This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Bifurcation of periodic orbits for the NN-body problem, from a non geometrical family of solutions.

Oscar Perdomo1, Andrés Rivera2, Johann Suárez3 1 Department of Mathematics. Central Connecticut State University. New Britain. CT 06050 perdomoosm@cssu.edu 2 Departamento de Ciencias Naturales y Matemáticas. Pontificia Universidad
Javeriana Cali, Facultad de Ingeniería. Calle 18 No. 118–250 Cali, Colombia
amrivera@javerianacali.edu.co 3 Departamento de Matemáticas Universidad del Valle, Facultad de Ciencias. Cali, Colombia johann.j.suarez@correounivalle.edu.co
(Date: August 10, 2025)
Abstract.

Given two positive real numbers MM and mm and an integer n>2n>2, it is well known that we can find a family of solutions of the (n+1)(n+1)-body problem where the body with mass MM stays put at the origin and the other nn bodies, all with the same mass mm, move on the xx-yy plane following ellipses with eccentricity ee. It is expected that this geometrical family that depends on ee, has some bifurcations that produces solutions where the body in the center moves on the zz-axis instead of staying put in the origin. By doing analytic continuation of a periodic numerical solution of the 44-body problem –the one displayed on the video http://youtu.be/2Wpv6vpOxXk –we surprisingly discovered that the origin of this periodic solution is not part of the geometrical family of elliptical solutions parametrized by the eccentricity ee. It comes from a not so geometrical but easier to describe family. Having notice this new family, the authors find an exact formula for the bifurcation point in this new family and use it to show the existence of non planar periodic solution for any pair of masses MM, mm and any integer nn. As a particular example we find a solution where three bodies with mass 33 move around a body with mass 77 that moves up and down.

Key words and phrases:
NN-body problem, periodic orbits, bifurcations, analytic continuation method.
2010 Mathematics Subject Classification:
70F10, 37C27, 34A12.

1. Introduction

The spatial isosceles solutions of the three body problem are solutions where two bodies of equal mass mm have initial positions and velocities symmetric with respect to the zz-axis which passes through their center of mass at the origin, the third body of mass MM moves on the zz-axis, and at any time the three bodies form an isosceles triangle. The youtube video https://www.youtube.com/watch?v=fSmQyeKcj5k shows some images of a periodic spatial solution. Due to the importance and several applications that the spatial three body configuration offers, nowadays there is a large numbers of papers devoted to this problem, see for example [1, 2, 10, 12, 13, 15, 17, 18] and specially [7] and the references therein. In all cited papers, the existence of periodic spatial isosceles solutions has been considered using variational methods, numerical methods or techniques of analytical continuation. In most of the cases it is assumed that the body on the zz-axis has a small mass MM or a mass M=0M=0, see [1, 3, 10, 15]. In the latter case we obtain the Sitnikov problem. For example in [1] using the analytical continuation method of Poincaré, the authors prove analitically the existence of periodic and quasi-periodic spatial isosceles solutions, for M>0M>0 sufficiently small as a continuation of some well known periodic solutions of the reduced circular Sitnikov problem (when the body in the zz-axis has infinitesimal mass M=0M=0 and the other bodies move in circular orbits of the two-body problem). Using variational methods and for M>0M>0 non necessarily small, the authors in [18] prove numerically the existence of spatial isosceles solutions as minimizers of the corresponding Lagragian action functional. A natural generalization of the spatial isosceles solutions are those where nn bodies move on the vertex of a regular polygon perpendicular to the zz-axis and the (n+1)(n+1)-th body moves along the zz-axis. An example of this type of solution can be seen in the youtube video https://www.youtube.com/watch?v=PtEMb6Rvflg.

In order to describe these motions more precisely, let us define

[2π/n]=(cos(2π/n)sin(2π/n)0cos(2π/n)sin(2π/n)0001).\mathcal{R}[2\pi/n]=\begin{pmatrix}\cos(2\pi/n)&-\sin(2\pi/n)&0\\ \cos(2\pi/n)&\sin(2\pi/n)&0\\ 0&0&1\\ \end{pmatrix}.

A direct computation (also see [11]), shows that

Theorem 1.1.

The functions

x(t)=(0,0,f(t)),yk(t)=k[2(k1)π/n]y1(t),k=1,,n,x(t)=\left(0,0,f(t)\right),\quad y^{k}(t)=\mathcal{R}^{k}[2(k-1)\pi/n]y^{1}(t),\quad k=1,\dots,n,

with

y1(t)=(r(t)cos(θ(t)),r(t)sin(θ(t)),Mmnf(t)),y^{1}(t)=\left(r(t)\cos(\theta(t))\,,\,r(t)\sin(\theta(t)),\,-\frac{M}{mn}\,f(t)\right),

satisfying

f(0)=0,f˙(0)=b,r(0)=r0,r˙(0)=0,θ(0)=0,θ˙(0)=ar02,f(0)=0,\quad\dot{f}(0)=b,\quad r(0)=r_{0},\quad\dot{r}(0)=0,\quad\theta(0)=0,\quad\dot{\theta}(0)=\frac{a}{r^{2}_{0}},

provides a solution of the (n+1)(n+1)-body problem with n2n\geq 2 and masses M0M\geq 0 for the body moving along x(t)x(t) and mass m>0m>0 for each body moving along the function yk(t)y^{k}(t), if and only if

(1.1) f¨=(M+mn)fh3,r¨=r02a2r3mλnr2Mrh3,θ˙=r0ar2,\begin{split}\ddot{f}&=-\frac{(M+mn)f}{h^{3}},\\ \ddot{r}&=\frac{r_{0}^{2}a^{2}}{r^{3}}-\frac{m\lambda_{n}}{r^{2}}-\frac{Mr}{h^{3}},\\ \dot{\theta}&=\frac{r_{0}a}{r^{2}},\end{split}

where h=r2+(M+nmmn)2f2\displaystyle{h=\sqrt{r^{2}+\Big{(}\frac{M+nm}{mn}\Big{)}^{2}f^{2}}} and λn=k=1n11ei2πkn|ei2πkn1|3=14k=1n1csc(2πkn).\displaystyle{\lambda_{n}=\sum_{k=1}^{n-1}\frac{1-\hbox{e}^{i\frac{2\pi k}{n}}}{|\hbox{e}^{i\frac{2\pi k}{n}}-1|^{3}}=\frac{1}{4}\sum_{k=1}^{n-1}\csc\Big{(}\frac{2\pi k}{n}\Big{)}.}

