Skip to main content
Log in

Linear Stability of an Elliptic Relative Equilibrium in the Spatial \(n\)-Body Problem via Index Theory

  • Published:
Regular and Chaotic Dynamics Aims and scope Submit manuscript

This article has been updated

Abstract

It is well known that a planar central configuration of the \(n\)-body problem gives rise to a solution where each particle moves in a Keplerian orbit with a common eccentricity \(\mathfrak{e}\in[0,1)\). We call this solution an elliptic relative equilibrium (ERE for short). Since each particle of the ERE is always in the same plane, it is natural to regard it as a planar \(n\)-body problem. But in practical applications, it is more meaningful to consider the ERE as a spatial \(n\)-body problem (i. e., each particle belongs to \(\mathbb{R}^{3}\)). In this paper, as a spatial \(n\)-body problem, we first decompose the linear system of ERE into two parts, the planar and the spatial part. Following the Meyer – Schmidt coordinate [19], we give an expression for the spatial part and further obtain a rigorous analytical method to study the linear stability of the spatial part by the Maslov-type index theory. As an application, we obtain stability results for some classical ERE, including the elliptic Lagrangian solution, the Euler solution and the \(1+n\)-gon solution.

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

Change history

  • 29 October 2023

    The numbering issue has been changed to 4-5.

References

  1. Abbondandolo, A. and Majer, P., Ordinary Differential Operators in Hilbert Spaces and Fredholm Pairs, Math. Z., 2003, vol. 243, no. 3, pp. 525–562.

    Article  MathSciNet  Google Scholar 

  2. Arnol’d, V. I., On a Characteristic Class Entering into Conditions of Quantization, Funct. Anal. Appl., 1967, vol. 1, no. 1, pp. 1–13; see also: Funktsional. Anal. i Prilozhen., 1967, vol. 1, no. 1, pp. 1-14.

    Article  MathSciNet  MATH  Google Scholar 

  3. Asselle, L., Portaluri, A., and Wu, L., Spectral Stability, Spectral Flow and Circular Relative Equilibria for the Newtonian \(n\)-Body Problem, J. Differential Equations, 2022, vol. 337, pp. 323–362.

    Article  MathSciNet  MATH  Google Scholar 

  4. Arnol’d, V. I., Mathematical Methods of Classical Mechanics, 2nd ed., Grad. Texts in Math., vol. 60, New York: Springer, 1989.

    Book  MATH  Google Scholar 

  5. Atiyah, M. F., Patodi, V. K., and Singer, I. M., Spectral Asymmetry and Riemannian Geometry: 3, Math. Proc. Cambridge Philos. Soc., 1976, vol. 79, no. 1, pp. 71–99.

    Article  MathSciNet  MATH  Google Scholar 

  6. Cappell, S. E., Lee, R., and Miller, E. Y., On the Maslov Index, Comm. Pure Appl. Math., 1994, vol. 47, no. 2, pp. 121–186.

    Article  MathSciNet  MATH  Google Scholar 

  7. Hu, X., Lei, L., Ou, Y., Salomão, A. S., and Yu, G., A Symplectic Dynamics Approach to the Spatial Isosceles Three-Body Problem, arXiv:2308.00338 (2023).

  8. Hu, X., Long, Y., and Sun, S., Linear Stability of Elliptic Lagrangian Solutions of the Planar Three-Body Problem via Index Theory, Arch. Ration. Mech. Anal., 2014, vol. 213, no. 3, pp. 993–1045.

    Article  MathSciNet  MATH  Google Scholar 

  9. Hu, X., Ou, Y., and Wang, P., Trace Formula for Linear Hamiltonian Systems with Its Applications to Elliptic Lagrangian Solutions, Arch. Ration. Mech. Anal., 2015, vol. 216, no. 1, pp. 313–357.

    Article  MathSciNet  MATH  Google Scholar 

  10. Hu, X. and Ou, Y., An Estimation for the Hyperbolic Region of Elliptic Lagrangian Solutions in the Planar Three-Body Problem, Regul. Chaotic Dyn., 2013, vol. 18, no. 6, pp. 732–741.

    Article  MathSciNet  MATH  Google Scholar 

  11. Hu, X. and Ou, Y., Collision Index and Stability of Elliptic Relative Equilibria in Planar \(n\)-Body Problem, Comm. Math. Phys., 2016, vol. 348, no. 3, pp. 803–845.

