Skip to main content
Log in

On the coplanar eccentric non-restricted co-orbital dynamics

  • Original Article
  • Published:
Celestial Mechanics and Dynamical Astronomy Aims and scope Submit manuscript

Abstract

We study the phase space of eccentric coplanar co-orbitals in the non-restricted case. Departing from the quasi-circular case, we describe the evolution of the phase space as the eccentricities increase. We find that over a given value of the eccentricity, around 0.5 for equal mass co-orbitals, important topological changes occur in the phase space. These changes lead to the emergence of new co-orbital configurations and open a continuous path between the previously distinct trojan domains near the \(L_4\) and \(L_5\) eccentric Lagrangian equilibria. These topological changes are shown to be linked with the reconnection of families of quasi-periodic orbits of non-maximal dimension.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16

Similar content being viewed by others

Notes

  1. The planar dynamics is decoupled from the dynamics of the inclinations at first order in the inclination, see Robutel and Pousse (2013).

  2. It is proven in Robutel et al. (2016) that the averaging process is not convergent in a neighbourhood of the collision between the two planets including the Hill sphere associated with this collision. The two Eulerian configurations that correspond to \(L_1\) and \(L_2\) are consequently excluded from the present study.

  3. As long as the Gascheau criterion is fulfilled.

  4. The \({{\mathcal {F}} ^{sc}}\) and \({{\mathcal {F}} ^{{ sf}}}\) correspond, respectively, to the \(\sigma -family\) and the \({\varDelta \varpi } -family\) studied in Giuppone et al. (2010).

  5. The red curves satisfy the relation (44), contain the collision point (0, 0), and are located at a relevant position for the collision manifold (see the purple curves in the unstable areas in Figs. 5, 6, 7, 8, 9, 10, 11, 12, 13, and 14). The collision manifold satisfies Eq. (44) if \(\frac{\partial }{\partial \zeta } {\mathcal {H}} _{{\mathcal {R}} {\mathcal {M}} }\) tends to \(- \infty \) when we get close to the collision from one side and \(+ \infty \) from the other side.

  6. When \(m_1=m_2\), we suppose that the reference manifold represents all the co-orbital configurations reaching \(a_1=a_2\) on their orbit, for a given value of the total angular momentum. However, the relative size of the section of two stability domains by the reference manifold is not necessarily representative of the relative volume of these two configurations in the phase space. For example, Fig. 2 shows that depending on the chosen section, the horseshoe domain may appear larger or smaller than the tadpole one (for \(\mu \approx 10^{-6}\)).

  7. Interestingly, while the position of \(\{\nu =0\}\cap {\mathcal {V}} \) for \({\varDelta \varpi } =0\) depends strongly of the value of \(e_1=e_2\), the position of \(\{g=0\}\cap {\mathcal {V}} \) for \({\varDelta \varpi } =0\) seems to occur around \(\zeta =40^\circ \) for any value of \(e_1=e_2 \gtrsim 0.5\), see Figs. 8, 9, 10 and 14.

  8. Orbits in the close neighbourhood of the unstable family can also verify the condition (47) when integrated over a duration of the order of \(1/{\varepsilon } \) because g tends to 0 for this family.

  9. The large amount of grey pixels in the quasi-satellite domain is due to numerical instabilities: between the blue and green domains on the right hand graph, each eccentricity vanishes periodically, while the other gets close to 0.99, so our integration step of 0.01 orbital periods is not adapted to such high eccentricities.

  10. Note that in this case the integration time is too short to properly account for the effect of the secular dynamics (which is also of the order of \(10^6\) orbital periods).

References

  • Batygin, K., Morbidelli, A.: Analytical treatment of planetary resonances. Astron. Astrophys. 556, A28 (2013)

    Article  ADS  Google Scholar 

  • Beaugé, C., Roig, F.: A semianalytical model for the motion of the trojan asteroids: proper elements and families. Icarus 153, 391–415 (2001)

    Article  ADS  Google Scholar 

  • Charlier, C.V.L.: Über den Planeten 1906 TG. Astron. Nachr. 171, 213 (1906)

    Article  ADS  Google Scholar 

  • Delisle, J.-B., Laskar, J., Correia, A.C.M.: Resonance breaking due to dissipation in planar planetary systems. Astron. Astrophys. 566, A137 (2014)