Related to the system (1.1) we mention some important results obtained in [11, 12, 13, 14] by the first author of this paper. Firstly, a reduction of the system (1.1) to a single second order differential equation, see [11]. Secondly, a rigorous mathematical proof of the periodicity of a solution of the 3-body problem describe by (1.1), see [12]. This was achieved through a numerical method that keeps track of the round-off error, developed in [14] and also by a lemma that can be viewed as a numerical version of the implicit function theorem, see [12]. Finally, in [13] the author describes a new family of symmetric periodic solutions of (1.1) that contains a nontrivial bifurcation point. The diagram of periodic solutions in the space of initial conditions (similar to the one in Figure 4.2) shows four branches emanating from this bifurcation solution. One is unbounded, another has as a limit point a solution where there is collision of two bodies, the third branch has as a limit point a solution where there is collision of the three bodies and the fourth one ends on a family of trivial solutions.

The main purpose of this paper is to provide an analysis of the properties of the family of periodic solutions in the fourth branch mentioned above and to provide an explicit formula for its limit point. To this end, assuming nonzero angular momentum for (1.1) we will consider an appropriate system which symmetries can be used to obtain the periodic solutions in the fourth branch; see Section 2. In Section 3, we prove Theorems 3.1 and 3.2 which are the statements of our main results. It is interesting to compare our results in this paper with those in [1]. Both papers analytically show the existence of periodic solutions using analytic continuation techniques. The new solutions in [1] emanate from solutions of the Sitnikov problem while those in this paper emanates from what we can call “part of circular solutions” where the two masses mm and MM, are arbitrary. We call them part of circular solutions because nn of the (n+1)(n+1) bodies move on a circle but they do not necessarily cover the whole circle. Finally, numerical validation of the theoretical results obtained in Section 3 are presented in Section 4.

2. The reduced problem and symmetries

A simple inspection of the system (1.1), shows that r2θ˙r^{2}\dot{\theta} is the first integral of the angular momentum C=r0aC=r_{0}a. Having said this, throughout this document we consider only solutions of (1.1) with nonzero angular momentum, i.e., C=r0a0C=r_{0}a\neq 0. This assumption allows us to consider only the initial value problem,

(2.1) f¨=(M+mn)fh3,f(0)=0,f˙(0)=b,r¨=r02a2r3mλnr2Mrh3,r(0)=r0,r˙(0)=0.\begin{split}\ddot{f}&=-\frac{(M+mn)f}{h^{3}},\qquad\qquad\,f(0)=0,\quad\dot{f}(0)=b,\\ \ddot{r}&=\frac{r_{0}^{2}a^{2}}{r^{3}}-\frac{m\lambda_{n}}{r^{2}}-\frac{Mr}{h^{3}},\quad\quad r(0)=r_{0},\quad\dot{r}(0)=0.\end{split}

We can check that, if ϕ(t)=(f(t),r(t))\phi(t)=(f(t),r(t)) is a solution of (2.1) and

θ(t)=0tr0ar2(s)𝑑s,\theta(t)=\int_{0}^{t}\frac{r_{0}a}{r^{2}(s)}ds,

then ϕ¯(t)=(f(t),r(t),θ(t))\overline{\phi}(t)=(f(t),r(t),\theta(t)) is a solution of (1.1). From now on, we denote by

F(a,b,t)=f(t),R(a,b,t)=r(t),Θ(a,b,t)=θ(t),F(a,b,t)=f(t),\quad R(a,b,t)=r(t),\quad\Theta(a,b,t)=\theta(t),

the solutions of the system (1.1) with initial conditions

(2.2) f(0)=0,f˙(0)=b,r(0)=r0,r˙(0)=0,θ(0)=0,θ˙(0)=ar02.f(0)=0,\quad\dot{f}(0)=b,\quad r(0)=r_{0},\quad\dot{r}(0)=0,\quad\theta(0)=0,\quad\dot{\theta}(0)=\frac{a}{r^{2}_{0}}.
Remark 2.1.

If for some aa and bb, f(t)f(t) and r(t)r(t) are TT-periodic of the system (2.1) then (f(t),r(t),θ(t))(f(t),r(t),\theta(t)) defines a periodic solution of the (n+1)(n+1)-th body, if and only if θ(T)\theta(T) is equal to n1π/n2n_{1}\pi/n_{2} with n1n_{1} and n2n_{2} whole numbers. See [13]. In general, TT-periodic solutions of the systems (2.1) define reduced-periodic solutions of the (n+1)(n+1)-body problem. This is, solutions with the property that every TT unites of time, the positions and velocities of the (n+1)(n+1) bodies only differ by an rigid motion in 3.\mathbb{R}^{3}.

The existence of periodic solutions of (1.1) becomes simpler if we restrict our seach to periodic solutions with symmetries. In such a case, the following lemma, see [13] provides a useful result.

Lemma 2.2.

Let ϕ(t)=(F(a,b,t),R(a,b,t))\phi(t)=(F(a,b,t),R(a,b,t)) be a solution of (2.1).

  1. \rhd

    If for some 0<T0<T we have

    F(a,b,T)=0andRt(a,b,T)=0,()F(a,b,T)=0\quad\text{and}\quad R_{t}(a,b,T)=0,\qquad(\dagger)

    then f(t)=F(a,b,t)f(t)=F(a,b,t) and r(t)=R(a,b,t)r(t)=R(a,b,t) are both 2T2T-periodic functions.

  2. \rhd

    If for some 0<T0<T we have

    Ft(a,b,T)=0andRt(a,b,T)=0,()F_{t}(a,b,T)=0\quad\text{and}\quad R_{t}(a,b,T)=0,\qquad(\dagger\dagger)

    then f(t)=F(a,b,t)f(t)=F(a,b,t) and r(t)=R(a,b,t)r(t)=R(a,b,t) are both 4T4T-periodic functions.

