Skip to main content
Log in

On the Minimality of Keplerian Arcs with Fixed Negative Energy

  • Published:
Qualitative Theory of Dynamical Systems Aims and scope Submit manuscript

Abstract

We revisit a classical result by Jacobi (J Reine Angew Math 17:68–82, 1837) on the local minimality, as critical points of the corresponding energy functional, of fixed-energy solutions of the Kepler equation joining two distinct points with the same distance from the origin. Our proof relies on the Morse index theorem, together with a characterization of the conjugate points as points of geodesic bifurcation.

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

Similar content being viewed by others

References

  1. Ambrosetti, A., Coti-Zelati, V.: Periodic solutions of singular Lagrangian systems. In: Progress in Nonlinear Differential Equations and their Applications 10, Birkhäuser Boston, Inc., Boston (1993)

  2. Bahri, A., Rabinowitz, P.H.: A minimax method for a class of Hamiltonian systems with singular potentials. J. Funct. Anal. 82, 412–428 (1989)

    Article  MathSciNet  Google Scholar 

  3. Barutello, V.L., Jadanza, R.D., Portaluri, A.: Morse index and linear stability of the Lagrangian circular orbit in a three-body-type problem via index theory. Arch. Ration. Mech. Anal. 219, 387–444 (2016)

    Article  MathSciNet  Google Scholar 

  4. Barutello, V., Terracini, S., Verzini, G.: Entire parabolic trajectories as minimal phase transitions. Calc. Var. Partial Differ. Equ. 49, 391–429 (2014)

    Article  MathSciNet  Google Scholar 

  5. Boscaggin, A., Dambrosio, W., Terracini, S.: Scattering parabolic solutions for the spatial \(N\)-centre problem. Arch. Ration. Mech. Anal. 223, 1269–1306 (2017)

    Article  MathSciNet  Google Scholar 

  6. Chen, K.-C.: Existence and minimizing properties of retrograde orbits to the three-body problem with various choices of masses. Ann. Math. (2) 167, 325–348 (2008)

    Article  MathSciNet  Google Scholar 

  7. Chenciner, A., Montgomery, R.: A remarkable periodic solution of the three-body problem in the case of equal masses. Ann. Math. (2) 152, 881–901 (2000)

    Article  MathSciNet  Google Scholar 

  8. Ferrario, D.L., Terracini, S.: On the existence of collisionless equivariant minimizers for the classical n-body problem. Invent. Math. 155, 305–362 (2004)

    Article  MathSciNet  Google Scholar 

  9. Fusco, G., Gronchi, G.F., Negrini, P.: Platonic polyhedra, topological constraints and periodic solutions of the classical \(N\)-body problem. Invent. Math. 185, 283–332 (2011)

    Article  MathSciNet  Google Scholar 

  10. Geiges, H.: The Geometry of Celestial Mechanics. London Mathematical Society Student Texts, Cambridge University Press, Cambridge (2016)

  11. Gordon, W.B.: A minimizing property of Keplerian orbits. Am. J. Math. 99, 961–971 (1977)

    Article  MathSciNet  Google Scholar 

  12. Gutzwiller, M.C.: Chaos in Classical and Quantum Mechanics, Interdisciplinary Applied Mathematics, vol. 1. Springer, New York (1990)

    Book  Google Scholar 

  13. Hu, X., Sun, S.: Morse index and stability of elliptic Lagrangian solutions in the planar three-body problem. Adv. Math. (1) 223, 98–119 (2010)

    Article  MathSciNet  Google Scholar 

  14. Jacobi, C.G.J.: Zur Theorie der variations-Rechnung und der differential-Gleichungen. J. Reine Angew. Math. 17, 68–82 (1837)

    MathSciNet  Google Scholar 

  15. Kavle, H., Offin, D., Portaluri, A.: Keplerian orbits through the Conley–Zehnder index, arXiv:1908.00075v1

  16. Klingenberg, W.P.A.: Riemannian geometry. Second edition, De Gruyter Studies in Mathematics 1, Walter de Gruyter & Co., Berlin (1995)

  17. Maderna, E., Venturelli, A.: Globally minimizing parabolic motions in the Newtonian \(N\)-body problem. Arch. Ration. Mech. Anal. 194, 283–313 (2009)

