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

Relative Equilibria and Periodic Orbits in a Binary Asteroid Model

Lennard F. Bakker, Nicholas J. Freeman
Brigham Young University, USA
Abstract

We present a planar four-body model, called the Binary Asteroid Problem, for the motion of two asteroids (having small but positive masses) moving under the gravitational attraction of each other, and under the gravitational attraction of two primaries (with masses much larger than the two asteroids) moving in uniform circular motion about their center of mass. We show the Binary Asteroid Model has (at least) 6 relative equilibria and (at least) 10 one-parameter families of periodic orbits, two of which are of Hill-type. The existence of six relative equilibria and 8 one-parameter families of periodic orbits is obtained by a reduction of the Binary Asteroid Problem in which the primaries have equal mass, the asteroids have equal mass, and the positions of the asteroids are symmetric with respect to the origin. The remaining two one-parameter families of periodic orbits, which are of comet-type, are obtained directly in the Binary Asteroid Problem.

1 Introduction

In December 2017, the asteroid 2017 YE5{\rm YE}_{5} was discovered by Claudine Rinner as part of the Morocco Oukaïmeden Sky Survey directed by Zouhair Benkhaldoun [7]. Only later in June 2018, when 2017 YE5{\rm YE}_{5} passed close enough to Earth was it determined by ground-based radar imaging that it was a binary asteroid where the two asteroids have nearly equal mass, i.e., the mass ratio of the two asteroids is close to 11 [5]. There are many known binary asteroids with a larger mass central asteroid and a smaller mass satellite asteroid [10], but only four known near-Earth binary asteroids with nearly equal mass, with 2017 YE5{\rm YE}_{5} being the fourth discovered. The other three known near-Earth binary asteroids with nearly equal mass are (69230) Hermes, (190166) 2005 UP156{\rm UP}_{156}, and 1994 CJ1{\rm CJ}_{1} [15]. In the main asteroid belt there are two more binary asteroids with nearly equal mass, and these are 90 Antiope and (300163) 2006 VW139{\rm VW}_{139} [10]. Of these six binary asteroids with nearly equal mass, (300163) 2006 VW163{\rm VW}_{163} is also classified as a comet with designation 288P [10], and 2017 YE5{\rm YE}_{5} is also considered a possible dormant Jupiter-family comet [15].

The standard model for the motion of an asteroid (be it a single asteroid, a binary asteroid with arbitrary mass ratio, etc.) is with six orbital elements (the orbital elements of known binary asteroids are listed in [10]). This model uses observational data of the asteroid and the integrable two-body problem to estimate the asteroid’s orbital elements (e.g., see [8]) that describe the ellipse the asteroid approximately makes around the Sun.

Another model for the motion of a binary asteroid is the Restricted Hill Full 44-Body Problem [17], which is known as the Binary Asteroid System. This model has one body with large mass (the Sun) at a large distance from the two bodies with smaller masses (the binary asteroid with arbitrary mass ratio) and a spacecraft (the zero-mass particle). It is assumed that two asteroids and the spacecraft remain close together. This model further assumes each of the asteroids has positive volume and a mass distribution. One of the asteroids is typically assumed to be a homogeneous sphere, and the other is typically assumed to be a constant density triaxial ellipsoid. A recent extension of the Binary Asteroid System model includes the Solar Radiation Pressure in the model, which, using an analytical solution as a seed for a multiple shooting method, has been applied to design a quasiperiodic forced hovering orbit above the one of asteroids for the unequal mass binary asteroid (66391) Moshup and Squannit [22]. When both asteroids have irregular shape and heterogeneous mass distributions, a nested interpolation method has been developed, based on a hybrid gravity model on one asteroid and a finite element model of the other, which has been applied to the unequal mass binary asteroid (66391) Moshup and Squannit [18].

In this paper we present a four point-mass model for the motion of a binary asteroid. The spatial version of this model was first proposed for the binary asteroid 2017 YE5{\rm YE}_{5} [4], but it applies to any two asteroids. The planar model, formulated here, has two bodies having large masses (called the primaries) and two bodies having small but positive masses (called the asteroids). We do not assume a priori that the masses of the primaries are equal nor do we assume a priori that the masses of the asteroids are equal. The primaries are prescribed to move in uniform circular motion around their center of mass (set at the origin), as is done in the Circular Restricted Planar Three-Body Problem. The two asteroids move under the gravitational attraction of the two primaries as well as the gravitational attraction of each other. We do not assume that the two asteroids are necessarily close to each other, that they are not a priori in the formation of a binary asteroid. The system of ODEs associated to the Hamiltonian in rotating coordinates (which Hamiltonian is derived in Section 2.2) we call the Binary Asteroid Problem. The analytic existence of 6 relative equilibria and 10 one-parameter families of periodic orbits in the Binary Asteroid Problem presented here form part of the second author’s senior thesis [9].

