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

Periodic oscillations in a 2N2N-body problem

Oscar Perdomo 1, Andrés Rivera 2, John A. Arredondo 3, Nelson Castañeda 4 1 Departament of Mathematics. Central Connecticut State University. New Britain. CT 06050. Connecticut, USA. 2 Departamento de Ciencias Naturales y Matemáticas. Pontificia Universidad Javeriana Cali, Facultad de Ingeniería y Ciencias, Calle 18 No. 118–250 Cali, Colombia. 3 Departamento de Matemáticas Fundación Universitaria Konrad Lorenz, Facultad de Ciencias e Ingeniería, Cra 9 bis 62-43, Bogotá, Colombia. 4 Departament of Mathematics. Central Connecticut State University. New Britain. CT 06050. Connecticut, USA. [email protected], [email protected], [email protected], [email protected]
Abstract.

Hip-Hop solutions of the 2N2N-body problem are solutions that satisfy at every instance of time, that the 2N2N bodies with the same mass mm, are at the vertices of two regular NN-gons, each one of these NN-gons are at planes that are equidistant from a fixed plane Π0\Pi_{0} forming an antiprism. In this paper, we first prove that for every NN and every mm there exists a family of periodic hip-hop solutions. For every solution in these families the oriented distance to the plane Π0\Pi_{0}, which we call d(t)d(t), is an odd function that is also even with respect to t=Tt=T for some T>0.T>0. For this reason we call solutions in these families, double symmetric solutions. By exploring more carefully our initial set of periodic solutions, we numerically show that some of the branches stablished in our existence theorem have bifurcations that produce branches of solutions with the property that the oriented distance function d(t)d(t) is not even with respect to any T>0T>0, we call these solutions single symmetry solutions. We prove that no single symmetry solution is a choreography. We also display explicit double symmetric solutions that are choreographies.

Key words and phrases:
N-body problem, periodic orbits, hip-hop solutions, choreographies, bifurcations.
2010 Mathematics Subject Classification:
70F10, 37C27, 34A12.

 

 

[Uncaptioned image]

1. Introduction

Hip-hop solutions of the 2N2N-body problem are solutions satisfying that: (i) The masses of all the bodies are the same. (ii) At every instance of time tt, NN of the bodies are at the vertices of a regular NN-gon contained in a plane Π1(t)\Pi_{1}(t) and the other NN bodies are at the vertices of a second regular NN-gon that differ from the first NN-gone by a translation and a rotation of 2πN\frac{2\pi}{N} radians and it is contained in a plane Π2(t)\Pi_{2}(t). (iii) For every tt, the planes Π1(t)\Pi_{1}(t) and Π2(t)\Pi_{2}(t) are parallel and they are equidistant from a fixed plane Π0\Pi_{0}. (iv) The center of the NN-gones are always in a fixed line l0l_{0} perpendicular to Π0\Pi_{0}. In particular when Π1(t)Π2(t)\Pi_{1}(t)\neq\Pi_{2}(t) the 2N2N bodies form an antiprism, and when Π1(t)=Π2(t)\Pi_{1}(t)=\Pi_{2}(t) the 2N2N bodies form a regular 2N2N-gon. For any hip-hop solution we can define the function d(t)d(t) that gives the oriented distance from the plane Π0\Pi_{0} to the plane Π1\Pi_{1} and the function r(t)r(t) that gives the distance from the line l0l_{0} to any of the bodies. We will assume that the line l0l_{0} is the zz-axis and the plane Π0\Pi_{0} is the xx-yy plane.

Refer to caption

Figure 1. Six bodies moving at the vertices of a regular antiprism all time. The function r(t)r(t) is the distance of each body respect to the zz-axis passing through the center of the two parallel planes where each group of three bodies moves. The function d(t)d(t) is the distance of the plane Π1\Pi_{1} to the plane Π0={(x,y,z):z=0}\Pi_{0}=\left\{(x,y,z):z=0\right\}.

We will call a hip-hop solution, double symmetric, if the function d(t)d(t) satisfies that for all tt, d(t)=d(t)d(-t)=-d(t) and d(Tt)=d(T+t)d(T-t)=d(T+t) for some nonzero TT. We will call a hip-hop solution, single symmetric, if the function d(t)d(t) satisfies that d(t)=d(t)d(-t)=-d(t) for all tt and there is not a TT such that, d(Tt)=d(T+t)d(T-t)=d(T+t) for all tt.

The differential equations for hip-hop solutions were provided by Meyer and Schmidt in 1993 [MR1205666] as a model for braided rings of a planet based on a previous model for the rings of Saturn. They called these solutions alternating solutions and the only difference with the hip-hop solutons is that the alternating solutions consider an additional motionless body in the center of mass of the 2N2N bodies. For the alternating solutions, the body in the center makes the role of the planet while the 2N2N bodies are the small bodies conforming the ring. One of the pioneer works that showed the periodicity of some hip-hop solutions is the work of Chenciner and Venturelli [MR1820355], where the authors used variational methods to prove existence of periodic solutions. Some other papers that show the periodicity of hip-hop solutions by variational methods are [MR1820355, MR2401905, MR2031430]. It is well known that the periodic solutions of the two body problem can be extended to periodic planar solutions of the nn body problem, with the trajectory of each body being a rotation of a single ellipse. The link \urlhttps://youtu.be/RjlZpqDFsDM leads to a video showing some of these periodic solutions. Using a Bolzano argument, in [MR2570295] the authors show the periodicity of hip-hop solution with motion close to ellipses with high eccentricity. In [MR2267950] the authors use a Poincaré analytical continuation method to show the existence of double symmetric hip-hop solutions with trajectories close to circles.