    Article  Google Scholar 

  • Delisle, J.-B., Laskar, J., Correia, A.C.M., Boué, G.: Dissipation in planar resonant planetary systems. Astron. Astrophys. 546, A71 (2012)

    Article  Google Scholar 

  • Dermott, S.F., Murray, C.D.: The dynamics of tadpole and horseshoe orbits. I - Theory. Icarus 48, 1–11 (1981)

    Article  ADS  Google Scholar 

  • Érdi, B.: An asymptotic solution for the trojan case of the plane elliptic restricted problem of three bodies. Celest. Mech. 15, 367–383 (1977)

    Article  ADS  MATH  Google Scholar 

  • Érdi, B., Nagy, I., Sándor, Z., Süli, Á., Fröhlich, G.: Secondary resonances of co-orbital motions. MNRAS 381, 33–40 (2007)

    Article  ADS  Google Scholar 

  • Ford, E.B., Gaudi, B.S.: Observational constraints on Trojans of transiting extrasolar planets. Astrophys. J. Lett. 652, 137–140 (2006)

    Article  ADS  Google Scholar 

  • Garfinkel, B.: Theory of the Trojan asteroids. I. Astron. J. 82, 368–379 (1977)

    Article  ADS  Google Scholar 

  • Gascheau, G.: Examen d’une classe d’équations différentielles et application à un cas particulier du problème des trois corps. C. R. Acad. Sci. Paris 16(7), 393–394 (1843)

    Google Scholar 

  • Gastineau, M., Laskar, J.: Trip: a computer algebra system dedicated to celestial mechanics and perturbation series. ACM Commun. Comput. Algebra 44(3/4), 194–197 (2011)

    MATH  Google Scholar 

  • Giuppone, C.A., Beaugé, C., Michtchenko, T.A., Ferraz-Mello, S.: Dynamics of two planets in co-orbital motion. MNRAS 407, 390–398 (2010)

    Article  ADS  Google Scholar 

  • Giuppone, C.A., Benitez-Llambay, P., Beaugé, C.: Origin and detectability of co-orbital planets from radial velocity data. MNRAS (2012)

  • Hadjidemetriou, J.D., Psychoyos, D., Voyatzis, G.: The 1/1 resonance in extrasolar planetary systems. Celest. Mech. Dyn. Astron. 104, 23–38 (2009)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Hadjidemetriou, J.D., Voyatzis, G.: The 1/1 resonance in extrasolar systems. Migration from planetary to satellite orbits. Celest. Mech. Dyn. Astron. 111, 179–199 (2011)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Henrard, J., Caranicolas, N.D.: Motion near the 3/1 resonance of the planar elliptic restricted three body problem. Celest. Mech. Dyn. Astron. 47, 99–121 (1989)

    Article  ADS  MathSciNet  Google Scholar 

  • Laskar, J.: The chaotic motion of the solar system: a numerical estimate of the size of the chaotic zone. Icarus 88, 266–291 (1990)

    Article  ADS  Google Scholar 

  • Laskar, J., Robutel, P.: Stability of the planetary three-body problem I: expansion of the planetary hamiltonian. Celest. Mech. Dyn. Astron. 62, 193–217 (1995)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Laskar, J., Robutel, P.: High order symplectic integrators for perturbed Hamiltonian systems. Celest. Mech. Dyn. Astron. 80, 39–62 (2001)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Leleu, A.: Dynamics of co-orbital exoplanets. PhD thesis (2016)

  • Leleu, A., Robutel, P., Correia, A.C.M.: Detectability of quasi-circular co-orbital planets: application to the radial velocity technique. Astron. Astrophys. 581, A128 (2015)

    Article  ADS  Google Scholar 

  • Leleu, A., Robutel, P., Correia, A.C.M., Lillo-Box, J.: Detection of co-orbital planets by combining transit and radial-velocity measurements. Astron. Astrophys. 599, L7 (2017)

    Article  ADS  Google Scholar 

  • Liouville, J.: Sur un cas particulier du problème des trois corps. C. R. Acad. Sci. Paris 14, 503–506 (1842)

    Google Scholar 

  • Meyer, K.R., Hall, G.R.: Introduction to Hamiltonian Dynamical Systems and the n-Body Problem. Springer, Berlin (1992)

    Book  MATH  Google Scholar 

  • Michtchenko, T.A., Ferraz-Mello, S., Beaugé, C.: Modeling the 3-d secular planetary three-body problem. Icarus 181, 555–571 (2006)