    Article  MathSciNet  Google Scholar 

  18. Montgomery, R.: Minimizers for the Kepler problem, preprint

  19. Ortega, R., Ureña, A.J.: Introducción a la Mecánica Celeste. Universidad de Granada, Granada (2010)

    Google Scholar 

  20. Piccione, P., Portaluri, A., Tausk, D.V.: Spectral flow, Maslov index and bifurcation of semi-Riemannian geodesics. Ann. Global Anal. Geom. 25, 121–149 (2004)

    Article  MathSciNet  Google Scholar 

  21. Portaluri, A., Waterstraat, N.: Yet another proof of the Morse index theorem. Expo. Math. 33, 378–386 (2015)

    Article  MathSciNet  Google Scholar 

  22. Soave, N., Terracini, S.: Symbolic dynamics for the \(N\)-centre problem at negative energies. Discrete Contin. Dyn. Syst. 32, 3245–3301 (2012)

    Article  MathSciNet  Google Scholar 

  23. Terracini, S., Venturelli, A.: Symmetric trajectories for the \(2N\)-body problem with equal masses. Arch. Ration. Mech. Anal. 184, 465–493 (2007)

    Article  MathSciNet  Google Scholar 

  24. Todhunter, M.A.: Researches in the Calculus of Variation, Principally on the Theory of Discontinuous Solutions. Macmillan, Cambridge (1871)

    MATH  Google Scholar 

  25. Wintner, A.: Analytical foundation of Celestial Mechanics, Princeton Mathematical series, vol. 5, Princeton University Press, Princeton (1947)

  26. Yu, G.: Simple choreographies of the planar Newtonian \(N\)-body problem. Arch. Ration. Mech. Anal. 225, 901–935 (2017)

    Article  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vivina Barutello.

Additional information

In memory of Florin Diacu.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Work supported by the ERC Advanced Grant 2013 No. 339958 Complex Patterns for Strongly Interacting Dynamical Systems - COMPAT. Work written under the auspices of the Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM). In particular, A.B. acknowledges the support of the INdAM-GNAMPA Project “Il modello di Born-Infeld per l’elettromagnetismo nonlineare: esistenza, regolarità e molteplicità di soluzioni”.

Appendix: A Brief Recap on Geodesics and their Morse Index

Appendix: A Brief Recap on Geodesics and their Morse Index

In this section, we collect some known facts about geodesics and conjugate points. We will work in the simplified setting when the manifold is an open set of the Euclidean space, endowed with a Riemannian metric conformal to the standard one. For a more general treatment, we refer to [16].

1.1 Definition and First Properties

Let M be an open subset of \(\mathbb {R}^N\); as such, M has a natural structure of differentiable manifold of dimension N. Given a smooth function \(W: M \rightarrow \mathbb {R}\) satisfying \(W(x) > 0\) for every \(x \in M\), we can endow M with the Riemannian structure given by

$$\begin{aligned} g_W(x)[u,v] = W(x) \langle u,v \rangle , \qquad x \in M, \,u,v \in \mathbb {R}^N, \end{aligned}$$

where \(\langle \cdot , \cdot \rangle \) denotes the Euclidean product on \(\mathbb {R}^N\). Notice that smooth curves on M are nothing but curves with values in M which are smooth in the usual sense; however, the velocity and the length of a curve are influenced by the function W.

A geodesic on the Riemannian manifold \((M,g_W)\) is a smooth curve \(\gamma : I \rightarrow M\) (with \(I \subset \mathbb {R}\) an interval) solving the equation

$$\begin{aligned} \frac{d}{ds} \left( W(\gamma ) {{\dot{\gamma }}} \right) = \frac{1}{2} \vert {{\dot{\gamma }}} \vert ^2 \nabla W(\gamma ), \end{aligned}$$
(18)

where \(\nabla W\) is the gradient of the function W with respect to the standard Euclidean structure. By the standard theory of ODEs, for any \(p \in M\) and \(v \in \mathbb {R}^N\), there exists a unique (maximal) geodesic \(\gamma _{p,v}\) satisfying \(\gamma _{p,v}(0) = p\) and \({{\dot{\gamma }}}_{p,v}(0) = v\).