Similar to the work in [MR2267950] our solutions emerge from planar periodic solutions where all the bodies move in the same circle. We study both, doubly and single symmetric solutions. The technique that we use is slightly different from the regular Poincaré analytical method. We can say that it is more direct. Let us give the idea behind the method. We consider the functions r(t)r(t) and d(t)d(t) defined above as functions of one of the initial conditions and the angular momentum, more precisely we consider R(a,b,t)=r(t)R(a,b,t)=r(t) and D(a,b,t)=d(t)D(a,b,t)=d(t) where b=d˙(0)b=\dot{d}(0) and aa is the angular momentum. All our solutions satisfy d(0)=0d(0)=0. For the circular solutions b=0b=0 because the solution are planar, d(t)0d(t)\equiv 0, and the value of a=a0a=a_{0} can be easily found from the condition that r(t)r(t) is constant. In the Euclidean space (a,b,T)(a,b,T) with TT a fixed value of time, we have that due to the symmetries of the ordinary differential equations that govern the hip-hop solutions, we have that points in the space that satisfy the system of equations Dt(a,b,T)=Rt(a,b,T)=0D_{t}(a,b,T)=R_{t}(a,b,T)=0 produce double symmetric hip-hop solutions with angular momentum aa and period 4T4T, while points (a,b,T)(a,b,T) that satisfy D(a,b,T)=Rt(a,b,T)=0D(a,b,T)=R_{t}(a,b,T)=0 produce hip-hop solutions with angular momentum aa and period 2T2T that are potentially single symmetric. With this idea in mind we have that double and single symmetric solutions are obtained from studying three surfaces: The surface in the space (a,b,T)(a,b,T) that satisfies Rt(a,b,T)=0R_{t}(a,b,T)=0, the surface in the space (a,b,T)(a,b,T) that satisfies Dt(a,b,T)=0D_{t}(a,b,T)=0 and the surface in the space (a,b,T)(a,b,T) that satisfies D(a,b,T)=0D(a,b,T)=0. Recall that the intersection of two surfaces is in general a curve and that if the gradient of the functions that define two surfaces are linearly dependent at a particular point in the intersection, then the solution of the system may be the union of two curves. As an example, Figure 5 show points (a,b,T)(a,b,T) that solve the system D(a,b,T)=Rt(a,b,T)=0D(a,b,T)=R_{t}(a,b,T)=0. In this paper we notice that the circular solutions of the 2N2N body problem produces a line Γ={(a0,0,T)3}\Gamma=\{(a_{0},0,T)\in\mathbb{R}^{3}\} in the space (a,b,T)(a,b,T) that solve the systems D(a,b,T)=Rt(a,b,T)=0D(a,b,T)=R_{t}(a,b,T)=0 and Dt(a,b,T)=Rt(a,b,T)=0D_{t}(a,b,T)=R_{t}(a,b,T)=0. By computing the gradient of the functions DD, DtD_{t} and RtR_{t} we find points in Γ\Gamma that potentially allows nearby points that solve both systems. Half of these (bifurcation) points were found in the paper [MR2267950] using the Poincare analytical method. The other half that we found were those bifurcation points that potentially may produce single symmetric solutions. It turned out that points in the smooth curves Ωi\Omega_{i} emanating from the new bifurcation points, did not produced single symmetric hip-hop solutions near the circular solutions but, after doing analytic continuation, we found out that for some values of NN, Ωi\Omega_{i} has a bifurcation point that generates a curve of points Φ\Phi in the space (a,b,T)(a,b,T) with the property that each of its point represents a single symmetric solution. We proved that an interesting difference between the single and double symmetric families of hip-hop solutions of the 2N2N body problem is the fact that no single symmetric solution can produce a choreographic while there are infinitely many choreographies in the family of double symmetric hip-hop solutions.

2. Preliminary results

The anti-prism 2N2N-body problem will be characterized in the following concrete framework: Consider Q1,Q2,,Q2NQ_{1},Q_{2},\ldots,Q_{2N} bodies of equal mass m>0m>0, located on the vertices of a regular anti-prism. If 𝐫j(t)\mathbf{r}_{j}(t) is the position of the body Qj,j=1,,2NQ_{j},j=1,\ldots,2N at each instant tt and satisfy

𝐫j(t)=j1𝐫1(t),j=1,,2N,\mathbf{r}_{j}(t)=\mathcal{R}^{j-1}\mathbf{r}_{1}(t),\quad j=1,\dots,2N,

where \mathcal{R} is a rotation/reflection matrix given by

=(cos(πN)sin(πN)0sin(πN)cos(πN)0001).\mathcal{R}=\begin{pmatrix}\cos(\frac{\pi}{N})&-\sin(\frac{\pi}{N})&0\\ \sin(\frac{\pi}{N})&\cos(\frac{\pi}{N})&0\\ 0&0&-1\end{pmatrix}.

Introducing cylindrical coordinates (r,θ,d)(r,\theta,d) it is shown that the equations of motions of the bodies are given by

{r¨=a2r32rmf(r,d),d¨=md2g(r,d),θ˙=a/r2,\begin{cases}\begin{split}\ddot{r}&=\frac{a^{2}}{r^{3}}-2rmf(r,d),\\ \ddot{d}&=-\frac{m\,d}{2}g(r,d),\\ \dot{\theta}&=a/r^{2},\end{split}\end{cases} (1)

where a\displaystyle{a} is the angular momentum of the system and

f(r,d)=k=12N1sin2(kπ/2N)[4r2sin2(kπ/2N)+((1)k1)2d2]3/2,g(r,d)=k=12N1((1)k1)2[4r2sin2(kπ/2N)+((1)k1)2d2]3/2.\begin{split}f(r,d)&=\sum_{k=1}^{2N-1}\frac{\sin^{2}(k\pi/2N)}{\left[4r^{2}\sin^{2}(k\pi/2N)+((-1)^{k}-1)^{2}d^{2}\right]^{3/2}},\\ g(r,d)&=\sum_{k=1}^{2N-1}\frac{((-1)^{k}-1)^{2}}{\left[4r^{2}\sin^{2}(k\pi/2N)+((-1)^{k}-1)^{2}d^{2}\right]^{3/2}}.\end{split}

From now on, for r0,mr_{0},m and NN fixed, we denote by

R(a,b,t)=r(t),D(a,b,t)=d(t),andΘ(a,b,t)=θ(t),R(a,b,t)=r(t),\quad D(a,b,t)=d(t),\quad\text{and}\quad\Theta(a,b,t)=\theta(t),

the solutions of the system (1) with initial conditions

r(0)=r0,r˙(0)=0,d(0)=0,d˙(0)=b,θ(0)=0.r(0)=r_{0},\quad\dot{r}(0)=0,\quad d(0)=0,\quad\dot{d}(0)=b,\quad\theta(0)=0. (2)

It is clear that r(t)r(t), d(t)d(t) solves the reduced problem

{r¨=a2r32rmf(r,d),d¨=md2g(r,d),\begin{cases}\begin{split}\ddot{r}&=\frac{a^{2}}{r^{3}}-2rmf(r,d),\\ \ddot{d}&=-\frac{m\,d}{2}g(r,d),\end{split}\end{cases} (3)

then r(t)r(t), d(t)d(t) and θ(t)\theta(t) solves (1) if

θ(t)=0tar2(s)𝑑s.\theta(t)=\int_{0}^{t}\frac{a}{r^{2}(s)}ds.

Notice that if for some a,ba,b\in\mathds{R} the functions R(a,b,t)R(a,b,t) and D(a,b,t)D(a,b,t) have the same period T>0T>0 then, they provide a periodic hip-hop solution of (1) if and only if θ(a,b,T)\theta(a,b,T) is equal to pπ/qp\,\pi/q with pp and qq whole numbers. In general, TT-periodic solutions of the systems (3) define reduced-periodic solutions of the 2N2N-body problem (1). That is, solutions with the property that every TT units of time, the positions and velocities of the 2N2N-bodies only differ by an rigid motion in 3.\mathds{R}^{3}.

Let us present some useful results on the existence of symmetric periodic solutions of (1)(\ref{cilindricas}). To this end, let r0,m+r_{0},m\in\mathds{R}^{+} and consider the following initial value problem

{r¨=F(a,r,d),r(0)=r0,r˙(0)=0,d¨=G(r,d)d(0)=0,d˙(0)=b.\begin{cases}\begin{split}\ddot{r}&=F(a,r,d),\quad r(0)=r_{0},\quad\dot{r}(0)=0,\\ \ddot{d}&=G(r,d)\quad\quad d(0)=0,\quad\dot{d}(0)=b.\\ \end{split}\end{cases} (4)

with

F(a,r,d)=a2r32rmf(r,d),andG(r,d)=md2g(r,d).F(a,r,d)=\dfrac{a^{2}}{r^{3}}-2rmf(r,d),\quad\text{and}\quad G(r,d)=-\frac{md}{2}g(r,d).