    Article  ADS  MATH  Google Scholar 

  • Mikkola, S., Innanen, K., Wiegert, P., Connors, M., Brasser, R.: Stability limits for the quasi-satellite orbit. Mon. Not. R. Astron. Soc 369, 15–24 (2006)

    Article  ADS  Google Scholar 

  • Morais, M.H.M.: A secular theory for Trojan-type motion. Astron. Astrophys. 350, 318–326 (1999)

    ADS  Google Scholar 

  • Morais, M.H.M.: Hamiltonian formulation of the secular theory for Trojan-type motion. Astron. Astrophys. 369, 677–689 (2001)

    Article  ADS  MATH  Google Scholar 

  • Morais, M.H.M., Namouni, F.: Retrograde resonance in the planar three-body problem. Celest. Mech. Dyn. Astron. 117, 405–421 (2013)

    Article  ADS  MathSciNet  Google Scholar 

  • Morbidelli, A.: Modern Celestial Mechanics : Aspects of Solar System Dynamics. Taylor & Francis, London (2002). ISBN 0415279399

    Google Scholar 

  • Namouni, F.: Secular interactions of coorbiting objects. Icarus 137, 293–314 (1999)

    Article  ADS  Google Scholar 

  • Nauenberg, M.: Stability and eccentricity for two planets in a 1:1 resonance, and their possible occurrence in extrasolar planetary systems. Astron. J. 124, 2332–2338 (2002)

    Article  ADS  Google Scholar 

  • Nesvorný, D., Thomas, F., Ferraz-Mello, S., Morbidelli, A.: A perturbative treatment of the co-orbital motion. Celest. Mech. Dyn. Astron. 82(4), 323–361 (2002)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Páez, R.I., Efthymiopoulos, C.: Trojan resonant dynamics, stability, and chaotic diffusion, for parameters relevant to exoplanetary systems. Celest. Mech. Dyn. Astron. 121, 139–170 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Pousse, A., Robutel, P., Vienne, A.: On the co-orbital motion in the planar restricted three-body problem: the quasi-satellite motion revisited. Celest. Mech. Dyn. Astron. (2017)

  • Roberts, G.: Linear stability of the elliptic Lagrangian triangle solutions in the three-body problem. JDIFE 182, 191–218 (2002)

    MathSciNet  MATH  Google Scholar 

  • Robutel, P., Gabern, F.: The resonant structure of Jupiter’s Trojan asteroids I: long-term stability and diffusion. MNRAS 372, 1463–1482 (2006)

    Article  ADS  Google Scholar 

  • Robutel, P., Laskar, J.: Frequency map and global dynamics in the solar system I: short period dynamics of massless particles. Icarus 152, 4–28 (2001)

    Article  ADS  Google Scholar 

  • Robutel, P., Niederman, L., Pousse, A.: Rigorous treatment of the averaging process for co-orbital motions in the planetary problem. Comput. Appl. Math. 35(3), 675–699 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  • Robutel, P., Pousse, A.: On the co-orbital motion of two planets in quasi-circular orbits. Celest. Mech. Dyn. Astron. 117, 17–40 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  • Sidorenko, V., Artemiev, A., Neishtadt, A., Zelenyi, L.: Quasi-satellite orbits in general context of dynamics at 1:1 mean motion resonance: a perturbative treatment (2014)

Download references

Acknowledgements

The authors acknowledge financial support from the Observatoire de Paris Scientific Council, CIDMA strategic project UID/MAT/04106/2013, ENGAGE SKA POCI-01- 0145-FEDER-022217 (funded by COMPETE 2020 and FCT, Portugal), and the Marie Curie Actions of the European Commission (FP7-COFUND). Parts of this work have been carried out within the frame of the National Centre for Competence in Research PlanetS supported by the SNSF. This work was granted access to the HPC resources of MesoPSL financed by the Region Ile de France and the project Equip@Meso (Reference ANR-10-EQPX-29-01) of the programme Investissements dAvenir supervised by the Agence Nationale pour la Recherche. The authors thank the referees for useful suggestions that greatly improved the description of the results.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to A. Leleu.