    Article  MathSciNet  MATH  Google Scholar 

  12. Hu, X., Long, Y., and Ou, Y., Linear Stability of the Elliptic Relative Equilibrium with \((1+n)\)-gon Central Configurations in Planar \(n\)-Body Problem, Nonlinearity, 2020, vol. 33, no. 3, pp. 1016–1045.

    Article  MathSciNet  MATH  Google Scholar 

  13. Hu, X. and Portaluri, A., Index Theory for Heteroclinic Orbits of Hamiltonian Systems, Calc. Var. Partial Differential Equations, 2017, vol. 56, no. 6, Paper No. 167, 24 pp.

    Article  MathSciNet  MATH  Google Scholar 

  14. Hu, X. and Sun, S., Morse Index and Stability of Elliptic Lagrangian Solutions in the Planar Three-Body Problem, Adv. Math., 2010, vol. 223, no. 1, pp. 98–119.

    Article  MathSciNet  MATH  Google Scholar 

  15. Hu, X., Wu, L., and Yang, R., Morse Index Theorem of Lagrangian Systems and Stability of Brake Orbit, J. Dynam. Differential Equations, 2020, vol. 32, no. 1, pp. 61–84.

    Article  MathSciNet  MATH  Google Scholar 

  16. Long, Y., Bott Formula of the Maslov-Type Index Theory, Pacific J. Math., 1999, vol. 187, no. 1, pp. 113–149.

    Article  MathSciNet  MATH  Google Scholar 

  17. Long, Y., Index Theory for Symplectic Paths with Applications, Progr. Math., vol. 207, Basel: Birkhäuser, 2002.

    Book  MATH  Google Scholar 

  18. Magnus, W. and Winkler, S., Hill’s Equation, Intersci. Tracts Pure Appl. Math., No. 20, New York: Wiley, 1966.

    MATH  Google Scholar 

  19. Meyer, K. R. and Schmidt, D. S., Elliptic Relative Equilibria in the \(N\)-Body Problem, J. Differential Equations, 2005, vol. 214, no. 2, pp. 256–298.

    Article  MathSciNet  MATH  Google Scholar 

  20. Martínez, R., Samà, A., and Simó, C., Analysis of the Stability of a Family of Singular-Limit Linear Periodic Systems in \(\mathbb{R}^{4}\): Applications, J. Differential Equations, 2006, vol. 226, no. 2, pp. 652–686.

    Article  MathSciNet  MATH  Google Scholar 

  21. Martínez, R. and Samà, A., On the Centre Manifold of Collinear Points in the Planar Three-Body Problem, Celestial Mech. Dynam. Astronom., 2003, vol. 85, no. 4, pp. 311–340.

    Article  MathSciNet  MATH  Google Scholar 

  22. Moeckel, R., On Central Configurations, Math. Z., 1990, vol. 205, no. 4, pp. 499–517.

    Article  MathSciNet  MATH  Google Scholar 

  23. Zhou, Q., Trace Estimation of a Family of Periodic Sturm – Liouville Operators with Application to Robe’s Restricted Three-Body Problem, J. Math. Phys., 2019, vol. 60, no. 5, 053503, 14 pp.

    Article  MathSciNet  MATH  Google Scholar 

  24. Zhou, Q. and Long, Y., Maslov-Type Indices and Linear Stability of Elliptic Euler Solutions of the Three-Body Problem, Arch. Ration. Mech. Anal., 2017, vol. 226, no. 3, pp. 1249–1301.

    Article  MathSciNet  MATH  Google Scholar 

Download references

ACKNOWLEDGMENTS

The authors sincerely thank the referee for his/her careful reading of the manuscript and valuable comments.

Funding

The work of all authors was supported by the National Key R&D Program of China (2020YFA0713303) and NSFC (No. 12071255). The second author was also supported by NSFC (No. 12371192) and the Qilu Young Scholar Program of Shandong University.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yuwei Ou.

Ethics declarations

The authors declare that they have no conflicts of interest.