Recall that, we are denoting the solutions of (4) as tR(a,b,t)t\to R(a,b,t) and tD(a,b,t)t\to D(a,b,t). Moreover, it is easy to check the following symmetries:

F(a,r,d)=F(a,r,d),andG(r,d)=G(r,d)F(a,r,d)=F(a,r,-d),\quad\text{and}\quad G(r,d)=-G(r,-d)

From previous symmetries it is clear that R(a,b,t)R(a,b,t) is even and D(a,b,t)D(a,b,t) is odd. Moreover, we have the following result.

Lemma 1.

Let tR(a,b,t)t\to R(a,b,t) and tD(a,b,t)t\to D(a,b,t) be a solution of (4).

  • \triangleright

    If for some 0<T0<T we have

    (I)Rt(a,b,T)=0,Dt(a,b,T)=0,\textbf{(I)}\quad R_{t}(a,b,T)=0,\quad D_{t}(a,b,T)=0,

    then R(a,b,t)R(a,b,t), D(a,b,t)D(a,b,t) are even functions respect to the line t=Tt=T. Moreover, both functions are 4T4T-periodic.

  • \triangleright

    If for some 0<T~0<\tilde{T} we have

    (II)Rt(a,b,T~)=0,D(a,b,T~)=0,\textbf{(II)}\quad R_{t}(a,b,\tilde{T})=0,\quad D(a,b,\tilde{T})=0,

    then with respect to the line t=T~t=\tilde{T}, D(a,b,t)D(a,b,t) is an odd function and R(a,b,t)R(a,b,t) is an even function. Moreover, both functions are 2T~2\tilde{T}-periodic.

Definition 1.

We say that a solution of (4) is of type I (type II) if it satisfies condition (I) (condition (II)) in Lemma 1. We also call the solutions of type I double symmetry solutions.

Remark 1.

If (a,b,T)(a,b,T) solve system (I) then (a,b,2T)(a,b,2T) solve system (II).

For a given N>1N>1 define the parameters

αN=116k=12N1((1)k1)2sin3(kπ2N),γN=14k=12N11sin(kπ2N).\alpha_{N}=\frac{1}{16}\sum_{k=1}^{2N-1}\frac{((-1)^{k}-1)^{2}}{\sin^{3}(\frac{k\pi}{2N})},\quad\gamma_{N}=\frac{1}{4}\sum_{k=1}^{2N-1}\frac{1}{\sin(\frac{k\pi}{2N})}. (5)
Proposition 2.

For any integer N>1N>1, the sum αN\alpha_{N} and γN\gamma_{N} defined in (5) satisfy

γNαN4+28.\frac{\gamma_{N}}{\alpha_{N}}\leq\frac{4+\sqrt{2}}{8}.
Proof.

This results is a direct consequence of Lemma 1 in [MR2570295]. ∎

Remark 2.

A direct computation shows that D(a,0,t)=0D(a,0,t)=0 for all a,t.a,t\in\mathds{R}. Moreover, if a=a0:=mγNr0a=a_{0}:=\sqrt{m\gamma_{N}r_{0}} then R(a0,0,t)=r0R(a_{0},0,t)=r_{0} for all tt\in\mathds{R}. Consequently, the points (a0,0,T)(a_{0},0,T) for all TT\in\mathds{R} satisfies the two systems of equations in three variables (I) and (II). We will call these solutions the trivial solutions of (I) and (II).

3. Existence of periodic solutions

In this section we state and prove the main theoretical result of this document.

Theorem 3.

Let N>1N>1 and m,r0>0m,r_{0}>0 fixed. Then there exists b^>0\hat{b}>0, and a pair of functions T,a:]b^,b^[T,a:]-\hat{b},\hat{b}[\to\mathds{R}, with

T(0)=π2r03mαN,anda(0)=mγNr0,T(0)=\frac{\pi}{2}\sqrt{\frac{r_{0}^{3}}{m\alpha_{N}}},\quad\text{and}\quad a(0)=\sqrt{m\gamma_{N}r_{0}},

such that R(a(b),b,t)=r(t)\displaystyle{R(a(b),b,t)=r(t)} and D(a(b),b,t)=d(t)D(a(b),b,t)=d(t) are 4T4T-periodic functions. Moreover, the points (a(b),b,T(b))(a(b),b,T(b)) solve the system (I).

Proof.

For fixed m,r0m,r_{0}, let R(a,b,t)=r(t)R(a,b,t)=r(t), D(a,b,t)=d(t)D(a,b,t)=d(t) be the solutions of (3). By Lemma 1 if for some a,ba,b and TT we have

Rt(a,b,T)=0,andDt(a,b,T)=0,R_{t}(a,b,T)=0,\quad\text{and}\quad D_{t}(a,b,T)=0, (6)

then r(t)r(t) is even respect to TT-axis whereas d(t)d(t) is an odd function and both are 4T4T-periodic. Let ζ(t)=(a0,0,t)\zeta(t)=(a_{0},0,t), t.t\in\mathds{R}. From Remark 2 it follows directly that

Rt(ζ(t))=0,andDt(ζ(t))=0,t.R_{t}(\zeta(t))=0,\quad\text{and}\quad D_{t}(\zeta(t))=0,\quad\forall t\in\mathds{R}.

In order to study possible bifurcations points in the system (6), we search a particular value t=T0t=T_{0} such that the set of vectors

{Rt(ζ(T0)),Dt(ζ(T0))},\Big{\{}\nabla R_{t}(\zeta(T_{0})),\nabla D_{t}(\zeta(T_{0}))\Big{\}},

are linearly dependent. To this end, since D(a,b,t)D(a,b,t) satisfies the initial value problem

Dtt=mD2g(R,D),D(0)=0,D˙(0)=b,D_{tt}=-\frac{mD}{2}g(R,D),\quad D(0)=0,\quad\dot{D}(0)=b, (7)

from the existence and uniqueness theorem of ordinary differential equations we deduce that D(a,0,t)=0\displaystyle{D(a,0,t)=0} for all (t,a)(t,a). From here,

D(a,b,t)=bD^(a,b,t),(a,b,t)D(a,b,t)=b\hat{D}(a,b,t),\qquad\forall(a,b,t) (8)

and

Dt(a,b,t)=bD^t(a,b,t),(a,b,t)D_{t}(a,b,t)=b\hat{D}_{t}(a,b,t),\qquad\forall(a,b,t) (9)

for some appropriate function D^\hat{D}. Moreover, direct computations shows that

Dt(ζ(t))=Da(ζ(t))=0,Db(ζ(t))=D^(ζ(t)),t.D_{t}(\zeta(t))=D_{a}(\zeta(t))=0,\quad D_{b}(\zeta(t))=\hat{D}(\zeta(t)),\quad\forall t\in\mathds{R}.
Dbt(ζ(t))=D^t(ζ(t)),t.D_{bt}(\zeta(t))=\hat{D}_{t}(\zeta(t)),\quad\forall t\in\mathds{R}.