The Binary Asteroid Problem serves as a “bridge” model between two uncoupled copies of the Circular Restricted Planar Three-Body Problem and the general Planar Four-Body Problem. The connection with the Circular Restricted Planar Three-Body Problem will be demonstrated in this paper by relative equilibria of the COM reduced Binary Asteroid Problem obtained by passing to center of mass (COM) coordinates, where the center of mass is with respect to the two asteroids. This COM reduction requires that the masses of the primaries to be equal, that the masses of the asteroids to be equal, and that the two asteroids to be symmetric with respect to each other through their center of mass which set at the origin (Section 2.3). As the common mass of the asteroids goes to 0 each relative equilibria of the COM reduced Binary Asteroid Problem converges to one of the five equilibria of the Generalized Copenhagen Problem (Theorem 2). In a forth-coming paper [4], we show that the spatial Binary Asteroid Problem inherits the 12 collinear relative equilibria from the general Four-Body Problem, and show that each of these relative equilibria limits to an equilibrium of the Generalized Copenhagen Problem as the masses of the two asteroids (not assumed equal) go to zero. These kinds of connections have been observed between relative equilibria in the equal mass Planar Four-Body Problem [20] and the Copenhagen Problem when the value of the mass for a pair of bodies with equal masses goes to zero [16].

The COM reduced Binary Asteroid Problem has an interpretation as a restricted four-body problem. The two primaries and a fictitious third body, placed at the origin, whose mass is that of the common mass of the two asteroids of equal mass, form a collinear central configuration. A zero-mass particle then moves in the gravitational field determined by the three collinear bodies that are moving in uniform circular motion about the origin. In rotating coordinates, the Hamiltonian of the COM reduced Binary Asteroid Problem is similar to that of the Circular Restricted Four-Body Problem with Three Equal Primaries in a Collinear Central Configuration in [11]. Since the COM reduced Binary Asteroid Problem has an interpretation as a restricted four-body problem, we will apply and/or adapt techniques used in restricted problems (e.g. [11, 2, 12, 14]) in our analysis.

Our motivations for developing the Binary Asteroid Problem consisted of following questions.

  1. (1)

    If two single asteroids (not in a binary asteroid formation) that are far away from the primaries eventually pass simultaneously close enough to one of the primaries, could the two asteroids be ejected from the close encounter as a binary asteroid?

  2. (2)

    If a binary asteroid far away from the primaries eventually passes close enough to one of the primaries, could the binary asteroid be pulled apart into two single asteroids going in separate directions?

  3. (3)

    If one asteroid is orbiting a primary and other asteroid that is far away eventually passes close enough to that primary and its orbiting asteroid, could the two asteroids be ejected as a binary asteroid?

  4. (4)

    If a binary asteroid far away from the primaries eventually passes close enough to one of the primaries, could the binary asteroid be pulled apart where one of the asteroid is captured into an orbit around that primary and the other asteroid is ejected?

  5. (5)

    Could a relative equilibrium be the alpha and/or omega limit for the two asteroids, for which the two asteroids are close together for a long time but are pulled apart as they approach the relative equilibria in backward or forward time?

  6. (6)

    Do there exist periodic solutions in a model for binary asteroids in which one of the asteroids orbits one primary and the other asteroid orbits the other primary (a Hill-type orbit)?

We leave questions (1), (2), (3), and (4) for future research, but based on numerical simulations we have realizing affirmative answers for these four questions in the Binary Asteroid Problem (see e.g. [9]), we believe these can be answered analytically through regularization. A partial answer for question (5) for the COM reduced Binary Asteroid Problem is that all the relative equilibria have stable and unstable manifolds and so provide alpha and omega limits for the binary asteroid (Section 3). In [4] we will show that the 1212 collinear relative equilibria in the Binary Asteroid Problem have stable and unstable manifolds. What we do not know, for both the Binary Asteroid Problem and the COM reduced Binary Asteroid Problem, is where those stable and unstable manifolds go in phase space and whether they intersect. An answer for question (6) in the COM reduced Binary Asteroid Problem is that there exists two families of periodic orbits in which one asteroid orbits one primary and the other asteroid orbits the other primary (Section 5).