Additional information

Dedicated to Professor Alain Chenciner on the occasion of his 80th birthday

MSC2010

37J25, 70F10, 53D12

APPENDIX

We start this subsection with a brief review of the Maslov index theory [2, 6]. A Lagrangian subspace \(V\) is an \(n\)-dimensional subspace with \(V|_{\Omega}=0\). We let \(\mathrm{Lag}(2n)\) denote the Lagrangian Grassmanian, i. e., the set of Lagrangian subspaces of \((\mathbb{R}^{2n},\Omega)\). For two continuous paths \(L_{1}(t)\) and \(L_{2}(t)\), \(t\in[a,b]\) in \(\mathrm{Lag}(2n)\), the Maslov index \(\mu(L_{1},L_{2})\) is an integer invariant. Here we use the definition from [6]. We list several properties of the Maslov index. Details can be found in [6].

Property I (nullity). For \(L_{1},L_{2}\in C\big{(}[a,b],\mathrm{Lag}(2n)\big{)}\) such that \(\dim L_{1}(t)\cap L_{2}(t)\) is independent of \(t\), then

$$\mu\big{(}L_{1}(t),L_{2}(t)\big{)}=0.$$

Property II (homotopy invariant). For two continuous families of the Lagrangian path \(L_{1}(s,t)\), \(L_{2}(s,t)\), \(0\leqslant s\leqslant 1\), \(a(s)\leqslant t\leqslant b(s)\), and satisfies \(\dim L_{1}\big{(}s,a(s)\big{)}\cap L_{2}\big{(}s,a(s)\big{)}\) and \(\dim L_{1}\big{(}s,b(s)\big{)}\cap L_{2}\big{(}s,b(s)\big{)}\) is constant, then

$$\mu\big{(}L_{1}(0,t),L_{2}(0,t)\big{)}=\mu\big{(}L_{1}(1,t),L_{2}(1,t)\big{)}.$$

Property III (path additivity). If \(a<c<b\), then

$$\mu\big{(}L_{1}(t),L_{2}(t)\big{)}=\mu(L_{1}(t),L_{2}(t)|_{[a,c]})+\mu(L_{1}(t),L_{2}(t)|_{[c,b]}).$$

Property IV (symplectic invariance) Let \(\gamma(t)\), \(t\in[a,b]\) be a continuous path in \(\mathrm{Sp}(2n)\), then

$$\mu\big{(}L_{1}(t),L_{2}(t)\big{)}=\mu\big{(}\gamma(t)L_{1}(t),\gamma(t)L_{2}(t)\big{)}.$$

Property V (symplectic additivity). Let \(L_{1},L_{2}\in C\big{(}[a,b],\mathrm{Lag}(W_{1})\big{)}\), \(\hat{L}_{1},\hat{L}_{2}\in C\big{(}[a,b], \mathrm{Lag}(W_{2})\big{)}\), where \(W_{i}\), \(i=1,2\) are symplectic spaces, then

$$\mu\big{(}L_{1}(t)\oplus\hat{L}_{1}(t),L_{2}(t)\oplus\hat{L}_{2}(t)\big{)}=\mu\big{(}L_{1}(t),L_{2}(t)\big{)}+\mu\big{(}\hat{L}_{1}(t),\hat{L}_{2}(t)\big{)}.$$

The Maslov index is easily explained for \(n=1\). Obviously, \(\mathrm{Lag}(2)\) consists of a straight line passing through the origin. We define a map \(\mathcal{F}:\mathrm{Lag}(2)\to\mathbb{U}\) as follows. For \(V=\mathbb{R}\big{(}\cos(\theta),\sin(\theta)\big{)}^{T}\), we let \(\mathcal{F}(V)=\exp(2i\theta)\). For two paths \(V_{0},V_{1}\in C\big{(}[a,b],\mathrm{Lag}(2)\big{)}\), then \(\mathcal{F}(V_{0})^{-1}\mathcal{F}(V_{1})(t)\in\mathbb{U}\). Let \(x(t)\) be a continuous path which satisfies

$$\exp\big{(}2\pi ix(t)\big{)}=\mathcal{F}(V_{0})^{-1}\mathcal{F}(V_{1})(t),t\in[a,b].$$
Set \([x]:=\sup\{z\in\mathbb{Z},z\leqslant x\}\). The Maslov index
$$\mu(V_{0},V_{1};[a,b])=[x(b)-\epsilon]-[x(a)-\epsilon]$$
for \(\epsilon\) small enough.