Taking the partial derivative with respect to bb on both sides in (7), the function Db(ζ(t))D_{b}(\zeta(t)) satisfies the initial value problem

x¨+mαNr03x=0,x(0)=0,x˙(0)=1.\ddot{x}+\frac{m\alpha_{N}}{r_{0}^{3}}x=0,\quad x(0)=0,\quad\dot{x}(0)=1.

Therefore,

Db(ζ(t))=(mαNr03)1/2sin[(mαNr03)1/2t],D_{b}(\zeta(t))=\Big{(}\frac{m\alpha_{N}}{r^{3}_{0}}\Big{)}^{-1/2}\sin\Big{[}\Big{(}\frac{m\alpha_{N}}{r^{3}_{0}}\Big{)}^{1/2}t\Big{]}, (10)

and

Dbt(ζ(t))=D^t(ζ(t))=cos[(mαNr03)1/2t].D_{bt}(\zeta(t))=\hat{D}_{t}(\zeta(t))=\cos\Big{[}\Big{(}\frac{m\alpha_{N}}{r^{3}_{0}}\Big{)}^{1/2}t\Big{]}.

in consequence,

Dt(ζ(T0))=0,ifT0=π2r03mαN.D_{t}(\zeta(T_{0}))=0,\quad\text{if}\quad T_{0}=\frac{\pi}{2}\sqrt{\frac{r^{3}_{0}}{m\alpha_{N}}}.

This shows that Dt(ζ(T0))=0\displaystyle{\nabla D_{t}(\zeta(T_{0}))=\textbf{0}}. The previous computations, suggest to study the solutions of the system

Rt(a,b,T)=0,andD^t(a,b,T)=0,R_{t}(a,b,T)=0,\quad\text{and}\quad\hat{D}_{t}(a,b,T)=0,

around the point (a0,0,T0)(a_{0},0,T_{0}).

Firstly, we need to compute Ra(ζ(t))R_{a}(\zeta(t)), Rat(ζ(t))R_{at}(\zeta(t)) and Rb(ζ(t))R_{b}(\zeta(t)). For this purpose, let us recall that

Rtt=F(a,R,D),R(0)=r0,Rt(0)=0.R_{tt}=F(a,R,D),\quad R(0)=r_{0},\quad R_{t}(0)=0.

In consequence, taking the partial derivative with respect to aa on both sides, direct computations shows that the function u(t)=Ra(ζ(t))u(t)=R_{a}(\zeta(t)) satisfies the equation

u¨=Fa(a0,0,r0)+FR(a0,0,r0)u,\ddot{u}=F_{a}(a_{0},0,r_{0})+F_{R}(a_{0},0,r_{0})u,

where

FR(a,R,D)=3a2R42m[f(R,D)+RfR(R,D)],Fa(a,R,D)=2aR3.\begin{split}F_{R}(a,R,D)&=-3\frac{a^{2}}{R^{4}}-2m\left[f(R,D)+Rf_{R}(R,D)\right],\\ F_{a}(a,R,D)&=\frac{2a}{R^{3}}.\end{split} (11)

From here, direct computations shows that

f(r0,0)=γN2r03,andfR(r0,0)=3γN2r04,f(r_{0},0)=\frac{\gamma_{N}}{2r_{0}^{3}},\quad\text{and}\quad f_{R}(r_{0},0)=\frac{-3\gamma_{N}}{2r_{0}^{4}},

with γN\gamma_{N} given in (5). Therefore, u(t)u(t) satisfies the initial value problem, but a02=mr0γN\displaystyle{a_{0}^{2}=mr_{0}\gamma_{N}}, therefore

u¨+w2u=2w/r0,u(0)=0,u˙(0)=0,\ddot{u}+w^{2}u=2w/r_{0},\quad u(0)=0,\quad\dot{u}(0)=0,

with w2=mγN/r03\displaystyle{w^{2}=m\gamma_{N}/r_{0}^{3}}, where the solution is given by

Ra(ζ(t))=2wr0(1cos(ωt)),R_{a}(\zeta(t))=\frac{2}{wr_{0}}(1-\cos(\omega t)),

implying that

Rat(ζ(t))=2r0sin(ωt).R_{at}(\zeta(t))=\frac{2}{r_{0}}\sin(\omega t).

In the same fashion, the function v(t)=Rb(ζ(t))v(t)=R_{b}(\zeta(t)) satisfies the differential equation

v¨=FD(a0,r0,0)Db(ζ(t))+FR(a0,r0,0)v.\ddot{v}=F_{D}(a_{0},r_{0},0)D_{b}(\zeta(t))+F_{R}(a_{0},r_{0},0)v.

Since r(0)=r0r(0)=r_{0} and r˙(0)=0\dot{r}(0)=0, the function v(t)v(t) is the solution of the initial value problem

v¨+w2v=0,v(0)=0,v˙(0)=0.\ddot{v}+w^{2}v=0,\quad v(0)=0,\quad\dot{v}(0)=0.

In consequence,

Rb(ζ(t))=0,andRbt(ζ(t))=0,t.R_{b}(\zeta(t))=0,\quad\text{and}\quad R_{bt}(\zeta(t))=0,\quad\forall t.

Finally, it follows directly that

Rtt(ζ(t))=F(r0,0,a0)=a02r03mγNr02=0t.R_{tt}(\zeta(t))=F\left(r_{0},0,a_{0}\right)=\frac{a_{0}^{2}}{r_{0}^{3}}-\frac{m\gamma_{N}}{r_{0}^{2}}=0\quad\forall t.

In brief, the previous calculation provide for Rt\nabla R_{t} at ζ(T0)\zeta(T_{0}) that

Rt(ζ(T0))=(2r0sin(ωT0),0,0),\nabla R_{t}\big{(}\zeta(T_{0})\big{)}=\big{(}\frac{2}{r_{0}}\sin(\omega T_{0}),0,0\big{)}, (12)

where ωT0=(πγN/αN)/2\displaystyle{\omega T_{0}=(\pi\sqrt{\gamma_{N}/\alpha_{N}}})/2. From Proposition 2 it follows that ωT0pπ\omega T_{0}\neq p\pi for every positive integer pp implying that Rt(ζ(T0))\nabla R_{t}(\zeta(T_{0})) does not vanish. Finally, we use the relation (9) to compute D^t(ζ(T0))\displaystyle{\nabla\hat{D}_{t}(\zeta(T_{0}))}. Notice that,

Dbt(ζ(t))=D^t(ζ(t)),andDba(ζ(t))=D^a(ζ(t)).D_{bt}(\zeta(t))=\hat{D}_{t}(\zeta(t)),\quad\text{and}\quad D_{ba}(\zeta(t))=\hat{D}_{a}(\zeta(t)).

Then, from (10) we have

D^t(ζ(T0))=1,andD^a(ζ(T0))=0.\hat{D}_{t}(\zeta(T_{0}))=-1,\quad\text{and}\quad\hat{D}_{a}(\zeta(T_{0}))=0.