Appendices

Appendix A: Time scales

The planar 3-body co-orbital problem has 3 time scales: the fast time scale, associated with the mean mean-motion \(\eta ={\mathcal {O}} (1)\), the semi-fast time scale of fundamental frequency \(\nu ={\mathcal {O}} (\sqrt{{\varepsilon } })\), associated with the evolution of the resonant angle \(\zeta \), and the secular time scale of fundamental frequency \(g_1\) and \(g_2\), of order \({\mathcal {O}} ({\varepsilon } )\), associated with the evolution of the eccentricities and the arguments of perihelia. The separation of these time scales is a classical approach for the study of mean motion resonances (Henrard and Caranicolas 1989; Morbidelli 2002; Batygin and Morbidelli 2013; Delisle et al. 2012, 2014).

In theory, this separation allows for two averaging of the Hamiltonian: a first averaging over the fast angle \(\lambda _2\) that we already considered in Sect. 2.3, and a second one over the semi-fast angle \(\zeta \), in order to obtain the secular Hamiltonian. In the double averaged reduced case, we would obtain a 1 degree of freedom Hamiltonian which would describe the secular dynamics of the resonance. It would add an additional parameter \(J_0\) (the action variable associated with the degree of freedom Z,\(\zeta \)). The canonical transformation for the variables \(\varPi \) and \({\varDelta \varpi } \) associated with this second averaging differs from the identity only with coefficients of the order of \({\mathcal {O}} (\sqrt{\varepsilon } )\) (Morbidelli 2002).

1.1 A.1 Adiabatic invariants

In practice, this second averaging is rather difficult because the variables Z and \(\zeta \) are not close to action-angle variables (Morais 1999, 2001; Beaugé and Roig 2001; Páez and Efthymiopoulos 2015). However, the possibility to do it gives us important information on the dynamics of the system: in the averaged reduced problem (2 degrees of freedom), the evolution of the variables \(\varPi \) and \({\varDelta \varpi } \) is of size \(\sqrt{{\varepsilon } }\) over durations of the order of \(1/\nu \), these variables can be considered as constant on a time scale short with respect to 1 / g. For sufficiently low-mass co-orbitals, we can hence consider that the variables \(\varPi \) and \({\varDelta \varpi } \) are adiabatic invariants.

1.2 A.2 Interpretation of numerical simulations

The change of coordinate (Eq. 15) from the variables of the planar 3-body problem to the variables of the averaged problem is \({\varepsilon } \) close from identity for all variables except \(\zeta _2\). Similarly, the perturbations of the semi-fast time scale on the secular variables are of size \(\sqrt{{\varepsilon } }\) (Morbidelli 2002). Thus, if we integrate numerically the full 3-body problem for co-orbital with low enough masses, for quasi-periodic orbits we can consider, on the one hand, the evolution of the variables (\(Z,\zeta \)) as their evolution in the averaged problem (they are \({\varepsilon } \) close), and on the other hand the evolution of the variables \(e_j\) and \(\varpi _j\) as their evolution in the secular problem (they are \(\sqrt{\varepsilon } \) close).

Appendix B: Reference manifold

In this section, we aim to verify that all the trajectories of the phase space pass as close as we want from the reference manifold \({\mathcal {V}} \) defined by Eq. (35).

1.1 B.1 At first order in \(e_j\)

Equation (25) holds at first order in \(e_j\). Hence, for \(m_1=m_2\), all trajectories go through the plane \(a_1=a_2\) twice per period \(2\pi /\nu \). On the other hand, near a solution of the circular coplanar case \(\zeta (t)\), the equation of variation in the direction (\(x_j,{\tilde{x}}_j\)), where the \(x_j\) are the canonical Poincaré variables defined in Eq. (4), is given by the matrix (Robutel and Pousse 2013):

$$\begin{aligned} X= \begin{pmatrix} x_1\\ x_2 \end{pmatrix} \text {et} M(t)= i{\varepsilon } \eta \frac{m'_1 m'_2}{m_0} \begin{pmatrix} \frac{A(\zeta (t))}{m'_1} &{} \frac{\bar{B}(\zeta (t))}{\sqrt{m'_1 m'_2}}\\ \frac{B(\zeta (t))}{\sqrt{m'_1 m'_2}} &{} \frac{A(\zeta (t))}{m'_2} \end{pmatrix}, \end{aligned}$$
(36)

where A and B depend on the considered trajectory and on the time. For a given trajectory, since \(\nu \gg g\), we can obtain an approximation of the secular dynamics in the direction (\(x_j,{\tilde{x}}_j\)) by averaging the expression of this matrix over a period \(2\pi /\nu \) with respect to the time t. For equal mass co-orbitals, the symmetries of this matrix give relations of the form:

