164
Views
9
CrossRef citations to date
0
Altmetric
Original Articles

The energy-free equations of the 3D inverse problem of dynamics

Pages 545-558 | Received 03 Nov 2004, Accepted 04 May 2005, Published online: 26 Jan 2007

Abstract

Given a two-parametric family of curves f(x,y,z) = c1, g(x,y,z) = c2, we look for potentials V under whose action a particle of unit mass can describe the curves of the family. Using a procedure inspired by Kasner, who studied the geometrical properties of the trajectories, we obtain two partial differential equations satisfied by V. These equations do not contain the total energy, which appeared in Szebehely-type ones, and relate merely the potential and the given family. Therefore they can be used when no information on the energy is given in advance, as well as for the direct problem of dynamics. The case of the general autonomous force fields is also discussed.

1. Introduction

The version of the inverse problem of dynamics considered here is: given a two-parametric family of curves in (---545----1) , find the force field (conservative or not) under whose action a material point of unit mass describes, with suitable initial conditions, the curves of the family.

Both the planar (2D) and the spatial (3D) inverse problems of dynamics for general force fields have been considered by Dainelli Citation[1] in the same article, from which Whittaker (Citation[2], p. 93) included in his book only the 2D version. The 2D problem, this time for conservative systems, has renewed the interest in the inverse problem of dynamics by means of Szebehely's Citation[3] partial differential equation relating the potential to a monoparametric family of curves and to the total energy. Similar results for the spatial problem were to appear almost a decade later. Another basic result for the 2D inverse problem was the energy-free partial differential equation obtained by Bozis Citation[4] from Szebehely's equation, and later derived directly by Anisiu Citation[5]. A similar derivation of the basic equations will be given here for the case of spatial families of curves.

Beside the 2D problem, Dainelli Citation[1] studied the existence of a general force field (not necessarily conservative), which can produce as orbits two-parametric 3D curves of the form (---545----2) or, more generally, (---545--1) Actually he referred to a single curve (c1 = c2 = 0), but working with the differentials of ϕ and ψ, respectively of f and g, his results are valid in fact for two-parametric families of curves.

For conservative systems, the 3D problem was tackled by Érdi Citation[6] for a monoparametric family of orbits, and then for two-parametric families by Váradi and Érdi Citation[7]. They obtained partial differential equations satisfied by the potential V, supposing that the total energy of the particle was known; therefore the energy appears in the equations, as in the case of Szebehely's equation for monoparametric 2D families. Puel Citation[8] used the least action principle of Maupertuis to obtain the equations of the 3D inverse problem of dynamics. Because in this case the potential appeared as a solution of two partial differential equations, the existence of such a potential and its relation with the energy was the subject of further studies, like those of Gonzáles-Gascón et al. Citation[9], Bozis and Nakhla Citation[10] and Shorokhov Citation[11]. Puel obtained in Citation[12] the intrinsic equations of the 3D inverse problem, using the Frenet reference frame and revealing interesting geometrical properties. Partial differential equations satisfied by the force components for general force fields were given in Citation[13]. A review of the basic results in the inverse problem of dynamics, including the 3D ones, was presented in Citation[14] by Bozis.

In what follows we derive in a systematic way, considering all the cases that might occur, the energy-free equations for the conservative 3D fields, as we have done for the 2D problem in Citation[5]. We present the spatial region where real motion is allowed, as well as the equations that relate the total energy E, the potential and the given family. EquationEquations (12) and Equation(14) were written as such and used recently to find homogeneous potentials compatible with homogeneous families of orbits by Bozis and Kotoulas Citation[15]. The inverse problem for 3D families of straight lines (Equationequations (10) and Equation(11)) was considered by Bozis and Kotoulas Citation[16]. The 3D inverse problem for a general force field was solved in 1880 by Dainelli Citation[1]. Bozis Citation[13] obtained Equationequation (46), which relates several derivatives of the components X,Y,Z of the force field. Maybe the recent results in the 2D inverse problem, based on the corresponding energy-free equation, as well as the interest in various 3D potentials, have stimulated the research on similar tools for the 3D inverse problem.

The energy-free equations derived in section 2 are applied to find the potentials, the energy and the regions where the motion is possible for two specific families of curves. In section 3 an application involving families of curves created by the 3D Hénon–Heiles type potential is given.

