated with the problem. Therefore, the Lagrangian of the problem,
, is defined, for each
, by
where
is given by (
11). We may remark that (
12) may be locally expressed by
Differentiating
L, we obtain
Furthermore, and, since this is the matrix that represents the Riemannian metric, we have the guarantee that the Lagrangian L is regular.
3.2. Intrinsic Version of the Euler–Lagrange Equation
Given a curve
in
, the tangent space to
at
,
, is constituted by smooth vector fields
X along
such that
for
, where
is the
jth-order lift of
X. The variation of the curve
is given by a smooth 1-parameter family of curves
with
, and the corresponding variation vector field
is defined by
. The first-order variation of
associated with
X takes the form
Hamilton’s variational principle establishes that a curve
is a critical curve of
if, for an arbitrary variation vector
, we have
. If the Lagrangian
L is regular (which is the case that we are considering), the arbitrariness of the variation vector field
X in the condition for the action integral to be stationary,
gives the geometric version of the Euler–Lagrange equation
where
is the Poincaré–Cartan 2-form on
and
is the energy function associated with
, defined, respectively, by
Proposition 4. The one-form on is semibasic of type k; that is, .
One calls
the
Jacobi–Ostrogradsky form associated with the Lagrangian
L. We have
. Locally,
where
,
.
The Euler–Lagrange Equation (
13) uniquely defines the vector field
on
since, due to the regularity of the Lagrangian
L,
is sympletic (see [
18]). Moreover, since
,
is a semispray on
of type 1, which represents the
kth-order differential Equation (
2). This means that the integral curves of
are lifts to
of the curves in
M satisfying the Euler–Lagrange Equation (
2).
We also remark that Equation (
2) can be rewritten in local coordinates as follows:
3.3. Generalized Legendre Transformation and the Hamiltonian Approach
Proposition 4 allows us to conclude that the Jacobi–Ostrogradsky form (
14),
, is semibasic of type
k and consequently determines, via the identity
the Legendre transformation
(see [
18] for more details),
where
represents the pairing duality of vectors and covectors on
, and
are the natural projections. Locally, we have
where
,
, are the real functions defined by (
14).
When
is a diffeomorphism, we say that the Lagrangian
L is
hyper-regular, and we have a symplectomorphism from
to
, where
is the symplectic canonical form on
. Under the hyper-regularity condition, we can consider the Hamiltonian energy function associated with
L given by
and the system
is associated with the Hamiltonian system
. The dynamics of the Hamiltonian system is described by
and the Hamiltonian vector field
defined by (
15) verifies
The Hamiltonian function
is locally given by