Moreover

D^t(ζ(t))=(Dtba(ζ(t)),12Dbbt(ζ(t)),Dbtt(ζ(t))).\nabla\hat{D}_{t}(\zeta(t))=\Big{(}D_{tba}(\zeta(t)),\frac{1}{2}D_{bbt}(\zeta(t)),D_{btt}(\zeta(t))\Big{)}. (13)

To sum up, from expressions (12) and (13) we obtain

Λ:=Rt(ζ(T0))×D^(ζ(T0))=(0,2Dbtt(ζ(T0))r01sin(ωT0),Dbbt(ζ(T0))r01sin(ωT0)).\Lambda:=\nabla R_{t}(\zeta(T_{0}))\times\nabla\hat{D}(\zeta(T_{0}))=\big{(}0,-2D_{btt}(\zeta(T_{0}))r_{0}^{-1}\sin(\omega T_{0}),D_{bbt}(\zeta(T_{0}))r_{0}^{-1}\sin(\omega T_{0})\big{)}.

Since,

Dbtt(ζ(t))=D^tt(ζ(t))=(mαNr03)1/2sin[(mαNr03)1/2t],D_{btt}(\zeta(t))=\hat{D}_{tt}(\zeta(t))=-\Big{(}\frac{m\alpha_{N}}{r^{3}_{0}}\Big{)}^{1/2}\sin\Big{[}\Big{(}\frac{m\alpha_{N}}{r^{3}_{0}}\Big{)}^{1/2}t\Big{]},

we have Dbtt(ζ(T0/2))=π/2T0.\displaystyle{D_{btt}(\zeta(T_{0}/2))=-\pi/2T_{0}}. This shows that the second component of Λ\displaystyle{\Lambda} is different from zero. Then, by the Implicit Function Theorem, there exists b^>0\hat{b}>0, and a unique pair of functions T,a:]b^,b^[T,a:]-\hat{b},\hat{b}[\to\mathds{R}, such that

bT(b),ba(b),b\to T(b),\quad b\to a(b),

for b]b^,b^[b\in]-\hat{b},\hat{b}[, with T(0)=T0T(0)=T_{0} and a(0)=a0a(0)=a_{0}, such that

Rt(λ(b))=0,D^t(λ(b))=0,R_{t}(\lambda(b))=0,\quad\hat{D}_{t}(\lambda(b))=0,

with

λ:]b^,\displaystyle\lambda:]-\hat{b}, b^[3\displaystyle\hat{b}\left[\longrightarrow\mathbb{R}^{3}\right.
bλ(s)=(a(b),b,T(b)).\displaystyle b\longrightarrow\lambda(s)=(a(b),b,T(b)).

Therefore, for each b]b^,b^[b\in]-\hat{b},\hat{b}[ it follows

Rt(λ(b))=0,andDt(λ(b))=bD^t(λ(b))=0.R_{t}(\lambda(b))=0,\quad\text{and}\quad D_{t}(\lambda(b))=b\hat{D}_{t}(\lambda(b))=0.

By Lemma 1 we get that for any b]b^,b^[b\in]-\hat{b},\hat{b}[ the functions

r(t)=R(a(b),b,t),andd(t)=D(a(b),b,t),r(t)=R(a(b),b,t),\quad\text{and}\quad d(t)=D(a(b),b,t),

provides a 4T(b)4T(b)-periodic solution of the reduced problem (3) with d(t)d(t) an odd function, whereas r(t)=R(a(b),b,t)r(t)=R(a(b),b,t) and d(t)=D(a(b),b,t)d(t)=D(a(b),b,t) are both even respect to the line t=T(b)t=T(b). ∎

Due to Remark 1 we have the following result:

Theorem 4.

Let N>1N>1 and m,r0>0m,r_{0}>0 fixed. Then there exists b^>0\hat{b}>0, and a pair of functions T,a:]b^,b^[T,a:]-\hat{b},\hat{b}[\to\mathds{R}, with

T(0)=πr03mαN,anda(0)=mγNr0,T(0)=\pi\sqrt{\frac{r_{0}^{3}}{m\alpha_{N}}},\quad\text{and}\quad a(0)=\sqrt{m\gamma_{N}r_{0}},

such that R(a(b),b,t)=r(t)\displaystyle{R(a(b),b,t)=r(t)} and D(a(b),b,t)=d(t)D(a(b),b,t)=d(t) are a 2T2T-periodic functions. Moreover, the points (a(b),b,T(b))(a(b),b,T(b)) solve the system (II).

4. Branches emanating from the bifurcations

In this section we use analytic continuation to extend the branches emanating form the bifurcations points

p0=(mγNr0, 0,π2r03mαN)andq0=(mγNr0, 0,πr03mαN),p_{0}=\left(\,\sqrt{m\gamma_{N}r_{0}},\,0,\,\frac{\pi}{2}\sqrt{\frac{r_{0}^{3}}{m\alpha_{N}}}\,\right)\quad\hbox{and}\quad q_{0}=\left(\,\sqrt{m\gamma_{N}r_{0}},\,0,\,\pi\sqrt{\frac{r_{0}^{3}}{m\alpha_{N}}}\,\right),

obtained in Theorem 3 and Theorem 4.

Notice that the solutions from p0p_{0}, q0q_{0} are essentially the same, all of them have double symmetry. With the purpose of searching for solutions of type (II) that are not of type (I) we extend the branch starting at the point q0q_{0} by applying analytic continuation and a method similar to the one presented in [MR2267950] to system (II). Taking N=3N=3, r0=2r_{0}=2 and m=1m=1 we have

q0=(52+23,0,4217π)(1.91173,0,4.31023).q_{0}=\left(\sqrt{\frac{5}{2}+\frac{2}{\sqrt{3}}},0,4\sqrt{\frac{2}{17}}\pi\right)\approx(1.91173,0,4.31023).

The analytic continuation method give us a table that we labeled DSP because, as explained above, near the point q0q_{0}, the solutions satisfy the double symmetry property. We found a bifurcation point –that we labeled BB– along this branch that gave us a new branch with single symmetric hip-hop solutions. Recall from the introduction that a solution is called single symmetric if d(t)d(t) is even and there is not T0T\neq 0 such that d(Tt)=d(T+t)d(T-t)=d(T+t) for all tt. Before we address the existence of this bifurcation point BB, let us elaborate on some properties of the DSP branch.

Figure 3 shows the table DSP={Q1=q0,Q2,,Q5806=qf}.DSP=\{Q_{1}=q_{0},Q_{2},\dots,Q_{5806}=q_{f}\}. It is a list of points in the a,b,Ta,b,T space that satisfies the conditions

|Rt(Qi)|<104|F(Qi)|<104for each QiDSP.|R_{t}(Q_{i})|<10^{-4}\ \ |F(Q_{i})|<10^{-4}\ \ \ \text{for each $Q_{i}\in DSP$}.

For the solutions along the DSP branch the system of particles reaches the maximum vertical expansion (that is expansion in the zz-direction) at the same time that it reaches a maximum contraction towards the zz-axis. In other words, when the particles are at their maximum height they are also the closest they can be to the zz-axis. This is illustrated by the joint graph of d(t)d(t) and r(t)r(t) for the points Q210Q_{210} and Q4225Q_{4225} shown in Figure 2.