It is worth to mentioning that the solutions ϕ(t)=(F(a,b,t),R(a,b,t))\phi(t)=(F(a,b,t),R(a,b,t)) that satisfies ()(\dagger) are called odd solutions because f(t)=F(a,b,t)f(t)=F(a,b,t) is an odd function with respect to t=0t=0. On the other hand, if ϕ(t)\phi(t) satisfies ()(\dagger\dagger) they are called odd/even solutions because f(t)=F(a,b,t)f(t)=F(a,b,t) is an odd function with respect to t=0t=0, but with respect to t=Tt=T, both functions f(t)=F(a,b,t)f(t)=F(a,b,t) and r(t)=R(a,b,t)r(t)=R(a,b,t) are even. Furthermore, we point out that every odd/even solution is also an odd solution.

3. Main results

3.1. Periodic solutions for the reduced problem

In this section we prove the existence of a one parametric a family of periodic solutions for the reduced (n+1)(n+1)-body problem (2.1).

Theorem 3.1.

For any n2n\geq 2, let us define λn=14k=1n1csc(kπ/n)\displaystyle{\lambda_{n}=\frac{1}{4}\sum_{k=1}^{n-1}\csc\big{(}k\pi/n\big{)}}. Assume that m,Mm,M and nn satisfy that λnnp2+Mm(p21)\lambda_{n}\neq np^{2}+\frac{M}{m}(p^{2}-1) for every positive integer pp. Then, for any positive real number r0r_{0} there exist b0b\neq 0 near 0, T>0T>0 near πr03nm+M\pi\sqrt{\frac{r_{0}^{3}}{nm+M}}, and a>0a>0 near λnm+Mr0\sqrt{\frac{\lambda_{n}m+M}{r_{0}}} that provides an odd 2T2T-periodic solution of the reduced (n+1)(n+1)-body problem with initial conditions described in Theorem 1.1.

Proof.

For fixed values of m,M,r0m,M,r_{0}, let F(a,b,T)F(a,b,T), R(a,b,T)R(a,b,T) be the solutions of (2.1)-(2.2) evaluated at t=Tt=T. By Lemma 2.2 it follows that if for some a,ba,b and TT we have that

F(a,b,T)=0andRt(a,b,T)=0,\displaystyle F(a,b,T)=0\quad\hbox{and}\quad R_{t}(a,b,T)=0,

then tf(t)=F(a,b,t)t\longrightarrow f(t)=F(a,b,t) and tr(t)=R(a,b,t)t\longrightarrow r(t)=R(a,b,t) are odd and even functions respectively and both functions are periodic with period 2T2T. An easy computation shows that

F(a,0,T)=0,T.F(a,0,T)=0,\quad\forall T\in\mathbb{R}.

From here we can deduce the following

  • a)

    For all (a,b,T)(a,b,T) we can express FF as

    (3.1) F(a,b,T)=bF~(a,b,T),\displaystyle F(a,b,T)=b\tilde{F}(a,b,T),

    with F~\tilde{F} some smooth function. Therefore, an easy a direct computation shows that F~(a,0,T)=Fb(a,0,T)\displaystyle{\tilde{F}(a,0,T)=F_{b}(a,0,T)}. Moreover, for all (a,T)(a,T) it follows

    F~t(a,0,T)=Fbt(a,0,T),F~a(a,0,T)=Fba(a,0,T),andF~b(a,0,T)=Fbb(a,0,T)/2.\tilde{F}_{t}(a,0,T)=F_{bt}(a,0,T),\quad\tilde{F}_{a}(a,0,T)=F_{ba}(a,0,T),\quad\text{and}\quad\tilde{F}_{b}(a,0,T)=F_{bb}(a,0,T)/2.
  • b)

    If a=a0=λnm+Mr0a=\displaystyle{a_{0}=\sqrt{\frac{\lambda_{n}m+M}{r_{0}}}}, we have that F(a0,0,T)=0F(a_{0},0,T)=0 and R(a0,0,T)=r0R(a_{0},0,T)=r_{0} solves (2.1). Therefore,

    F(α(T))=Rt(α(T))=0,withα(T)=(a0,0,T),T.F(\alpha(T))=R_{t}(\alpha(T))=0,\quad\text{with}\quad\alpha(T)=(a_{0},0,T),\quad T\in\mathbb{R}.

    The path α(T)\alpha(T), more precisely, the pseudo periodic solutions of the (n+1)(n+1)-body problem induced by the equations ()(\dagger), can be viewed geometrically as the pseudo periodic solutions where nn of the (n+1)(n+1)-bodies move along part of a circle and the (n+1)(n+1)-body stays put in the center. From this collection we will find a bifurcation point, a particular value for TT, that will provide the nontrivial periodic solutions.

The previous observations suggest to study the solutions of the system

(3.2) F~(a,b,T)=0andRt(a,b,T)=0.\displaystyle\tilde{F}(a,b,T)=0\quad\hbox{and}\quad R_{t}(a,b,T)=0.

To this end we will use relation (3.1) to compute derivatives of F~\tilde{F} and RtR_{t}. Recall that

(3.3) Ftt\displaystyle F_{tt} =\displaystyle= (M+mn)Fh3,\displaystyle-\frac{(M+mn)F}{h^{3}},
(3.4) Rtt\displaystyle R_{tt} =\displaystyle= a2r02R3mλnR2MRh3,\displaystyle\frac{a^{2}r_{0}^{2}}{R^{3}}-\frac{m\lambda_{n}}{R^{2}}-\frac{MR}{h^{3}},

where h=R2+(M+nmmn)2F2\displaystyle{h=\sqrt{R^{2}+\Big{(}\frac{M+nm}{mn}\Big{)}^{2}F^{2}}}. Taking the partial derivative with respect to bb on both sides of Equation (3.3) and evaluating at α(t)\alpha(t) give us that Fb(α(t))F_{b}(\alpha(t)) satisfies

u¨=(M+mn)r03u,withu(0)=0,u˙(0)=1.\ddot{u}=-\frac{(M+mn)}{r_{0}^{3}}u,\quad\text{with}\quad u(0)=0,\,\,\dot{u}(0)=1.