We collect in the next proposition some simple properties which can be readily obtained from this definition.

Proposition 4.1

The following facts hold true:

  1. (i)

    Geodesics have constant velocity: namely, for any geodesic \(\gamma \), there exists \(c \in \mathbb {R}\) such that \(\vert \dot{\gamma }(s) \vert \sqrt{W(\gamma (s))} \equiv c\);

  2. (ii)

    For every geodesic \(\gamma \) and for every \(a,b \in \mathbb {R}\) with \(a \ne 0\), the curve \({{\tilde{\gamma }}}(s) = \gamma (as + b)\) is a geodesic, as well;

  3. (iii)

    (Rescaling Lemma) For any \(\lambda \in \mathbb {R}\) with \(\lambda \ne 0\), it holds that \(\gamma _{p,\lambda v}(s) = \gamma _{p,v}(\lambda s)\) for any \(s \in \mathbb {R}\) such that both sides of the equation are defined.

In view of this proposition, geodesics are often considered as geometric objects, rather than as parameterized curves. In some textbooks, curves which can be parameterized as geodesics are called pre-geodesics.

1.2 The Energy Functional and the Morse Index Theorem

Equation (18) can be meant as the Euler–Lagrange equation of the Lagrangian \(L(\gamma ,{{\dot{\gamma }}}) = \vert {{\dot{\gamma }}} \vert ^2 W(\gamma )\); thus, geodesics can be regarded as critical points of the functional \(E(\gamma ) = \int L(\gamma ,{{\dot{\gamma }}}) \,ds\). Below, we describe in more details this variational principle, by focusing on the case of geodesics joining two points \(p,q \in M\) (a similar discussion can be made when periodic boundary conditions are taken into account, leading to the so-called closed geodesics).

Let us consider the functional

$$\begin{aligned} E(\gamma ) = \int _0^1 \vert {{\dot{\gamma }}}(s)\vert ^2 W(\gamma (s))\,ds, \end{aligned}$$

defined on the Hilbert manifold \(\Omega _{p,q}\) of \(H^1\)-paths \(\gamma : [0,1] \rightarrow M\) satisfying \(\gamma (0) = p\) and \(\gamma (1) = q\). It is standard to verify that E is a smooth functional on \(\Omega _{p,q}\) with first Frechet differential given by

$$\begin{aligned} dE(\gamma )[\xi ] = \int _0^1 \left( 2 \langle {{\dot{\gamma }}}(s),{{\dot{\xi }}}(s)\rangle W(\gamma (s)) + \vert {{\dot{\gamma }}}(s) \vert ^2 \langle \nabla W(\gamma (s)),\xi (s)\rangle \right) \,ds, \end{aligned}$$

for any \(\xi \in H^1_0([0,1];\mathbb {R}^2)\). A well-known argument based on integration by parts yields:

Proposition 4.2

A curve \(\gamma : [0,1] \rightarrow M\) satisfying \(\gamma (0) = p\) and \(\gamma (1) = q\) is a geodesic if and only if it is a critical point of the functional E.

We are now in a position to define the crucial concept of this section. By Morse index\(\mu (\gamma )\) of a geodesic \(\gamma \) we mean its Morse index as a critical point of the functional E, that is, the dimension of the maximal subspace \(V \subset H^1_0([0,1];\mathbb {R}^2)\) such that \(d^2 E(\gamma )[\xi ,\xi ] < 0\) for any \(\xi \in V\). Since

$$\begin{aligned} d^2 E(\gamma )[\xi ,\xi ]&= 2\int _0^1 \vert {{\dot{\xi }}}(s)\vert ^2 W(\gamma (s))\,ds + 4 \int _0^1 \langle {{\dot{\gamma }}}(s),{{\dot{\xi }}}(s) \rangle \langle \nabla W(\gamma (s),\xi (s)\rangle \,ds \\&+ \int _0^1 \vert {{\dot{\gamma }}}(s) \vert ^2 \langle \nabla ^2 W(\gamma (s))\xi (s),\xi (s)\rangle \,ds \end{aligned}$$