$$\begin{aligned} \begin{aligned} x_1&=\alpha \sqrt{m_2} {{\mathrm{e}}}^{i \left( \frac{\pi }{3}+g t\right) } + \beta \sqrt{m_1} {{\mathrm{e}}}^{i\frac{\pi }{3}}, \\ x_2&=- \alpha \sqrt{m_1} {{\mathrm{e}}}^{i g t} + \beta \sqrt{m_2}, \end{aligned} \end{aligned}$$
(37)

with \(\alpha \) and \(\beta \) complexes. Replacing these expression in the one of \(\varPi \) (Eq. 32), and noting \(\alpha \bar{\beta }= C {\text {e}}^{ic}\), we obtain:

$$\begin{aligned} \begin{aligned} \varPi&= (\alpha \bar{\alpha }-\beta \bar{\beta })(m_1-m_2)-2C\sqrt{m_1 m_2} \cos (g t + c). \end{aligned} \end{aligned}$$
(38)

On the other hand, we have:

$$\begin{aligned} \begin{aligned} {\varDelta \varpi }&=\arg (x_1\bar{x}_2), \ \ \text {where}\ \\ x_1\bar{x}_2&=\left[ \sqrt{m_1m_2}(\beta \bar{\beta }-\alpha \bar{\alpha })+Cm_2{{\mathrm{e}}}^{i(gt+c)}-Cm_1{{\mathrm{e}}}^{-i(gt+c)}\right] {{\mathrm{e}}}^{i\pi /3}. \end{aligned} \end{aligned}$$
(39)

When \(m_1=m_2\), \(\varPi \) librates around 0 with a frequency g. Using once more the expression (32), we obtain that the quantity \(e_1^2-e_2^2\) behaves like an harmonic oscillator, librating around 0 with the frequency \(g=2|c|={\mathcal {O}} ({\varepsilon } )\). All the trajectories of the phase space hence go through the plane \(e_1=e_2\) twice per period \(2\pi /g\).

As long as \(\nu \) and g are non-resonant, all trajectories get as close as we want to the manifold \({\mathcal {V}} \) in a finite time.

Fig. 17
figure 17

Minimal value of the quantity \(\log \left( (a_1/\bar{a}-a_2/\bar{a})^2+(e_1-e_2)^2 \right) \) over \(10\times 10^{5}\) orbital periods with a step of 0.01 orbital period and a fixed value of the angular momentum \(J_1(e_1=e_2=0.4)\), with the following initial conditions: \(a_1=a_2=1\), \(m_1=m_2=10^{-5} m_0\) for the top row and \(m_2=3m_1=1.5\, 10^{-5}\) for the bottom one. a \(e_1=0.00\) and \(e_2 \approx 0.55\); b \(e_1=0.10\) and \(e_2 \approx 0.54\); c \(e_1=0.20\) and \(e_2 \approx 0.52\); d \(e_1=0.30\) and \(e_2 \approx 0.47\). The trajectories ejected before the end of the integration are identified by white pixels. See Sect. 4 for more details about the integrations

1.2 B.2 Large eccentricities

We check numerically if the definition \({\mathcal {V}} =\{a_1=a_2,e_1=e_2\}\) holds for higher eccentricities. Note that we consider only trajectories that reach \(a_1=a_2\) on their orbit. To perform this check, we take grids of initial conditions for \(\zeta \in [ 0^\circ : 360^\circ ] \) and \({\varDelta \varpi } \in [-180^\circ :180^\circ ]\) and several values of \(\varPi \) for a fixed value of \(J_1\) such that \(J_1=J_1(e_1=e_2=0.4)\). The corresponding values of the eccentricities are given by:

$$\begin{aligned} \begin{aligned} e_2=\sqrt{1-\frac{2J_1}{\varLambda ^0_1}-\sqrt{1-e_1^2}}. \end{aligned} \end{aligned}$$
(40)