Therefore,

(3.5) Fb(α(t))=(M+mnr03)1/2sin((M+mnr03)1/2t).\displaystyle F_{b}(\alpha(t))=\left(\frac{M+mn}{r_{0}^{3}}\right)^{-1/2}\sin\left(\left(\frac{M+mn}{r_{0}^{3}}\right)^{1/2}t\right).

The equation above shows that, if

(3.6) T0=πr03mn+M,\displaystyle T_{0}=\pi\sqrt{\frac{r_{0}^{3}}{mn+M}}\,,

then

F~(α(T0))=Fb(α(T0))=0.\tilde{F}(\alpha(T_{0}))=F_{b}(\alpha(T_{0}))=0.

In consequence, from a)a) and b)b) it follows that α(T0)\alpha(T_{0}) satisfies

F~(α(T0))=0,andRt(α(T0))=0,\tilde{F}(\alpha(T_{0}))=0,\quad\hbox{and}\quad R_{t}(\alpha(T_{0}))=0,

showing that the point (a0,0,T0)(a_{0},0,T_{0}) solves (3.2). Finally, from Equation (3.5) we get that

(3.7) F~t(α(T0))=Fbt(α(T0))=1.\displaystyle\tilde{F}_{t}(\alpha(T_{0}))=F_{bt}(\alpha(T_{0}))=-1.

Now, we take the derivative with respect to aa on both sides of Equation (3.4) and evaluate at α(t)\alpha(t). Then Ra(α(t))R_{a}(\alpha(t)) satisfies

v¨=2(λnm+Mr03)1/2(λnm+M)r03v,withv(0)=0,v˙(0)=0,\ddot{v}=2\left(\frac{\lambda_{n}m+M}{r_{0}^{3}}\right)^{1/2}-\frac{(\lambda_{n}m+M)}{r_{0}^{3}}v,\quad\text{with}\quad v(0)=0,\,\,\dot{v}(0)=0,

therefore

Ra(α(t))=2(λnm+Mr03)1/2(1cos((λnm+Mr03)1/2t))R_{a}(\alpha(t))=2\left(\frac{\lambda_{n}m+M}{r_{0}^{3}}\right)^{-1/2}\left(1-\cos\left(\left(\frac{\lambda_{n}m+M}{r_{0}^{3}}\right)^{1/2}t\right)\right)

From this last expression we deduce

Rat(α(t))=2sin((λnm+Mr03)1/2t).R_{at}(\alpha(t))=2\sin\left(\left(\frac{\lambda_{n}m+M}{r_{0}^{3}}\right)^{1/2}t\right).

Applying the same ideas, the function Rb(α(t))\displaystyle{R_{b}(\alpha(t))} can be obtained by solving

w¨(t)=(λnm+M)r03w,withw(0)=0,w˙(0)=0,\ddot{w}(t)=-\frac{(\lambda_{n}m+M)}{r_{0}^{3}}w,\quad\text{with}\quad w(0)=0,\,\,\dot{w}(0)=0,

therefore

Rb(α(t))=Rbt(α(t))=0,R_{b}(\alpha(t))=R_{bt}(\alpha(t))=0,

for all t.t\in\mathbb{R}. Further, from (3.4) we can deduce that Rtt(α(t))=0\displaystyle{R_{tt}(\alpha(t))=0} for all t.t\in\mathbb{R}. These computations shows that the gradient of the function RtR_{t} at α(T0)\alpha(T_{0}) is given by

(3.8) Rt(α(T0))=(2sin(πλnm+Mmn+M),0,0).\displaystyle\nabla R_{t}(\alpha(T_{0}))=\left(2\sin\left(\pi\sqrt{\frac{\lambda_{n}m+M}{mn+M}}\right),0,0\right).

Notice that our condition of λn\lambda_{n} guarantees that Rt(α(T0))\nabla R_{t}(\alpha(T_{0})) does not vanish. On the other hand, using a)a) and the equation (3.7) we obtain

(3.9) F~(α(T0))=(Fba(α(T0)),12Fbb(α(T0)),1).\displaystyle\nabla\tilde{F}(\alpha(T_{0}))=\left(F_{ba}(\alpha(T_{0})),\frac{1}{2}F_{bb}(\alpha(T_{0})),-1\right).

Since

η=F~(α(T0))×Rt(α(T0))=(0,Rta(α(T0)),12Rta(α(T0))Fbb(α(T0))),\eta=\nabla\tilde{F}(\alpha(T_{0}))\times\nabla R_{t}(\alpha(T_{0}))=\left(0,R_{ta}(\alpha(T_{0})),\frac{1}{2}R_{ta}(\alpha(T_{0}))F_{bb}(\alpha(T_{0}))\right),

has its second entry different from zero, by the Implicit Function Theorem there exists ϵ>0\epsilon>0 and a pair of continuous functions such that T,a:(ϵ,ϵ)T,a:(-\epsilon,\epsilon)\to\mathbb{R}

bT=T(b),ba=a(b),b\longrightarrow T=T(b),\quad b\longrightarrow a=a(b),

with T(0)=T0T(0)=T_{0}, a(0)=a0a(0)=a_{0}, such that

F~(γ1(b))=0,andRt(γ1(b))=0,\tilde{F}(\gamma_{1}(b))=0,\quad\text{and}\quad R_{t}(\gamma_{1}(b))=0,

with γ1:(ϵ,ϵ)3,\gamma_{1}:(-\epsilon,\epsilon)\to\mathbb{R}^{3}, given by

(3.10) bγ1(b)=(a(b),b,T(b)),γ1(0)=(a0,0,T0).b\to\gamma_{1}(b)=(a(b),b,T(b)),\quad\quad\gamma_{1}(0)=(a_{0},0,T_{0}).

Therefore, for each b(ϵ,ϵ)b\in(-\epsilon,\epsilon) it follows

F(γ1(b))=bF~(γ1(b))=0,andRt(γ1(b))=0.\begin{split}F(\gamma_{1}(b))=b\tilde{F}(\gamma_{1}(b))&=0,\quad\text{and}\quad R_{t}(\gamma_{1}(b))=0.\end{split}

In consequence, using Lemma 2.2 we get that for any bb, the functions f(t)=F(a,b,t)f(t)=F(a,b,t) y r(t)=R(a,b,t)r(t)=R(a,b,t) define a 2T(b)2T(b)-periodic solution of the reduced problem (2.1).