In order to study the case for \(\mathfrak{e}\to 1\), we need the index theory for heteroclinic orbits [11, 13]. We consider the Hamiltonian flow induced by

$$\dot{z}=JB(\tau)z,\ \ \tau\in\mathbb{R}.$$
(A.1)
We assume

(H1) \(JB(\pm\infty)=\lim_{\tau\rightarrow\pm\infty}JB(\tau)\) is hyperbolic, that is, there are no eigenvalues of \(JB(\pm\infty)\) on the imaginary line. It follows that \(\mathbb{R}^{2n}=V^{\pm}_{s}\oplus V^{\pm}_{u}\), where \(V^{\pm}_{s}(V^{\pm}_{u})\) is the stable subspace(resp. unstable subspace) of the equilibria which is spanned by the generalized eigenvector of the eigenvalue with the negative real part (positive real part) of \(JB(\pm\infty)\). Both the stable subspace \(V^{\pm}_{s}\) and the unstable subspace \(V^{\pm}_{u}\) are Lagrangian subspaces of \((\mathbb{R}^{2n},\Omega)\).

Under the condition of (H1), \(\mathcal{A}:=-J\frac{d}{d\tau}-B\) is a Fredholm operator on \(L^{2}(\mathbb{R},\mathbb{R}^{2n})\) with domain \(W^{1,2}(\mathbb{R},\mathbb{R}^{2n})\). In some cases, we need to consider the solution \(z\) of (A.1) defined on \((-\infty,0]\) and satisfying \(\lim_{\tau\to-\infty}z(\tau)=0\), \(z(0)\in\Lambda\) for \(\Lambda\in \mathrm{Lag}(2n)\). Let

$$D(\Lambda,\mathbb{R}^{-}):=\{x\in W^{1,2}(\mathbb{R}^{-},\mathbb{R}^{2n}),x(0)\in\Lambda\},$$
then \(\mathcal{A}|_{D(\Lambda,\mathbb{R}^{-})}\) is a Fredholm self-adjoint operator on \(L^{2}(\mathbb{R},\mathbb{R}^{2n})\).

Let \(\gamma(t,\tau)\) satisfy (A.1) with \(\gamma(\tau,\tau)=I_{2n}\). In what follows we set \(\gamma(t):=\gamma(t,0)\). Clearly, \(\gamma\) satisfies a semigroup property, that is, \(\gamma(t,\nu)\gamma(\nu,\tau)=\gamma(t,\tau)\). For \(\tau\in\mathbb{R},\) define

$$V_{s}(\tau)=\{\zeta\ \ |\ \ \zeta\in\mathbb{R}^{2n}\ \ \text{and}\ \ \lim_{t\rightarrow\infty}\gamma(t,\tau)\zeta=0\},\quad V_{u}(\tau)=\{\zeta\ \ |\ \ \zeta\in\mathbb{R}^{2n}\ \ \text{and}\ \ \lim_{t\rightarrow-\infty}\gamma(t,\tau)\zeta=0\}.$$
We remark that
$$\lim_{\nu\rightarrow\infty}V_{s}(\tau)=V^{+}_{s}\ \ \text{and}\ \ \lim_{\nu\rightarrow-\infty}V_{u}(\tau)=V^{-}_{u}.$$
It is well known that both \(V_{s}(\tau)\) and \(V_{u}(\tau)\) are Lagrangian subspaces of \((\mathbb{R}^{2n},\Omega)\). An important property from [1] is the following: if \(V\) is transversal to \(V_{s}(0)\), then
$$\lim_{t\rightarrow+\infty}\gamma(t,0)V=V^{+}_{u}.$$
Similarly, if is \(V\) transversal to \(V_{u}(0)\), then
$$\lim_{t\rightarrow-\infty}\gamma(t,0)V=V^{-}_{s}.$$

For \(\Lambda\in \mathrm{Lag}(2n)\), \(V_{u}\) is the unstable path of (A.1), the index for the past half-clinic orbit is defined by

$$\iota_{-}(\Lambda,B)=\mu(\Lambda,V_{u}(\tau),\tau\in(-\infty,0]),$$
which is well defined since \(V_{u}(-\infty)\) exist. Note that
$$\mu(\Lambda,V_{u}(\tau),\tau\in(-\infty,0])=-\mu\big{(}\Lambda,V_{u}(-\tau);\tau\in[0,+\infty)\big{)},$$
so \(\iota_{-}(\Lambda,B)\) is just the definition of the geometrical index given in [13]. We also denote \(\nu(\mathcal{A})=\dim\ker(\mathcal{A}),\) then
$$\nu(\mathcal{A})=\dim V_{s}(0)\cap V_{u}(0)\leqslant n.$$
Similarly,
$$\nu(\mathcal{A}|_{D(\Lambda,\mathbb{R}^{-})})=\dim\Lambda\cap V_{u}(0)\leqslant n.$$