2. The equations of the 3D inverse problem for conservative fields

We consider the 3D family of curves Equation(1) with f, g of C3-class (continuous with continuous derivatives up to the third order on a domain in (---545----3) ) and such that at least one 2 × 2 minor of the Jacobian matrix (∂(F,G)/∂(x,y,z)) does not vanish. In what follows we shall suppose that locally (---545--2) where the subscripts denote the corresponding partial derivatives. Of course, if any other determinant (containing derivatives with respect to x and y, or to x and z) is different from zero, one can proceed similarly.

The inverse problem under consideration is: find the potential V(x,y,z) of C2-class under whose action, for appropriate initial conditions, a material point of unit mass, whose motion is described by (---545--3) will trace the curves of the family Equation(1) (the dots denote derivatives with respect to the time t). The potential will be defined on a 3D domain whose points belong to one and only one curve from the family Equation(1).

In order to obtain the equations satisfied by V, we shall eliminate the time derivatives of x, y, and z following the procedure used by Kasner Citation[17]. To this aim we differentiate Equationequations (1) with respect to t, and from (---545---1) we get (---545--4) where (---545--5)

The notation Equation(5) was introduced in Citation[16], where it was emphasized that the family Equation(1) leads to a unique pair α, β and, conversely, the pair α, β determines uniquely the family Equation(1).

Differentiating both relations in Equation(4) we get (---545--6) This is denoted by (---545--7) and reminding Equation(4) into account, we obtain from relations Equation(6) (---545--8) Making use of Equation(4) again and the equations of motion Equation(3), we get (---545--9)

Our aim being to obtain differential equations satisfied by V, we have to analyse some special cases.

Case 1 A = B = 0. It is obvious that, in view of relation Equation(8), it also follows that (---545----4) , hence the curvature (---545----5) of each member of the family Equation(1) vanishes. We denote (---545----6) by (---545----7) , where (---545----8) are unit vectors along the axes Ox, Oy, Oz.

It follows that we have A = B = 0 if and only if the family Equation(1) consists of straight lines. This case was analysed in detail by Bozis and Kotoulas Citation[16]. Relations Equation(9) give rise to two linear partial differential equations to be necessarily satisfied by V, namely (---545--10) These equations will admit a solution only if α and β satisfy, besides the two equations obtained from Equation(7) for A = B = 0, a supplementary equation (---545--11) So, generally, the inverse problem is not expected to have a solution for arbitrary families of straight lines.

Case 2 A ≠ 0 and B ≠ 0. By eliminating (---545----9) between the two relations in Equation(9) we obtain first a necessary condition to be satisfied by V, (---545--12) where α,β from Equation(5) and A, B from Equation(7) depend on the derivatives of f and g up to the second order. Because of (---545----10) , it follows that the motion is possible only in the region determined by (---545--13)

Differentiating both members of the equality (---545----11) with respect to t and replacing (---545----12) from the first equation in Equation(3), respectively (---545----13) and (---545----14) from Equation(4), we obtain a second differential relation to be satisfied by V (---545--14) where (---545--15)

Summarizing the above reasoning, we assert that a potential that produces as orbits the curves of the family Equation(1) satisfies by necessity the two differential relations Equation(12) and Equation(14), the motion of the particle being possible in the region determined by inequality Equation(13). We remark that Equationequation (14) is of second order in V and does not involve the energy (constant on each curve of the family), hence it is the corresponding equation for the 3D case of Bozis' equation Citation[4] satisfied by planar potentials. The planar equation was derived directly by an elimination process of the derivatives of x(t) and y(t) between some differential relations based on the equations of motion and on the given planar family by Anisiu Citation[5].

In the following we shall derive the equation from which the total energy can be expressed. Denoting by (---545--16) one can check by direct calculation that Equation(14) is equivalent to (---545--17) The characteristic system for Equation(17) is (---545---2) and one easily obtains (---545----15) and (---545----16) . It follows that f(x,y,z) = c1 and g(x,y,z) = c2 are integrals, hence the general solution of Equation(17) is W = E(f,g) with E an arbitrary function.

In view of relations Equation(4) and Equation(9), we get from Equation(16) that (---545--18) i.e. W = E(f, g) is the total energy, constant on each curve of the family Equation(1). It follows that the equation (---545--19) which was derived by Váradi and Érdi Citation[7] using the energy integral (and which corresponds to Szebehely's planar equation), can be obtained as a consequence of the second order partial differential Equationequation (14).

The two Equationequations (12) and Equation(14) for a single unknown function V will not always have a solution; the compatibility conditions are to be checked. The advantage of this formulation consists in the fact that it is free of energy, and one can search for potentials in several classes, which are of astronomical or physical interest, as for example homogeneous (Bozis and Kotoulas Citation[15]) or quasihomogeneous ones.