Theorem 3.2.

For any n2n\geq 2, let us define λn=14k=1n1csc(kπ/n)\displaystyle{\lambda_{n}=\frac{1}{4}\sum_{k=1}^{n-1}\csc\big{(}k\pi/n\big{)}}. Assume that m,Mm,M and nn satisfy that λnnq2+Mm(q21)\lambda_{n}\neq nq^{2}+\frac{M}{m}(q^{2}-1) for every positive even integer qq. Then, for any positive real number r0r_{0} there exist b0b\neq 0 near 0, T>0T_{*}>0 near π2r03nm+M\dfrac{\pi}{2}\sqrt{\frac{r_{0}^{3}}{nm+M}} and a>0a>0 near λnm+Mr0\sqrt{\frac{\lambda_{n}m+M}{r_{0}}} that provides an ood/even 4T4T_{*}-periodic solution of the reduced (n+1)(n+1)-body problem with initial conditions described in Theorem 1.1.

Proof.

We follows the same lines of the proof of Theorem 3.1, but now considering consider the system

(3.11) Ft(a,b,T)=0andRt(a,b,T)=0.F_{t}(a,b,T)=0\quad\text{and}\quad R_{t}(a,b,T)=0.\\

By (3.1) there exists a function W(t,a,b)W(t,a,b) such that

Ft(a,b,t)=bW(a,b,t),Fbt(a,b,t)=W(a,b,t)+bWb(a,b,t),\begin{split}F_{t}(a,b,t)&=bW(a,b,t),\\ F_{bt}(a,b,t)&=W(a,b,t)+bW_{b}(a,b,t),\end{split}

for all (a,b,t).(a,b,t). In particular

Fbt(a,0,t)=W(a,0,t),F_{bt}(a,0,t)=W(a,0,t),

for all (a,t)(a,t). Now we search for points (a,b,T)(a,b,T_{*}) such that

W(a,b,T)=0andRt(a,b,T)=0.W(a,b,T_{*})=0\quad\text{and}\quad R_{t}(a,b,T_{*})=0.

To this end, we need to study the zeroes of the function Fbt(a,b,T)F_{bt}(a,b,T_{*}). From (3.5) we have

Fbt(α(t))=cos((M+mnr03)1/2t),F_{bt}(\alpha(t))=\cos\left(\left(\frac{M+mn}{r_{0}^{3}}\right)^{1/2}t\right),

Therefore, Fbt(α(T0/2))=0F_{bt}(\alpha(T_{0}/2))=0 and,

W(a0,0,T0/2)=0andRt(a0,0,T0/2)=0.W(a_{0},0,T_{0}/2)=0\quad\text{and}\quad R_{t}(a_{0},0,T_{0}/2)=0.

Once again, with the aim of applying the Implicit Function Theorem we consider the gradiente vector W(a0,0,T0/2)\nabla W(a_{0},0,T_{0}/2) and Rt(a0,0,T0/2)\nabla R_{t}(a_{0},0,T_{0}/2). A direct computation shows that

W(α(T0/2))×Rt(α(T0/2))=(0,πT0Rta(α(T0/2)),12Rta(α(T0/2))Fbbt(α(T0/2))).\nabla W(\alpha(T_{0}/2))\times\nabla R_{t}(\alpha(T_{0}/2))=\left(0,-\frac{\pi}{T_{0}}R_{ta}(\alpha(T_{0}/2)),\frac{1}{2}R_{ta}(\alpha(T_{0}/2))F_{bbt}(\alpha(T_{0}/2))\right).

Notice that the second entry of above vector is given by

Rta(α(T0/2))=2sin(π2λnm+Mmn+M),R_{ta}(\alpha(T_{0}/2))=2\sin\left(\frac{\pi}{2}\sqrt{\frac{\lambda_{n}m+M}{mn+M}}\right),

which by hypothesis is different from zero. Then, there exist δ>0\delta>0 and two continuous functions T,a:(δ,δ)T_{*},a_{*}:(-\delta,\delta)\rightarrow\mathbb{R} such that

bT=T(b),ba=a(b),b\rightarrow T_{*}=T_{*}(b),\quad b\rightarrow a_{*}=a_{*}(b),

with T(0)=T0/2T_{*}(0)=T_{0}/2, a(0)=a0a_{*}(0)=a_{0} and

W(γ2(b))=0andRt(γ2(b))=0,W(\gamma_{2}(b))=0\quad\text{and}\quad R_{t}(\gamma_{2}(b))=0,

with γ2:(δ,δ)3,\gamma_{2}:(-\delta,\delta)\to\mathbb{R}^{3},

(3.12) bγ2(b)=(a(b),b,T(b)),γ2(0)=(a0,0,T0/2).b\to\gamma_{2}(b)=(a_{*}(b),b,T_{*}(b)),\quad\quad\gamma_{2}(0)=(a_{0},0,T_{0}/2).

Then for b(δ,δ)b\in(-\delta,\delta) it follows

Ft(γ2(b))=bW(γ2(b))=0,andRt(γ2(b))=0,\begin{split}F_{t}(\gamma_{2}(b))=bW(\gamma_{2}(b))&=0,\quad\text{and}\quad R_{t}(\gamma_{2}(b))=0,\end{split}

Once again, by Lemma 2.2 we have that for any bb, the functions f(t)=F(a,b,t)f(t)=F(a,b,t) y r(t)=R(a,b,t)r(t)=R(a,b,t) define an odd/even 4T(b)4T_{*}(b)-periodic solution of the reduced problem (2.1). ∎

3.2. Periodic solutions for the (n+1)(n+1)-body problem

We will only consider the case when the solutions of the (n+1)(n+1)-body problem comes from solutions of Equation ()(\dagger). The case when the solutions of the (n+1)(n+1)-body problem come from solutions of equation ()(\dagger\dagger) is similar.