for any \(\xi \in H^1_0([0,1];\mathbb {R}^2)\), functions \(\xi \) lying in the kernel of \(d^2 E(\gamma )\) are nothing but solutions of the linear equation

$$\begin{aligned}&\frac{d}{ds}\Big ( W(\gamma (s)) {{\dot{\xi }}} + \langle \nabla W(\gamma (s),\xi \rangle {{\dot{\gamma }}}(s) \Big ) \nonumber \\&\quad = \frac{1}{2} \vert \dot{\gamma }(s) \vert ^2 \nabla ^2 W(\gamma (s))\xi + \langle {{\dot{\gamma }}}(s),{{\dot{\xi }}} \rangle \nabla W(\gamma (s)) \end{aligned}$$
(19)

satisfying the boundary condition \(\xi (0) = \xi (1) = 0\).

The celebrated Morse index theorem asserts that the Morse index \(\mu (\gamma )\) equals the number of some special points along \(\gamma \), the so-called conjugate points. Precisely, we say that the point \(\gamma (s^*) = p_{s^*}\), with \(s^* \in \mathopen {]}0,1\mathclose {]}\), is conjugate to \(\gamma (0) = p\) along \(\gamma \) if the linear equation (19) admits nontrivial solutions \(\xi \) satisfying \(\xi (0) = \xi (s^*) = 0\). The multiplicity \(m(p_{s^*})\) of the conjugate point is, by definition, the dimension of the space of such solutions. Incidentally, notice that \(\gamma (1) = q\) is conjugate to \(\gamma (0)\) along \(\gamma \) if and only if the kernel of \(d^2 E(\gamma )\) is not trivial (that is, if and only if \(\gamma \) is a degenerate critical point of E).

With this in mind, the Morse index theorem reads as follows (see for instance [21] and the references therein).

Theorem 4.3

(Morse Index Theorem) Let \(\gamma : [0,1] \rightarrow M\) be a non-constant geodesic. Then the set of conjugate points along \(\gamma \) is finite and

$$\begin{aligned} \mu (\gamma ) = \sum _{s^* \in \,\mathopen {]}0,1\mathclose {[}} m(p_{s^*}). \end{aligned}$$

Remark 4.4

The choice of parameterizing the geodesic \(\gamma \) on the interval [0, 1] is completely conventional: any other interval could be used, since geodesics are preserved by affine reparameterizations (recall Proposition 4.1). Of course, conjugate points and the Morse index do not depend on this choice.

1.3 Bifurcation of Geodesics

A key role in our arguments will be played by the notion of geodesic bifurcation, as introduced in the paper [20].

Definition 4.5

Let \(\gamma : [0,1] \rightarrow M\) be a non-constant geodesic. A point \(\gamma (s^*)\), with \(s^* \in \mathopen {]}0,1\mathclose {[}\), is a bifurcation point along \(\gamma \) if there exists a sequence \(\{s_n\}_n \subset [0,1]\) with \(s_n \rightarrow s^*\) and a sequence of geodesics \(\{\gamma _n\}_n\) (defined on [0, 1]), with \(\gamma _n \ne \gamma \), such that

  1. (i)

    \(\gamma _n(0) = \gamma (0)\), for every n,

  2. (ii)

    \(\gamma _n(s_n) = \gamma (s_n)\), for every n,

  3. (iii)

    \({{\dot{\gamma }}}_n(0) \rightarrow {{\dot{\gamma }}}(0)\), for \(n \rightarrow +\infty \).

According to the discussion in [20, p. 122] together with [20, Corollary 5.6], the relationship between conjugate points and bifurcation points can be stated as follows.

Theorem 4.6

Let \(\gamma : [0,1] \rightarrow M\) be a non-constant geodesic. Then, a point \(\gamma (s^*)\), with \(s^* \in \mathopen {]}0,1\mathclose {[}\), is a bifurcation point if and only if it is conjugate to \(\gamma (0)\) along \(\gamma \).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Barutello, V., Boscaggin, A. & Dambrosio, W. On the Minimality of Keplerian Arcs with Fixed Negative Energy. Qual. Theory Dyn. Syst. 19, 42 (2020). https://doi.org/10.1007/s12346-020-00362-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s12346-020-00362-9

Keywords

Navigation