Refer to caption
Figure 2. The graph of the functions d(t)d(t) and r(t)r(t) for two solutions in the branch DSP. Since the point Q210Q_{210} is close to the point q0q_{0} then r(t)r(t) is almost constant and the function d(t)d(t) does not oscillate much. Recall that the solution q0q_{0}, d(t)d(t) vanishes and r(t)r(t) are constant. Both solutions show that after a quarter of a period of the function d(t)d(t), r(t)r(t) reaches a minimum while d(t)d(t) reaches a maximum. For this reason these solutions have an additional symmetry.
Refer to caption
Figure 3. Points in the space aa, bb, TT found by doing analytic continuation near the bifurcation point q0q_{0} provided by Theorem 3. All these points are associated with solutions that have double symmetry and for this reason we have labelled this family of solutions as DSP. We show the bifurcation point BB, and the points Q210Q_{210} and Q4225Q_{4225}. We will show later that: (i)(i) we will have another family of solutions with not double symmetry emanating from the point BB, (ii)(ii) the point Q210Q_{210} will represent a solution where bodies 1, 3 and 5 share the same trajectory and (iii)(iii) bodies 2, 4 and 6 also share the same trajectory. For the solution represented by the point Q4225Q_{4225}, all the six trajectories are different.

The branch DSP ends at the point

qf=(0.259786,0.780202,5.80955),q_{f}=(0.259786,0.780202,5.80955),

a point with angular moment aa near zero, close to a solution that represents a collision, with three of the bodies colliding at maximum height.

Refer to caption
Figure 4. This is the graph of the functions d(t)=D(0.259786,0.780202,t)d(t)=D(0.259786,0.780202,t) and r(t)=R(0.259786,0.780202,t)r(t)=R(0.259786,0.780202,t) which is the solution associated with point qf=Q5806q_{f}=Q_{5806} located at the end of the branch that starts at the point q0q_{0}. As we can see, this solution has double symmetry, both functions have derivatives vanishing at 0.5809552\frac{0.580955}{2}. We can also see how this solution is near a solution with a collision of 33 bodies near the point (x,y,z)=(0,0,1.39)(x,y,z)=(0,0,1.39) and the collision of the other 33 bodies near the point (x,y,z)=(0,0,1.39).(x,y,z)=(0,0,-1.39).

The name for the set DSPDSP stands for “double symmetry points” because each one of these points represent a solution with period 2T2T with dd and rr even with respect to t=T/2t=T/2, rr even with respect to t=Tt=T and dd odd with respect to t=Tt=T.

4.1. Bifurcation point along the DSP branch

By the implicit function theorem, we have that as long as the vectors D(a0,b0,T0)\nabla D(a^{0},b^{0},T^{0}) and Rt(a0,b0,T0)\nabla R_{t}(a^{0},b^{0},T^{0}) are linearly independent for points (a0,b0,T0)(a^{0},b^{0},T^{0}) that satisfy

{Rt(a,b,T)= 0,F(a,b,T)= 0,\begin{cases}\begin{split}R_{t}(a,b,T)=&\,0,\\ F(a,b,T)=&\,0,\end{split}\end{cases} (14)

then, the solution of the system (14) is given by a smooth curve (not bifurcation points near that point) near (a0,b0,T0)(a^{0},b^{0},T^{0}). We have noticed that at the point B=(1.34958,0.727361,7.05373)B=(1.34958,0.727361,7.05373), which is one of the points in the set DSPDSP, satisfies that

D(B)=(4.58986,17.2712,0.727943),Rt(B)=(1.44703,5.44591,0.229381),\nabla D(B)=(4.58986,17.2712,-0.727943),\quad\nabla R_{t}(B)=(1.44703,5.44591,-0.229381),

and

D(B)×Rt(B)=(0.0026399,0.000527769,0.0041234).\nabla D(B)\times\nabla R_{t}(B)=(0.0026399,-0.000527769,0.0041234).

The information above provided numerical evidence of the possible existence of a point in the branch where the gradients D(a,b,T)\nabla D(a,b,T) and Rt(a,b,T)\nabla R_{t}(a,b,T) are linearly dependent. Therefore we search for solution of system (14) near BB but away from the smooth curve suggested by the points in the set DSPDSP. Indeed we were able to find a point that satisfy the system (14), away from the trajectory of the points in DSP but near the point BB. After doing analytic continuation to this new point, we were able to find the collection of points

SSP={W1,W2,,W14154},SSP=\{W_{1},W_{2},\dots,W_{14154}\},

all of them satisfying |Rt(Wi)|<104|R_{t}(W_{i})|<10^{-4} and |F(Wi)|<104|F(W_{i})|<10^{-4}. The name SSPSSP stand for “single symmetry points” because the solutions associated with these points do not satisfy that Rt(a,b,T/2)=0R_{t}(a,b,T/2)=0 and F(a,b,T/2)=0F(a,b,T/2)=0. Figure 5 shows how the two branches and the point BB.

Therefore, we have numerically found solution with only one symmetry.

Refer to caption
Figure 5. Points of the two branches SSP and DSP. The branch DSP starts at q0q_{0} while the branch SSP emanated from point that satisfies the system (14) near BB and away from the trajectory DSP. We have highlighted the point W1257W_{1257} because this point represent a periodic solution where the bodies 1, 3 and 5 share the same trajectory and the bodies 2, 4 and 6 also share the same trajectory.

The solutions that correspond to the SSP branch are characterized by the fact that the maximum contraction toward the zz-axis occurs when the system is still expanding (for the points before the point of bifurcation) or already contracting (for the points after the point of bifurcation) in the vertical direction. This is illustrated in Figure 6.

Refer to caption
Figure 6. The image on the left shows the graph of the solutions r(t)r(t) and d(t)d(t) associated with the point W3000W_{3000}\in SSP. We can see how the maximum of d(t)d(t) is obtained before the minimum of the function rr. This means that when the bodies reaches the maximum vertical distance, they continue to move closer to the zz-axis. The image in the center shows the graph of the solutions r(t)r(t) and d(t)d(t) associated with the bifurcation point. These two functions have double symmetry. The image on the right shows the graphs of the solutions r(t)r(t) and d(t)d(t) associated with the point W6000W_{6000}\in SSP. We can see how the maximum of d(t)d(t) is obtained after the minimum of the function rr.

4.2. On the number of trajectories and choreographic solutions

In theory, the six bodies of the periodic solutions can follow 6,3,2,6,3,2, or one single trajectory. In the last case, when all the bodies follow the same orbit, the solution is called a choreography. In this subsection, we show examples of the four cases.

If we label the bodies 1, 2, 3, 4, 5, 6 according to the order of their projections on the xx-yy plane in the direction of rotation about the zz-axis then the number of trajectories is determined by which of the starting positions will be reached first by the body number 1. More precisely,

Let 1j1\rightarrow j denote the fact that for some instant of time t>0t>0 the condition

C(t,k):r1(t)=rk(0),andr˙1(t)=r˙k(0),C(t,k):r_{1}(t)=r_{k}(0),\ \text{and}\ \ \dot{r}_{1}(t)=\dot{r}_{k}(0),

is met by k=jk=j and that for every 0<τ<t0<\tau<t and all k{1,2,,6}k\in\{1,2,...,6\} the condition C(τ,k)C(\tau,k) is false.