In order to show that we can find a non trivial periodic solution with b>0b>0 we just need to check that the function Θ(a,b,t)\Theta(a,b,t) is not constant along the curve γ1(b)=(a(b),b,T(b))\gamma_{1}(b)=(a(b),b,T(b)) defined in Equation (3.10). This is true because every triple (a(b),b,T(b))(a(b),b,T(b)) solves the equation F=0F=0 and Rt=0R_{t}=0 and by continuity, if bΘ(a(b),b,T(b)))b\to\Theta(a(b),b,T(b))) is not constant, then we can find a bb such that γ1(b)\gamma_{1}(b) satisfies that Θ(γ1(b))=n1π/n2\Theta(\gamma_{1}(b))=n_{1}\pi/n_{2} with n1n_{1} and n2n_{2} integers. See Remark 2.1.

One way to study the behavior of the function Θ\Theta along the curve γ1(b)\gamma_{1}(b) is by reparametrizing the curve γ1(b)\gamma_{1}(b) as β(τ)=γ1(b(τ))\beta(\tau)=\gamma_{1}(b(\tau)) where β\beta is an integral curve of the vector field XX defined as

X=F~×Rt=(F~bRttF~tRtb,F~tRtaF~aRtt,F~aRtbF~bRta).X=\nabla\tilde{F}\times\nabla R_{t}=\left(\tilde{F}_{b}R_{tt}-\tilde{F}_{t}R_{tb},\tilde{F}_{t}R_{ta}-\tilde{F}_{a}R_{tt},\tilde{F}_{a}R_{tb}-\tilde{F}_{b}R_{ta}\right).

As we have done before, we can explicitly compute as many partial derivatives of the functions F~\tilde{F} and RtR_{t} at the bifurcation point p0=β(0)=γ(0)p_{0}=\beta(0)=\gamma(0). Therefore we can compute as many derivatives as needed for the curve β\beta at τ=0\tau=0 and since we can also compute as many partial derivatives of the function Θ\Theta at p0p_{0} then using the chain rule we can compute the first and second derivative of the function

ξ(τ)=Θ(β(τ)),\xi(\tau)=\Theta(\beta(\tau)),

at τ=0\tau=0. A direct computation shows that ξ(0)=0\xi^{\prime}(0)=0 and a long direct computation, see [16], shows that

ξ′′(0)\displaystyle\xi^{\prime\prime}(0) =\displaystyle= (A(n,m,M,r0)+B(n,m,M,r0)Rat(α(T0)))Rat2(α(T0)),\displaystyle\big{(}A(n,m,M,r_{0})+B(n,m,M,r_{0})R_{at}(\alpha(T_{0}))\big{)}R^{2}_{at}(\alpha(T_{0})),

with A=A(n,m,M,r0)A=A(n,m,M,r_{0}) and B=B(n,m,M,r0)B=B(n,m,M,r_{0}) given by

A=3r05/2π[9(λnm+M)(λnm+24M4mn+16Mλnm+Mr03)8M(M+mn)(9+8λnm+Mr03)]16m2n2(λnm+4mn+3M)(λnm+M)1/2
B
=9Mr05/2(M+mn)5/2(3+2λnm+Mr03)m2n2(λnm+M)(λnm3M4mn)2.
\begin{split}A&=\frac{3r_{0}^{5/2}\pi\bigg{[}9(\lambda_{n}m+M)\Big{(}\lambda_{n}m+24M-4mn+16M\sqrt{\frac{\lambda_{n}m+M}{r_{0}^{3}}}\Big{)}-8M(M+mn)\Big{(}9+8\sqrt{\frac{\lambda_{n}m+M}{r_{0}^{3}}}\Big{)}\bigg{]}}{16m^{2}n^{2}(-\lambda_{n}m+4mn+3M)(\lambda_{n}m+M)^{1/2}}\\ \vskip 19.91684ptB&=\frac{9Mr_{0}^{5/2}(M+mn)^{5/2}\Big{(}3+2\sqrt{\frac{\lambda_{n}m+M}{r_{0}^{3}}}\Big{)}}{m^{2}n^{2}(\lambda_{n}m+M)(\lambda_{n}m-3M-4mn)^{2}}.\end{split}

Which implies that for most of the choices of nn, MM and mm the second derivative of ξ(τ)=Θ(β(τ))\xi(\tau)=\Theta(\beta(\tau)) at τ=0\tau=0 is different from zero.

4. Numerical solutions

In this section we present the analytic continuation of a periodic solution that is close to the solution displayed in the youtube video http://youtu.be/2Wpv6vpOxXk. It can be checked that m=92m=92, M=242M=242, n=3n=3, r0=11r_{0}=11 and

q0=(a0,b0,T0)=(1.84153,3.79392,7.31715),q_{0}=(a_{0},b_{0},T_{0})=(1.84153,3.79392,7.31715),

provides a periodic solution. Figure 4.1 shows the trajectory of one of the three bodies with mass 92.

Refer to caption
Figure 4.1. Trajectory of one of the bodies with mass 92 for the solution of the 4 body problem with initial conditions q0=(a0,b0,T0)=(1.84153,3.79392,7.31715)q_{0}=(a_{0},b_{0},T_{0})=(1.84153,3.79392,7.31715)

When we extend this solution, similar to the results in [13], the collection of points (a,b,T)(a,b,T) that satisfy the equations F(a,b,T)=0F(a,b,T)=0 and Rt(a,b,T)=0R_{t}(a,b,T)=0 with b0b\neq 0 have the shape of a fork, where two of the edges goes to a points with a=0a=0 corresponding to periodic solutions with collisions, the other edge seems to be unbounded and the fourth edge goes to a point where b=0b=0. The value of aa for this limit point with b=0b=0 is a0=λ3m+Mr0=22+921135.17965a_{0}=\sqrt{\frac{\lambda_{3}m+M}{r_{0}}}=\sqrt{22+\frac{92}{11\sqrt{3}}}\approx 5.17965 which was somehow expected because this is the aa corresponding to the solution where the mass in the center stays put and the other masses move along a circle. The value of TT for this limit point is near 5.03224. In other words the limit point with b=0b=0 that we found by analytic continuation of the periodic solution displayed in the video is

q1=(a1,b1,T1)=(5.17965, 0, 5.03224).q_{1}=(a_{1},b_{1},T_{1})=(5.17965,\,0,\,5.03224).