Remark The fact that Equationequations (12) and Equation(14) do not contain the energy makes them suitable for the direct problem of dynamics: given a 3D potential, find families of orbits of the form Equation(1) generated by it. We can rearrange the mentioned equations and obtain a linear partial differential equation of first order in α and β (---545--20) and a nonlinear one of second order (---545--21) Case 3 B = 0 and A ≠ 0. From the first relation in Equation(9) we obtain again inequality Equation(13); instead of Equation(12), the relation (---545--22) holds, beside the second order partial differential Equationequation (14), which can be derived as in Case 2. The energy is given by Equation(19).

Case 4 A = 0 and B ≠ 0. The inequality to be satisfied is (---545--23) and Equation(12) is replaced by (---545--24) Starting with (---545----17) , we follow the steps from the case when both A and B were different from zero and obtain instead of Equation(14) (---545--25) where (---545--26) The energy is given in this case by (---545---3) Synthesis Given a two-parametric family of curves Equation(1), we compute α and β from Equation(5), and A and B from Equation(7). We obtain the partial differential equations satisfied by the potential V, which are:

Case 1 (A = B = 0): Equation(10);

Case 2 (AB ≠ 0): Equation(12) and Equation(14);

Case 3 (A ≠ 0, B = 0): Equation(22) and Equation(14);

Case 4 (A = 0, B ≠ 0): Equation(24) and Equation(25).

Example 2.1 We consider the two-parametric family of curves Equation(1) with (---545--27) For this family we have (---545---4) The first-order Equationequation (12) in Case 2 reads (---545---5) and has the general solution (---545--28) The second-order Equationequation (14) is (---545--29) Substituting V from Equation(28) in Equation(29), and denoting by (---545---6) we obtain (---545--30) It follows that we must have (---545---7) which gives v(u,w) = F1(u + 2w) + uF2(u + 2w), with F1,F2 arbitrary functions. Introducing this value of v(u, w) in the right-hand side of Equation(30) and equating it with zero leads to (---545----18) where C is an arbitrary constant. We finally get (---545---8) From Equation(19) we obtain the energy E = 4C; the inequality Equation(13) becomes (---545----19) .

We remark that for (---545----20) and C = 0 we have (---545--31) and E = 0; condition Equation(13) gives no restriction for the domain where the particle moves. In three curves of the family Equation(27) are plotted traced under the action of the potential Equation(31). The curves pass respectively through the points (1,1,1), (1.4,1,1) and (1.6,0.8,1), and the initial velocities for the system Equation(3) are obtained from Equation(9) and Equation(4).

Figure 1. Curves of the family Equation(27) produced by the potential Equation(31).

Figure 1. Curves of the family Equation(27)(---545--27) produced by the potential Equation(31)(---545--31) .

Example 2.2 For the two-parametric family of elliptic orbits given by (---545--32) with a priori energy given (---545----21) , Váradi and Érdi Citation[7] and Bozis and Nakhla Citation[10], have found the central potential (---545----22) , which gives rise to the curves of the family Equation(32). For this family we have (---545---9) hence it enters Case 3. Relation Equation(22) becomes (---545---10) with the general solution (---545--33) We substitute V from Equation(33) in Equation(14) and get (---545--34) where w = x2 + z2, (---545----23) .

It follows that d11 = 0; its general solution is (---545----24) . By substituting this in Equation(34) we obtain (---545--35) This equality is possible if and only if each member is equal to a constant k0. It follows that (---545---11) or (---545--36) From Equation(19) we obtain the energy (---545---12) Real motion is possible only in the region Equation(13), which is described in this case by (---545---13) For the particular case k0 = 4, C1 = C2 = 1/2, the potential (---545--37) will give rise to the curves of the family Equation(32). In the curves of that family passing through the points (1.1,1,1), (1.05,1.05,1) and (1,1,1.1) are shown. They are traced by a unit mass particle with initial velocities calculated from Equation(9) and Equation(4).

Figure 2. Curves of the family Equation(32) produced by the potential Equation(37).

Figure 2. Curves of the family Equation(32)(---545--32) produced by the potential Equation(37)(---545--37) .

It can be checked that the potential and the prescribed energy given by Váradi and Érdi Citation[7] follow from our result for k0 = 4, C1 = C2 = 0. Using the energy-free equations we have obtained a larger class of potentials, which can give rise to the elliptic orbits Equation(32).