Then the number of trajectories are:

Six in the case 111\rightarrow 1

One in the case 121\rightarrow 2 or 15.1\rightarrow 5.

Two in the case 131\rightarrow 3

Three in the case 14.1\rightarrow 4.

4.2.1. Number of orbits for solutions on the branch DSPDSP.

For the solutions in the branch DSPDSP we have that the number of trajectories of the solution depends on θ0=θ(T)\theta_{0}=\theta(T): the angle of rotation of the solution after half of the common period of the functions r(t)r(t) and d(t)d(t). Since for every solution on this branch, the function rr is even with respect to t=T/2t=T/2, then we have that r(T)=r(0)=2r(T)=r(0)=2. We also have that for any integer kk, θ(kT)=kθ0\theta(kT)=k\theta_{0}. Notice that when the solution is periodic, we can find three integers k,jk,j and ll that satisfy the following integer equation

IE(k,j,l):kθ0=jπ3+2πl.\displaystyle IE(k,j,l):k\theta_{0}=j\frac{\pi}{3}+2\pi l. (15)

Recall that d(kT)=0d(kT)=0 for every integer kk and d˙(kT)<0\dot{d}(kT)<0 if kk is odd and d˙(kT)>0\dot{d}(kT)>0 if kk is even. Also notice that kθ0=jπ3+2πlk\theta_{0}=j\frac{\pi}{3}+2\pi l for some integer k,jk,j and ll indicates that the body starting at (r0,0,0)(r_{0},0,0) moves, after kTkT units of time to the initial position of the body starting at (r0cos(jπ2),r0sin(jπ2),0)(r_{0}\cos(j\frac{\pi}{2}),r_{0}\sin(j\frac{\pi}{2}),0). Each body rotates jπ3j\frac{\pi}{3} radians during their first kTkT units of time.

For a periodic solution, let us denote k0k_{0} smallest positive integer such that IE(k0,j,l)IE(k_{0},j,l) is satisfied for a pair of integers jj and ll with j>0j>0 and k0+jk_{0}+j even. We denote by j0j_{0} the smallest positive integer with j0+k0j_{0}+k_{0} even such that IE(k0,j0,l)IE(k_{0},j_{0},l) is satisfied for some integer ll. We have that if j0j_{0} is 11 or 55, then the solution is a choreography, this is, we have only one trajectory that is share by the six bodies. If j0j_{0} is 22 or 44 then, there are two trajectories. If j0j_{0} is 33, then there are three trajectories and if j0j_{0} is 66 then, there are 66 trajectories.

As explain above, the angle θ0(T)\theta_{0}(T) plays an important role determining the number of trajectories. Figure 7 shows how the angle θ0\theta_{0} changes for different values of TT in the branch DSPDSP. Since the first point in DSPDSP is near a trivial solution, we can compute the starting value for these curve. In this particular case, it is 151(43+15)π2.06\sqrt{\frac{1}{51}\left(4\sqrt{3}+15\right)}\pi\approx 2.06.

Refer to caption
Figure 7. Graph of the points (i,θ0(Ti))(i,\theta_{0}(T_{i})) where TiT_{i} the third entry of QiQ_{i} in the branch DSP={Q1=q0,Q2,,Q5806=qf}DSP=\{Q_{1}=q_{0},Q_{2},\dots,Q_{5806}=q_{f}\}. This graph points out the fact that 5π/35\pi/3 is in the range of the function θ0\theta_{0}.

As pointed out in Figure 7, 5π/35\pi/3 is in the range of the function θ\theta as a function of TT. When TT moves along the smooth curve of points (a,b,T)(a,b,T) extending the points in DSP. The closest value of θ(T)\theta(T) to 5π/35\pi/3 in the branch DSP={Q1,Q5806}DSP=\{Q_{1}\dots,Q_{5806}\} happens for the solution 50105010, we have that θ(T5009)<5π/3<θ(T5010)\theta(T_{5009})<5\pi/3<\theta(T_{5010}) and θ(T5010)5π/3<0.00033\theta(T_{5010})-5\pi/3<0.00033. The intermediate value theorem shows that there is choreographic in the family of solution with double symmetry. Let us explain the orbit of this choreographic, assuming for practical reasons that it is given by the solution 5010 in the branch DSP. The initial configuration of the six bodies is displayed in Figure 8

Refer to caption
Figure 8. Labelling of the bodies. Initially they form a regular dodecagon.

The arrows indicate if the bodies start going up or down. Notice that those bodies that go up, they do not do it vertically due to the rotation motion that all of them are doing. The point Q5010=(a,b,T)=(0.581691,0.810807,6.53465)Q_{5010}=(a,b,T)=(0.581691,0.810807,6.53465), this means that after T=6.53465T=6.53465 units of time, the body 1 will move to the initial position of body 6. Figure 9 shows the rest of the details of this motion. In this case, in terms of the equation IE(k,j,l)IE(k,j,l) in (15) we have that k0=1k_{0}=1, j0=5j_{0}=5 and l=5l=5.

Refer to caption
Figure 9. The first image shows the trajectory of the body 1 in the solution represented by Q5010Q_{5010} for values of tt between 0 and T=6.53465T=6.53465. The second image shows the trajectories of bodies 1 and 6 in the same interval of time. The third image show the trajectories of all 6 bodies in the same interval of time. This third image is also the trajectory of each one of the bodies between t=0t=0 and t=6Tt=6T.

In the same way, 5π/35\pi/3 is in the range of θ0(T)\theta_{0}(T), we can check that π\pi is in the range of this function. This time the closest value of θ(T)\theta(T) to π\pi in the branch DSPDSP happens for the solution represented with the point Q2878Q_{2878}, we have that 0<πθ0(T2878)<0.000070<\pi-\theta_{0}(T_{2878})<0.00007. Let us explain the trajectory of this solution, assuming for practical reasons that the solution represented by Q2878Q_{2878} in the branch DSP satisfies θ0(T)=π\theta_{0}(T)=\pi. The point Q2878=(1.37188,0.717167,6.95687)Q_{2878}=(1.37188,0.717167,6.95687), this means that after T=6.95687T=6.95687 units of time, the body 1 will move to the initial position of body 4. The rest of the orbits are explained in Figure 10.

Refer to caption
Figure 10. The image on the left shows the trajectory of the body 1 in the solution represented by Q2878Q_{2878} for values of tt between 0 and T2878=6.95687T_{2878}=6.95687 and the second image shows the trajectory in the same interval for the bodies 1 and 4. This second image represents the whole orbit for the first and fourth body. They share the same orbit. Bodies 2 and 5 share the same trajectory and the bodies 3 and 6 also share the same orbit. The third image shows the three trajectories.