Figure 4.2 shows the points q0q_{0} and q1q_{1} as part of the family of reduced periodic solutions of the 4-body problem. This paper comes from an effort to understand the reason of the value of TT in the point above. Initially we were expecting a value for TT equal to 1133923+726π6.6717911\sqrt{\frac{33}{92\sqrt{3}+726}}\pi\approx 6.67179 which is half of the period of the circular solution where the mass in the center stays still. The result of this explanation is Theorem 3.1, it explain how actually this limit point is a bifurcation of the solutions coming from the line

L={(a0,0,t)=(22+92113,0,t):t},L=\Big{\{}(a_{0},0,t)=(\sqrt{22+\frac{92}{11\sqrt{3}}},0,t):t\in\mathbb{R}\Big{\}},

which even though they do not provide geometric periodic solutions, they satisfy the equation F(a,b,t)=0F(a,b,t)=0 and Rt(a,b,t)=0R_{t}(a,b,t)=0. Our main theorem predicts the value of bifurcation for TT at 1111518π5.0358611\sqrt{\frac{11}{518}}\pi\approx 5.03586 which completely explains our numerically value of TT in the point q1q_{1}. Figure 4.2 shows the solutions with b0b\neq 0 that form a fork shape curve and the solution with b=0b=0, the line LL defined above. It is worth pointing out that this line LL is the image of the curve α\alpha that we used in the proof of the main results of this paper.

Refer to caption
Figure 4.2. Two bifurcation points for the spatial 4-body problem. For these solutions the masses of three of the bodies is 92 and the mass of the fourth body is 242. One of the bifurcations is explained in the paper [13], and the other bifurcation, the one that takes place on the line LL, is explained in this paper.

Just to give an application to our Theorem 3.1, we decided to consider periodic solutions of the 4-body problem produced when m=3m=3, M=7M=7 and r0=11r_{0}=11. According to Theorem 3.1, when we consider the system

(4.1) {F(a,b,T)=0,Rt(a,b,T)=0,\begin{cases}\begin{split}F(a,b,T)&=0,\\ R_{t}(a,b,T)&=0,\end{split}\end{cases}

it follows that, along the line of solutions of this system

{(a0,b,T):a0=111(3+7),b=0},\Big{\{}(a_{0},b,T):a_{0}=\sqrt{\frac{1}{11}\left(\sqrt{3}+7\right)},\,b=0\Big{\}},

there is a bifurcation at

p0=(a0,0,T0)=(111(3+7),0,1111π4)(0.890967,0,28.6536).p_{0}=(a_{0},0,T_{0})=\Big{(}\sqrt{\frac{1}{11}\left(\sqrt{3}+7\right)},0,\frac{11\sqrt{11}\pi}{4}\Big{)}\approx(0.890967,0,28.6536).

In order to find a non trivial solution near p0p_{0}, we took b=0.05b=0.05 and we did a search by doing small changes of a0=0.890967a_{0}=0.890967 and T0=28.6536T_{0}=28.6536 to solve the system (4.1). We found that the point

p1=(a1,b1,T1)=(0.8892815,0.05,28.708),p_{1}=(a_{1},b_{1},T_{1})=(0.8892815,0.05,28.708),

Satisfies that |F(a1,b1,T1)||F(a_{1},b_{1},T_{1})| and |Rt(a1,b1,T1)||R_{t}(a_{1},b_{1},T_{1})| are smaller than 10610^{-6}. Using the point p1p_{1} we did an analytic continuation to obtain solutions of the system (4.1) with b0b\neq 0. Figure 4.3 shows about 12,000 solutions found numerically, along with the solutions along the line b=0b=0.

Refer to caption
Figure 4.3. Solutions of the system (4.1) near the bifurcation point p0p_{0}. The point p1p_{1} allowed us to start the analytic continuation by moving along the integral curve of the vector field F×Rt\nabla F\times\nabla R_{t}. This figure shows about 12,000 points with b0b\neq 0. The points p1p_{1}, p2p_{2} and p3p_{3} are points where Θ\Theta equals to 3π4\frac{3\pi}{4}, 4π5\frac{4\pi}{5}, and π\pi respectively.

We can compute the exact value for Θ(p0)\Theta(p_{0}), it is a0T011=π7+342.32086\frac{a_{0}T_{0}}{11}=\frac{\pi\sqrt{7+\sqrt{3}}}{4}\approx 2.32086. We also have that Θ(p1)2.32327\Theta(p_{1})\approx 2.32327, this time the computation has to be done numerically.

Refer to caption
Figure 4.4. This is the graph of Θ\Theta against TT along the solution of the system F(a,b,T)=0,Rt(a,b,T)=0F(a,b,T)=0,\,R_{t}(a,b,T)=0. We have highlighted three points that represents periodic solutions.

Since the function Θ\Theta along the solutions of the system (4.1) changes, by continuity we have that for some solution p=(a,b,T)p=(a,b,T) of the system Θ(p)=uv2π\Theta(p)=\frac{u}{v}2\pi with uu and vv whole numbers. Therefore we get that the functions

x(t)\displaystyle x(t) =\displaystyle= (0,0,F(a,b,t))\displaystyle\left(0,0,F(a,b,t)\right)
y1(t)\displaystyle y^{1}(t) =\displaystyle= (R(a,b,t)cos(Θ(a,b,t)),R(t)sin(Θ(a,b,t)),119F(a,b,t))\displaystyle\left(R(a,b,t)\cos(\Theta(a,b,t))\,,\,R(t)\sin(\Theta(a,b,t)),\,-\frac{11}{9}\,F(a,b,t)\right)
y2(t)\displaystyle y^{2}(t) =\displaystyle= (R(a,b,t)cos(Θ(a,b,t)+2π3),R(t)sin(Θ(a,b,t)+2π3),119F(a,b,t))\displaystyle\left(R(a,b,t)\cos(\Theta(a,b,t)+\frac{2\pi}{3})\,,\,R(t)\sin(\Theta(a,b,t)+\frac{2\pi}{3}),\,-\frac{11}{9}\,F(a,b,t)\right)
y3(t)\displaystyle y^{3}(t) =\displaystyle= (R(a,b,t)cos(Θ(a,b,t)+4π3),R(t)sin(Θ(a,b,t)+4π3),119F(a,b,t)),\displaystyle\left(R(a,b,t)\cos(\Theta(a,b,t)+\frac{4\pi}{3})\,,\,R(t)\sin(\Theta(a,b,t)+\frac{4\pi}{3}),\,-\frac{11}{9}\,F(a,b,t)\right),