Now we consider the family of the linear Hamiltonian system

$$\dot{z}=JB_{\lambda}(\tau)z\quad\tau\in\mathbb{R}\quad\lambda\in[a,b].$$
We assume that \(B_{\lambda}(\tau)\) converges to \(B_{\lambda}(\pm\infty)\) uniformly with respect to \(\lambda\), and that \(JB_{\lambda}(\pm\infty)\) is hyperbolic. It is well known that \(\mathcal{A}_{\lambda}:=-J\frac{d}{d\tau}-B_{\lambda}\) and \(\mathcal{A}_{\lambda}|_{D(\Lambda,\mathbb{R}^{-})}\) is a continuous path of Fredholm operators, so its spectral flow is well defined. The spectral flow was first introduced by Atiyah, Patodi and Singer [5] in their study of index theory on manifolds with boundary. Let \(\{A(\lambda),\lambda\in[0,1]\}\) be a continuous path of self-adjoint Fredholm operators on a Hilbert space \(E\). Roughly speaking, the spectral flow of path \(\{A(\lambda),\lambda\in[0,1]\}\) counts the net change in the number of negative eigenvalues of \(A(\lambda)\) as \(\lambda\) goes from \(0\) to \(1\), where the enumeration follows from the rule that each negative eigenvalue crossing the positive axis contributes \(+1\) and each positive eigenvalue crossing the negative axis contributes \(-1\), and for each crossing the multiplicity of the eigenvalue is counted. We refer the reader to [13] for a brief introduction.

From [13], we have the index theorem

Theorem 8

Under the above notation,

$$-sf(\mathcal{A}_{\lambda}|_{D(\Lambda,\mathbb{R}^{-})})=\iota_{-}(\Lambda,B_{b})-\iota_{-}(\Lambda,B_{a})+\mu(\Lambda,V^{\lambda}_{u}(-\infty);[a,b]).$$
(A.2)

The Maslov-type index for a symplectic path is a very useful tool for the study of the stability problem of periodic orbits. In this section, we give only a brief review. Details can be found in [17]. For \(\tau>0\) we are interested in paths in \(\mathrm{Sp}(2n)\):

$$\mathcal{P}_{\tau}(2n)=\left\{\gamma\in C\big{(}[0,\tau],\operatorname{Sp}(2n)\big{)}\mid\gamma(0)=I_{2n}\right\}.$$
For any \(\omega\in\mathbb{U}\)(the unit circle in \(\mathbb{C}\)), the following \(\omega\)-degenerate hypersurface of codimension one in \(\mathrm{Sp}(2n)\) is defined [17]:
$$\mathrm{Sp}(2n)_{\omega}^{0}=\{M\in \mathrm{Sp}(2n)|\det(M-\omega I_{2n})=0\}.$$
Moreover, the \(\omega\)-regular set of \(\mathrm{Sp}(2n)\) is defined by \(\mathrm{Sp}(2n)^{*}_{2n}=\mathrm{Sp}(2n)\setminus \mathrm{Sp}(2n)_{\omega}^{0}\).

For \(M\in \mathrm{Sp}(2n)_{\omega}^{0}\), we define a coorientation of \(\mathrm{Sp}(2n)_{\omega}^{0}\) at \(M\) by the positive direction \(\frac{d}{dt}Me^{tJ}|_{t=0}\) of the path \(Me^{tJ}\) with \(|t|\) sufficiently small. Now we will give a definition of the \(\omega\)-index [17].