3. Application to cubic potentials

The family of Hénon–Heiles 3D potentials (---545--38) with at least one of a, b, c different from zero, consists of cubic perturbations of harmonic oscillators. These potentials have been intensively studied lately, due to their applicability in various fields of physics; e.g., they describe the motion of the nuclei in small molecules Citation[18], as well as the motion in the central region of a galaxy Citation[19].

Considering the 2D case, Anisiu and Pal Citation[20] showed that, out of the families of curves (---545----25) , only two are compatible with potentials of Hénon–Heiles type, namely:

1.

the family (---545----26) with (---545----27) ;

2.

the family (---545----28) with (---545----29) .

We shall apply the results in section 2 to study the following problem: are there two-parametric families of curves Equation(1) with (---545--39) compatible with a cubic potential of the form Equation(38)?

For the family Equation(39) we have (---545---14)

First we consider the case r = s ≠ 1, for which A = 0 and B ≠ 0. The first-order equation satisfied by V is in this case Equation(24); it holds if and only if (---545----30) and b = c. The second order Equationequation (25) is satisfied for:

1.

(---545----31) ; this means that the family (---545----32) is compatible with the potential

(---545--40)

2.

(---545----33) ; in this case the family (---545----34) is compatible with

(---545--41)

If r ≠ s,r ≠ 1 we have A ≠ 0,B ≠ 0, and from Equationequation (12) it follows a = b = c = 0, hence no cubic potential is compatible with a family of the form Equation(39). The same situation arises when Equationequation (10) are used for the family of straight lines obtained for r = s = 1, as well as for 1 = r ≠ s (when B = 0 and A ≠ 0, and we use Equationequation (22) from Case 3 in section 2).

We remark that both cubic potentials Equation(40) and Equation(41) that are compatible with families of the form Equation(39) consist of an anisotropic harmonic oscillator, the first one having an axially symmetric perturbation and the second one a 1D cubic perturbation.

4. The case of a general force field

Let us consider a system of the form (---545--42) the force components X , Y and Z being of C1-class on a domain in (---545----35) . The differential equations of the trajectories of Equation(42), and the geometrical properties of these trajectories, were obtained by Kasner Citation[17].

The reasoning in section 2 can be applied for the system Equation(42) too; we remind the notations α, β in Equation(5), and A, B in Equation(7), which are related to the given family of curves Equation(1).

If the family Equation(1) consists of straight lines, i.e. A = B = 0, instead of Equation(10) we have (---545--43) These are the relations satisfied in this special case by the components X, Y, Z of the force. It follows that X can be an arbitrary function, and then Y = αX, Z = βX.

Let us consider now A ≠ 0, B ≠ 0. A first relation is obtained from (---545----36) , namely (---545--44) The positivity of (---545----37) implies that the region where the motion is allowed is (---545--45) By differentiating (---545----38) with respect to t and by using Equation(42) and Equation(4) we get (---545--46) where p, q, l, m are given by Equation(15). This partial differential equation, as well as the algebraic relation Equation(44), has been obtained, with different notation in Citation[13]. We remark that, if we denote by (---545--47) Equationequation (46) can be written as (---545---15) or (---545--48) From relation Equation(47) we then get (---545--49) and from Equation(44) (---545--50) Therefore in the general case the inverse problem always has solutions; for an arbitrary positive function ξ, we obtain the components of the force, given by Equation(48)Equation(50), which can produce as orbits the curves of the two-parametric family Equation(1).

The situation when only one of A and B is identically null can be treated similar to that in section 2 for conservative fields.

To end this journey through the 3D inverse problem we remind the formulae for the force components as given in Citation[1], namely (---545--51) where (---545----39) is an arbitrary function of the three variables x, y, z, and (---545----40) . We put a tilde on Dainelli's notation in order to avoid confusion with ours. The expressions in Equation(51) can be obtained from Equation(48) to Equation(50) by replacing (---545----41) with δ from Equation(2); it will follow that (---545----42) and (---545----43) , with (---545----44) from Equation(5).

5. Conclusions

Generally, the 3D inverse problem of dynamics gives rise to two energy-free partial differential equations (of first and second order) satisfied by the potential, and to an inequality, which determines the region in space where real motion is possible. These are deduced in a natural way following the ideas of Kasner Citation[17]. The relation between conservative and nonconservative case is also clarified.

The examples in section 2 use the energy-free equations to obtain all the potentials, which give rise to a given two-parametric 3D family of curves. In section 3 we provide two families of curves compatible with 3D cubic potentials (perturbations of anisotropic oscillators).