Figure 17 shows the value of \((a_1/\bar{a}-a_2/\bar{a})^2+(e_1-e_2)^2\) for several values of \(\varPi \) for \(m_1=m_2\) (top line) and \(m_1 \ne m_2\) (bottom line). The integrations are conducted over \(10/{\varepsilon } \) orbital periods, hence only a few times \(2\pi /g\) at best. For all initial conditions when \(m_1=m_2\), the criterion

$$\begin{aligned} \frac{(a_1-a_2)^2}{\bar{a}^2}+(e_1-e_2)^2 < \epsilon _\varSigma \end{aligned}$$
(41)

is met for \(\epsilon _\varSigma \approx 10^{-8}\). Although this verification is not exhaustive, it suggests that the chosen reference manifold represents a significant part of the phase space of the averaged reduced problem. However it is possible that, especially at high eccentricities, stable domains appear for which the orbits never reach \(e_1=e_2\) even in the case \(m_1=m_2\), but none was discovered during this study. A study performed in the case \(e_1=e_2=0.7\) yielded similar results.

In order to compare with the case \(m_1 \ne m_2\), the bottom line of Fig. 17 shows that there are areas of the phase space where the criterion (41) is not verified for \(\epsilon '_\varSigma =10^{-4}\). There are hence orbits in the phase space that are not represented by the trajectories taking their initial conditions on the manifold \({\mathcal {V}} =\{a_1=a_2,e_1=e_2\}\). This is not surprising: we know, for example, that, at least for moderate eccentricities, the position of the \(AL_4\) equilibrium is approximated by \(m_1e_1=m_2e_2\). In the case \(m_1 \ne m_2\), any trajectory librating sufficiently close to this equilibrium would never cross the \(e_1=e_2\) manifold.

Appendix C: Identification of the \({\mathcal {F}} \) families

We show here how the separation of the time scales allows us to identify the position of the \({\mathcal {F}} \) anywhere in the phase space.

1.1 C.1 Identification of the \({{\mathcal {F}} ^{sc}}\) families

The \({{\mathcal {F}} ^{sc}}\) families are families of periodic orbit of the reduced averaged problem, whose period is associated with the secular time scale. The position of the \({{\mathcal {F}} ^{sc}}\) families can be identified by studying the critical points of the averaged Hamiltonian. Let us use the hypothesis of adiabatic invariant for the variables \(\varPi \) and \({\varDelta \varpi } \) (see Appendix A): on a short time scale with respect to 1 / g, \({{\mathcal {F}} ^{sc}}\) is made of orbits that behave as fixed points of the reduced averaged problem. The orbits belonging to \({{\mathcal {F}} ^{sc}}\) are thus orbits which satisfy:

$$\begin{aligned} \frac{\partial }{\partial Z}{\mathcal {H}} _{{\mathcal {R}} {\mathcal {M}} } = \frac{\partial }{\partial \zeta } {\mathcal {H}} _{{\mathcal {R}} {\mathcal {M}} }= 0. \end{aligned}$$
(42)

where Z and \(\zeta \) are conjugated canonical variables. This is equivalent to:

$$\begin{aligned} \dot{\zeta }= \dot{Z}= 0. \end{aligned}$$
(43)

Starting from the reduced Hamiltonian (Sect. 3.1), we can estimate the value of the averaged Hamiltonian at any point of the phase space by doing a numerical averaging over the fast angle Q. We can identify the orbits belonging to \({{\mathcal {F}} ^{sc}}\) by finding the orbits for which \(\dot{Z} = 0\) on the manifold \(\dot{\zeta }=0\). For a given value of the constants \(\varPi \) and \({\varDelta \varpi } \), we take a grid of values for \(\zeta \) and estimate the averaged Hamiltonian at each point. We can then have the approximate position of the points where \(\dot{Z} = 0\) by finding the positions on the grid where the equation

$$\begin{aligned} \frac{\partial }{\partial \zeta } {\mathcal {H}} _{{\mathcal {R}} {\mathcal {M}} }|_{\zeta =\zeta _k} \times \frac{\partial }{\partial \zeta } {\mathcal {H}} _{{\mathcal {R}} {\mathcal {M}} }|_{\zeta =\zeta _{k+1}} < 0 \end{aligned}$$
(44)

is satisfied, with

$$\begin{aligned} \frac{\partial }{\partial \zeta } {\mathcal {H}} _{{\mathcal {R}} {\mathcal {M}} }|_{\zeta =\zeta _k} = \frac{{\mathcal {H}} _{{\mathcal {R}} {\mathcal {M}} }(Z,\zeta _{k+1},{\varDelta \varpi } ,\varPi ) -{\mathcal {H}} _{{\mathcal {R}} {\mathcal {M}} }(Z,\zeta _{k-1},{\varDelta \varpi } ,\varPi ) }{|\zeta _{k+1}-\zeta _{k-1}|}. \end{aligned}$$
(45)

Note that it is not guaranteed that the associated trajectory in the full 3-body problem is quasi-periodic.