Definition 1

For \(\omega\in\mathbb{U}\), \(\gamma\in\mathcal{P}_{\tau}\), the \(\omega\)-index of \(\gamma\) can be defined as

$$i_{\omega}(\gamma)=\begin{cases}[e^{-\epsilon J}\gamma:\mathrm{Sp}(2n)_{\omega}^{0}]-n&\text{ if }\omega=1\\ [e^{-\epsilon J}\gamma:\mathrm{Sp}(2n)_{\omega}^{0}] &\text{ if }\omega\neq 1.\end{cases}$$
For \(\epsilon>0\) small enough, where \([\cdot:\cdot]\) is the intersection number, we also denote
$$\nu_{\omega}(\gamma)=\dim_{\mathbb{C}}\ker_{\mathbb{C}}(\gamma(\tau)-\omega I).$$

The Maslov-type index satisfies the homotopy invariant property, that is, for a family of \(\gamma_{s}\in\mathcal{P}_{\tau}(2n)\) continuously depending on \(s\in[0,1]\), we have that, if \(\nu_{\omega}(\gamma_{s})\) is constant, then \(i_{\omega}(\gamma_{0})=i_{\omega}(\gamma_{1})\). Note that, for \(M\in \mathrm{Sp}(2n)\), \(\mathrm{Gr}(M):=\{(x,Mx)|x\in\mathbb{R}^{2n}\}\) is a Lagrangian subspace of the symplectic vector space \((\mathbb{R}^{2n}\oplus\mathbb{R}^{2n},-\Omega\oplus\Omega)\). For \(\gamma\in\mathcal{P}_{\tau}(2n)\), we have

$$i_{1}(\gamma)+n=\mu \Big{(}\mathrm{Gr}(I),\mathrm{Gr}\big{(}\gamma(t)\big{)}\Big{)},$$
and
$$i_{-1}(\gamma)=\mu\Big{(}\mathrm{Gr}(-I),\mathrm{Gr}\big{(}\gamma(t)\big{)}\Big{)}.$$

The Maslov-type index is essentially equal to the Morse index. Consider a second-order system \(\ddot{x}=D(t)x\), \(t\in[0,\tau]\). Let \(A=-\frac{d^{2}}{dt^{2}}+D\), which is a self-adjoint operator on \(L^{2}([0,\tau],\mathbb{C}^{n})\) with domain

$$\bar{D}(\omega,\tau)=\{W^{2,2}([0,\tau],\mathbb{C}^{n})|y(\tau)=\omega y(0),\dot{y}(\tau)=\omega\dot{y}(0)\}.$$
We set \(v_{\omega}(A)=\dim\ker(A)\) and denote by \(\phi_{\omega}(A)\) its Morse index, then from [17], we have the fact
$$\phi_{\omega}(A)=i_{\omega}(\gamma),\ \ v_{\omega}(A)=\nu_{\omega}(\gamma),$$
(A.3)
where \(\gamma\) is the fundamental solution of the corresponding Hamiltonian systems, that is, \(\dot{\gamma}=J\left(\begin{array}[]{cc}I_{n}&0_{n}\\ 0_{n}&-D(t)\end{array}\right)\gamma,\) with \(\gamma(0)=I\). For a formula for general boundary conditions, see [15].

For \(M_{0},M_{1}\in \mathrm{Sp}(2n)\), we write \(M_{0}\approx M_{1}\) if they are symplectically similar. For \(M\in \mathrm{Sp}(2)\), it is symplectically similar to one of the following normal forms:

$$\begin{array}[]{c}D(\lambda)=\left(\begin{array}[]{cc}\lambda&0\\ 0&\lambda^{-1}\end{array}\right),\quad\lambda\in\mathbb{R}\setminus\{0\}\\ N_{1}(\lambda,a)=\left(\begin{array}[]{cc}\lambda&a\\ 0&\lambda\end{array}\right),\quad\lambda=\pm 1,a=\pm 1,0\\ R(\theta)=\left(\begin{array}[]{cc}\cos\theta&-\sin\theta\\ \sin\theta&\cos\theta\end{array}\right),\quad\theta\in(0,\pi)\cup(\pi,2\pi).\end{array}$$
For \(\gamma\in\mathcal{P}_{\tau}(2)\), Y. Long has given a nice picture to explain the index [17]. In this case, we can get the normal form of \(\gamma(\tau)\) by its index \(i_{\omega}(\gamma)\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Hu, X., Ou, Y. & Tang, X. Linear Stability of an Elliptic Relative Equilibrium in the Spatial \(n\)-Body Problem via Index Theory. Regul. Chaot. Dyn. 28, 731–755 (2023). https://doi.org/10.1134/S1560354723040135

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1134/S1560354723040135

Keywords

Navigation