will eventually close and therefore the solution of the 44-body problem given by these functions will be, not only pseudo periodic, but periodic. We have noticed that along the branch of solutions of the system (4.1) with b0b\neq 0 (see Figure 4.3), Θ\Theta is a function of TT, see Figure 4.4. After noticing that the values of Θ\Theta increase up to numbers bigger that π\pi, we searched for the solutions that satisfy θ(T)=3π/4\theta(T)=3\pi/4, θ(T)=4π/5\theta(T)=4\pi/5 and θ(T)=π\theta(T)=\pi. We found that the points

p2\displaystyle p_{2} =\displaystyle= (a,b,T)=(0.866953,0.187583,29.4405),\displaystyle(a,b,T)=(0.866953,0.187583,29.4405),
p3\displaystyle p_{3} =\displaystyle= (a,b,T)=(0.775642,0.400635,32.6636),\displaystyle(a,b,T)=(0.775642,0.400635,32.6636),
p4\displaystyle p_{4} =\displaystyle= (a,b,T)=(0.547954,0.634946,41.1787),\displaystyle(a,b,T)=(0.547954,0.634946,41.1787),

satisfy that Θ(p1)=3π4\Theta(p_{1})=\frac{3\pi}{4}, Θ(p3)=4π5\Theta(p_{3})=\frac{4\pi}{5}, and Θ(p4)=π\Theta(p_{4})=\pi. Figure 4.5 shows the image of the function y1(t)y^{1}(t) for these three solutions p1p_{1}, p2p_{2} and p3p_{3}.

Refer to caption
Refer to caption
Refer to caption
Figure 4.5. This figure shows the graph of the function y1(t)y^{1}(t) for the periodic solutions provided by the points p2p_{2}, p3p_{3} and p4p_{4} respectively.

For the periodic solution given by the solution p4p_{4}, Figure 4.6 shows the image of the functions x(t)x(t), y1(t)y^{1}(t), y2(t)y^{2}(t) and y3(t)y^{3}(t).

Refer to caption
Figure 4.6. For this periodic solution m=3m=3, M=7M=7, r0=11r_{0}=11 and n=3n=3. This solutions was found by first finding the bifurcation point given by Theorem 3.1 and then finding a solution with b0b\neq 0 that is nearby, and doing analytic continuation of this latter point. For this solution θ(T)=π\theta(T)=\pi.

References

  • [1] Corbera, M. & LLibre, J. (2000) Families of periodic orbits for the spatial isosceles 33-body problem. Siam J. Math. Anal. 35. No 5, 1311-1346.
  • [2] Devaney, R. Triple collision in the planar isosceles three-body problem. Inventions Math 60 (1980) pp. 249-267
  • [3] Galán, J., Núñez D. & Rivera, A. (2018) Quantitative stability of certain periodic solutions in the Sitnikov problem. Siam J. Applied Dynamical Systems. 17 No 1, 52-77.
  • [4] Hénon, H. A family of periodic solutions of the planar three-body problem, and their stability. Celestial Mechanics 13 (1976) pp. 267-285
  • [5] Martinez, R. & Simo, C. [1987] Qualitative study of the planar isosceles three-body problem. Celestial Mechanics 41, (1987), pp 179-251.
  • [6] McGehee, R. [1974] Triple collision in the collinear three-body problem. Inventions Math. 27, (1974), pp 191-227.
  • [7] Meyer, K. & Wang, Q. [1995] The global phase structure of the three-dimensional isosceles three-body problem with zero energy, in Hamiltonian Dynamical Systems (Cincinnati, OH, 1992). IMA Vol. Math. Appl. 63, Springer-Verlag, New York, 1995, pp 265-282.
  • [8] Meyer, K. & Schmidt, D [1993] Libration of central configurations and braided saturn rings. Celestial Mechanics and Dynamical Astronomy 55: 289-303.
  • [9] Offin, D. & Cabral, H. Hyperbolicity for symmetric periodic orbits in the isosceles three body problem. Discrete and continuous dynamical systems series S. 2 No 2 (2009), 379-392
  • [10] Ortega, R. & Rivera, A. Global bifurcation from the center of mass in the sitnikov problem. Discrete and continuous dynamical systems series B. 14 No 2 (2010), 719-732.
  • [11] Perdomo [2014] A family of solution of the n body problem. http://arxiv.org/pdf/1507.01100.pdf
  • [12] Perdomo A small variation of the Taylor method and periodic solution of the 3-body problem. http://arxiv.org/pdf/1410.1757.pdf
  • [13] Perdomo Bifurcation in the Family of Periodic Orbits for the Spatial Isosceles 3 Body Problem Qual. Theory Dyn. Syst. (2017). https://doi.org/10.1007/s12346-017-0244-1.
  • [14] Perdomo (2019) The round Taylor Method The American Mathematical Monthly (2019). 126:3, 237-251, DOI: 10.1080/00029890.2019.1546087
  • [15] Rivera, A. (2013)Periodic solutions in the generalized Sitnikov (N+1)(N+1)-body problem. Siam J. Applied Dynamical Systems. 12 No 3, 1515-1540.
  • [16] Suárez, Johann. (2020) Sobre la existencia de soluciones periódicas en el problema de NN-cuerpos polygonal. Ph.D. Thesis. Universidad del Valle, Cali-Colombia. 2020.
  • [17] Xia, Z. The Existence of Noncollision Singularities in Newtonian Systems. Annals of Mathematics (1992), 135, 411-468.
  • [18] Yan, D., Ouyang, T. New phenomenons in the spatial isosceles three-body problem. Internet. J. Bifur. Chaos Appl. Sci. Engrg. 25, (2015), 9, 15p