Alternatively, numerical integrations allow us to determine an empiric criterion for a numerical determination of the position of \({{\mathcal {F}} ^{sc}}\). In the various integrations that we computed through this study, we noted that the amplitude of variation of Z (hence \(a_1-a_2\)) seems not to be impacted much by the frequency g. We hence make the hypothesis that if an orbit in a regular area of the phase space verifies the condition

$$\begin{aligned} (\max {(Z)}-\min {(Z)} ) < \epsilon _\nu , \end{aligned}$$
(46)

with \(\epsilon _\nu \propto \sqrt{{\varepsilon } }\), this orbit is in the neighbourhood of the manifold \({{\mathcal {F}} ^{sc}}\). One can check in Figs. 5, 6, 7, 8, 9, 10, 11, 12, 13 and 14, where the quasi-periodic orbits that verify Eq. (46) are identified by brown pixels, and the point of the phase space satisfying Eq. (42) is identified by purple dots that both methods yield very similar results in the regular area of the phase space.

1.2 C.2 Application in the case \(m_1=m_2\)

We can apply the research of the critical points of the Hamiltonian to identify the position of \({{\mathcal {F}} ^{sc}}\) in the case \(m_1=m_2\). We assume that, as it is the case for circular co-orbitals, the manifold \(\dot{\zeta }=0\) is located at \(Z=0\). We know that the equilibriums \(L_k\) and \(AL_k\) are all located in the plane \(\varPi =0\) (\(e_1=e_2\)). We can hence explore the manifold \({\mathcal {V}} =\{Z,\zeta , \varPi ,{\varDelta \varpi } / Z=\varPi =0 \}\). We chose a grid of initial condition for \(\zeta \) and \({\varDelta \varpi } \) with a step of \(0.5^\circ \), and we compute numerically the averaged Hamiltonian at each point of the grid. In Fig. 4, we show all the points of \({\mathcal {V}} \) that verify the condition (44). Each graph corresponds to a different value of the total angular momentum (different value of \(e_1=e_2\)).

1.3 C.3 Identification of the \({{\mathcal {F}} ^{{ sf}}}\) families

The \({{\mathcal {F}} ^{{ sf}}}\) families are families of periodic orbits of the reduced averaged problem, whose period is associated with the semi-fast (resonant) time scale. The method developed in Sect. C.1 cannot be used directly to determine the position of the \({{\mathcal {F}} ^{{ sf}}}\) manifold because it requires to numerically average the Hamiltonian over the semi-fast angle \(\zeta \), which is somehow laborious, see Sect. A.

However, the evolution of the variables \({\varDelta \varpi } \) and \(\varPi \) during the numerical integrations of the 3-body problem is \({\mathcal {O}} (\sqrt{\varepsilon } )\) close to their evolution in the secular problem (see Sect. A). Since the orbits belonging to \({{\mathcal {F}} ^{{ sf}}}\) are fixed points of the 1-degree of freedom secular problem, we make the following hypothesis: all orbits in a regular area of the phase space (far from the separatrix, the chaotic and the unstable areas) that verify

$$\begin{aligned} (\max {({\varDelta \varpi } )}-\min {({\varDelta \varpi } )} ) < \epsilon _g, \end{aligned}$$
(47)

with \(\epsilon _g \propto \sqrt{{\varepsilon } }\) are in the neighbourhood of \({{\mathcal {F}} ^{{ sf}}}\). One can check that such orbits (represented by black pixel in Figs. 5, 6, 7, 8, 9, 10, 11, 12, 13 and 14) are indeed in the neighbourhood of the analytical approximation of the positions of the \({{\mathcal {F}} ^{{ sf}}}\) families, see Leleu (2016, Sect. 2.7.2).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Leleu, A., Robutel, P. & Correia, A.C.M. On the coplanar eccentric non-restricted co-orbital dynamics. Celest Mech Dyn Astr 130, 24 (2018). https://doi.org/10.1007/s10569-017-9802-8

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10569-017-9802-8

Keywords

Navigation