Families of more complicated orbits can be described in parametric form; this fact justifies further studies in order to extend the 2D results of Bozis and Borghero Citation[21] for conservative systems, respectively of Anisiu and Pal Citation[22] for general force fields.

Acknowledgement

The author thanks the referees for their valuable suggestions.

Additional information

Notes on contributors

Mira-Cristiana Anisiu *

Notes

  • Dainelli, U, 1880. Sul movimento per una linea qualunque, Giornale di Matematiche 18 (1880), pp. 271–300.
  • Whittaker, ET, 1904. Analytical Dynamics of Particles and Rigid Bodies. 1904, C. J. Clay and Sons.
  • Szebehely, V, 1974. "On the determination of the potential by satellite observations". In: Proc. Int. Meeting on Earth's Rotation by Satellite Observation. Bologna, Italy. 1974. p. pp. 31–35, In: G. Proverbio (Ed.).
  • Bozis, G, 1984. Szebehely inverse problem for finite symmetrical material concentrations, Astronomy and Astrophysics 134 (1984), pp. 360–364.
  • Anisiu, M-C, 2004. An alternative point of view on the equations of the inverse problem of dynamics, Inverse Problems 20 (2004), pp. 1865–1872.
  • Érdi, B, 1982. A generalization of Szebehely's equation for three dimensions, Celestial Mechanics 28 (1982), pp. 209–218.
  • Váradi, F, and Érdi, B, 1983. Existence and solution of Szebehely's equation in three dimensions using a two-parametric family of orbits, Celestial Mechanics 30 (1983), pp. 395–405.
  • Puel, F, 1984. Équations de Szebehely et principes variationnels, Celestial Mechanics 32 (1984), pp. 349–353.
  • Gonzáles-Gascón, FJ, Gonzáles-López, A, and Pascual-Broncano, PJ, 1984. On Szebehely's equation and its connection with Dainelli's-Whittaker's equations, Celestial Mechanics 33 (1984), pp. 85–97.
  • Bozis, G, and Nakhla, A, 1986. Solutions of the three-dimensional inverse problem, Celestial Mechanics and Dynamical Astronomy 38 (1986), pp. 357–375.
  • Shorokhov, SG, 1988. Solution of an inverse problem of the Dynamics of a particle, Celestial Mechanics and Dynamical Astronomy 44 (1988), pp. 193–206.
  • Puel, F, 1992. Explicit solutions of the three-dimensional inverse problem of Dynamics, using the Frenet reference frame, Celestial Mechanics and Dynamical Astronomy 53 (1992), pp. 207–218.
  • Bozis, G, 1983. Determination of autonomous three-dimensional force fields from a two-parameter family of orbits, Celestial Mechanics 31 (1983), pp. 43–51.
  • Bozis, G, 1995. The inverse problem of dynamics: basic facts, Inverse Problems 11 (1995), pp. 687–708.
  • Bozis, G, and Kotoulas, TA, 2005. Homogeneous two-parametric families of orbits in three-dimensional homogeneous potentials, Inverse Problems 21 (2005), pp. 343–356.
  • Bozis, G, and Kotoulas, TA, 2004. Families of straight lines (FSL) in three-dimensional potentials, Rendiconti del Seminario della Facoltà di Scienze dell'Università di Cagliari 75 (2004), (to appear).
  • Kasner, E, 1907. Dynamical trajectories: the motion of a particle in an arbitrary field of force, Transactions of the American Mathematical Society 8 (1907), pp. 135–158.
  • Gutzwiller, MC, 1990. "Chaos in classical and quantum mechanics". In: Interdisciplinary Applied Mathematics 1. New York. 1990.
  • Ferrer, S, Hanssmann, H, Palacián, J, and Yanguas, P, 2002. On perturbed oscillators in 1-1-1 resonance: the case of axially symmetric cubic potentials, Journal of Geometry and Physics 40 (2002), pp. 320–369.
  • Anisiu, M-C, and Pal, A, 1999. Special families of orbits for the Hénon-Heiles type potential, Romanian Astronomical Journal 9 (1999), pp. 179–185.
  • Bozis, G, and Borghero, F, 1998. A new formulation of the two-dimensional inverse problem of dynamics, Inverse Problems 14 (1998), pp. 41–51.
  • Anisiu, M-C, and Pal, A, 1999. Two-dimensional inverse problem of dynamics for families in parametric form, Inverse Problems 15 (1999), pp. 135–140.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.