The solution represented by Q210=(1.88461,0.175173,4.41712)Q_{210}=(1.88461,0.175173,4.41712) provides a solution that has two trajectories. We have that θ0(T210)2π3\theta_{0}(T_{210})\approx\frac{2\pi}{3}. In this case after T210=4.41712T_{210}=4.41712 the body one reaches the initial position of body 3 but it does not reach it with the right direction of the velocity, after 2T2T units of time, the body 1 reaches the initial position of body 5 with the same initial velocity as well. For this solution k0=2k_{0}=2, j0=4j_{0}=4 and l=0l=0, see Figure 11.

Refer to caption
Figure 11. The first image shows the trajectory of the body 1 in the solution represented by Q210Q_{210} for values of tt between 0 and 2T2102T_{210} where T210=4.41712T_{210}=4.41712, the second image shows the trajectory in the same interval for the bodies 1 and 5 and the third image shows the trajectory in the same interval for the bodies 1, 5 and 3. This third image represent the whole orbit for the first, third and fifth body. They share the same orbit. The bodies 2, 4 and 6 also share the same trajectory. The fourth image shows the two trajectories.

The solution represented by Q4225=(0.827163,0.825182,7.28011)Q_{4225}=(0.827163,0.825182,7.28011) provides a solution that has six trajectories. We have that θ0(T4225)3π2\theta_{0}(T_{4225})\approx\frac{3\pi}{2}. In this case the body one reaches back its initial position and velocity after 4T42254T_{4225} units of time. Moreover, it never reaches the initial position and velocity of the other five bodies. For this solution k0=4k_{0}=4, j0=6j_{0}=6 and l=0l=0, see Figure 12.

Refer to caption
Figure 12. The second image shows the trajectory of the body 1 in the solution represented by Q4225Q_{4225} for values of tt between 0 and T4225=7.28011T_{4225}=7.28011. The first image shows the trajectory of the body 1 for values of tt between 0 and 2T42252T_{4225}. Notice that even though the ending point of this portion of this trajectory is the starting position for the fourth body, the velocity is not the initial velocity of the fourth body. The fourth image shows the trajectory of the body 1 for values of tt between 0 and 3T42253T_{4225} and the third image shows the trajectory of the body 1 for values of tt between 0 and 4T42254T_{4225}.

4.2.2. Number of orbits for solutions on the branch SSP

Before we start this subsection, let us point out the following:

Proposition 5.

Let us assume that a periodic solution of the Hip-Hop problem is given by the functions (d(t),r(t),θ(t))(d(t),r(t),\theta(t)) satisfying the following condition:

  1. (1)

    d(t)d(t) and r(t)r(t) are periodic with period 2T2T for some T>0T>0,

  2. (2)

    d(t)=0d(t)=0 only for values of t=kTt=kT with kk an integer,

  3. (3)

    r(kT)r((k+1)T)r(kT)\neq r((k+1)T) for all kk an integer,

then, the solution cannot be a choreography.

Proof.

Let us argue by contradiction. If we have a choreography, then there exist a time T~\tilde{T} when the body 1 reaches the initial position and velocity of the body 2. By hypothesis (2) T~=kT\tilde{T}=kT for some integer kk. On the other hand our two conditions on the function rr, (period equal to 2T2T and r(lT)r((l+1)T)r(lT)\neq r((l+1)T) with ll an integer) implies that r(lT)=r(0)r(lT)=r(0) if ll is even and r(lT)r(0)r(lT)\neq r(0) if ll is odd. Since r(kT)=r0r(kT)=r_{0} then kk must be even. Since dd has period 2T2T then d˙(kT)=d˙(0)\dot{d}(kT)=\dot{d}(0). This is a contradiction because the initial velocity of the body 2 goes in the “opposite” directions. If body 1 starts going up, the body 2 starts going down. ∎

One feature of the solutions of the SSP branch (other than the one that corresponds to the point of bifurcation BB) is that any body intersect the plane z=0z=0 only for values of t=kTt=kT where kk is an integer. More importantly, the distance to the zz-axis of any of the bodies changes when tt increases changes from t=kTt=kT to t=(k+1)Tt=(k+1)T, this is, r(kT)r((k+1)T)r(kT)\neq r((k+1)T). In particular the condition 1j1\longrightarrow j never occurs if jj is even. Therefore, in the SSP branch the only possible periodic solutions will have two trajectories with the bodies 1, 3, 5 sharing one orbit and 2, 4, 6 the other.

In the equation (15) when 131\rightarrow 3 we have j=2j=2 and when 151\rightarrow 5 we have j=4.j=4. We set k=2q,j=2pk=2q,j=2p in the condition IE(k,j,l)IE(k,j,l)

IE(k,j,l):2qθ0=2pπ3+2πl,\displaystyle IE(k,j,l):2q\theta_{0}=2p\frac{\pi}{3}+2\pi l, (16)

that simplifies to

qθ0=pπ3+πl,\displaystyle q\theta_{0}=p\frac{\pi}{3}+\pi l, (17)

where qq and ll are positive integers and p=1p=1 or p=2.p=2.

Assume p=1.p=1. The closest value of qθ(T)q\theta(T) to π/3+πl\pi/3+\pi l in the branch SSP={P1,P14154}SSP=\{P_{1}\dots,P_{14154}\} happens at P1257=(0.886201,0.557961,3.61393)P_{1257}=(0.886201,0.557961,3.61393), with q=l=13.q=l=13. have that

13θ(T1257)(π3+13π)<0.000102,13\theta(T_{1257})-\left(\frac{\pi}{3}+13\pi\right)<0.000102,

and

13θ(T1258)(π3+13π)>0.000133.13\theta(T_{1258})-\left(\frac{\pi}{3}+13\pi\right)>-0.000133.

The intermediate value theorem guarantees that there is solution with single symmetry that has exactly two orbits with the bodies 1, 3, 5 in one and 2, 4, 6 in the other. The first image on the Figure between the abstract and the Introduction shows both trajectories. For practical purposes we assume that this solution is given by the solution 1257 in the SSP branch. Figures 13 and 14 explain the trajectories in this particular solution.

Refer to caption
Figure 13. Images for the solution given by the point P1257P_{1257} in SSPSSP. The second image shows the trajectory of body 1 from t=0t=0 to t=T1257=3.61393t=T_{1257}=3.61393. Notice that the body comes back to the plane z=0z=0 with a value of rr different from 22. The first image show the trajectory of body. From t=0t=0 to t=2T1257t=2T_{1257}. This time the end position of the body is 2 units away from the zz-axis. The third image shows the trajectory of body 1 from t=0t=0 to t=26T1257t=26T_{1257}. We can see how the trajectory of the body ends at the initial position of body 3.
Refer to caption
Figure 14. The first image shows the trajectory of body 1 from t=26T1257t=26T_{1257} to t=52T1257t=52T_{1257}. We can see how the trajectory of the body ends at the initial position of body 5. This trajectory agrees with the trajectory of body 3 from t=0t=0 to t=26T1257t=26T_{1257}. The third image shows the trajectory of body 1 from t=52T1257t=52T_{1257} to t=78T1257t=78T_{1257}. The second image shows the trajectory of body 1 from t=0t=0 to t=78T1257t=78T_{1257}. This is a closed embedded curve which is also the trajectory of bodies 3 and 5.

References