The Binary Asteroid Problem readily extends to 33 or more primaries and 33 or more asteroids. The 33 or more primaries, with large masses, set in a central configuration, are prescribed a uniform circular motion about their center of mass, set at the origin. The 33 or more asteroids, with small but positive small masses relative to the masses of the primaries, move either in a spatial or planar setting in relation to the primaries. Some research with 33 primaries and 22 asteroids has been done where the three primaries are in an equilateral triangle with two of the masses of the primaries relatively small with respect to the the remaining primary so that the barycenter of the three primaries is close to the dominant mass primary. In this situation we numerically found relative equilibria on each of the lines that pass through the barycenter of the primaries and the primaries [3].

This paper is structured as follows. In Section 2 we derive a time-dependent Hamiltonian for the Binary Asteroid Model in a fixed coordinate system. Passing to rotating coordinates, we obtain a time-independent Hamiltonian for the Binary Asteroid Problem. Using COM coordinates we derive the Hamiltonian for the COM reduced Binary Asteroid Problem. In Section 3 we use an amended potential associated to the Hamiltonian of the COM reduced Binary Asteroid Problem, and two symmetries of the amended potential, to prove the existence of six relative equilibria, four of which are collinear with the two primaries, and two of which form convex kite configurations (and are each a rhombus) with the two primaries and have two axes of symmetries. We show how each of the these six relative equilibria limits to the corresponding relative equilibrium in the Generalized Copenhagen Problem. We also show that the four collinear relative equilibria are saddle-centers and that the two kites are hyperbolic. In Section 4 we prove that in the COM reduced Binary Asteroid Problem there exist two one-parameter families of periodic orbits near the origin (the point of collision of the two asteroids). The two asteroids on these periods orbits are close to each other and so form a binary asteroid. In Section 5 we prove in the COM reduced Binary Asteroid Problem the existence of two one-parameter families of near-circular Hill-type orbits in which one asteroid orbits one primary and the other asteroid orbits the other primary. We show numerically that these families continue when the masses of the two primaries vary slightly from being equal. In Section 6 we prove the existence of two one-parameter families of near-circular comet-type orbits in the Binary Asteroid Problem where the angle between the position vectors for the two asteroids is approximately π\pi. In Section 7 we state our conclusions and future research.

2 A Binary Asteroid Model

2.1 Initial Hamiltonian

For two bodies with masses M1,M2>0M_{1},M_{2}>0 (hereafter called primaries) having respective positions 𝔛1=(X1,Y1)\mathfrak{X}_{1}=(X_{1},Y_{1}), 𝔛2=(X2,Y2)\mathfrak{X}_{2}=(X_{2},Y_{2}) in the plane, Newton’s law of gravitation says their motion is governed by the equations

𝔛¨1=M2(𝔛1𝔛2)𝔛1𝔛23,\displaystyle\ddot{\mathfrak{X}}_{1}=-\frac{M_{2}(\mathfrak{X}_{1}-\mathfrak{X}_{2})}{\|\mathfrak{X}_{1}-\mathfrak{X}_{2}\|^{3}}, 𝔛¨2=M1(𝔛2𝔛1)𝔛1𝔛23.\displaystyle\ddot{\mathfrak{X}}_{2}=-\frac{M_{1}(\mathfrak{X}_{2}-\mathfrak{X}_{1})}{\|\mathfrak{X}_{1}-\mathfrak{X}_{2}\|^{3}}. (1)

The time unit is chosen so that the gravitation constant is 1. The well-known circular solution to (1) is given by

X1(t)=aM2Mcos(2πtP),\displaystyle X_{1}(t)=-\frac{aM_{2}}{M}\cos\left(\frac{2\pi t}{P}\right), Y1(t)=aM2Msin(2πtP),\displaystyle Y_{1}(t)=-\frac{aM_{2}}{M}\sin\left(\frac{2\pi t}{P}\right),
X2(t)=aM1Mcos(2πtP),\displaystyle X_{2}(t)=\frac{aM_{1}}{M}\cos\left(\frac{2\pi t}{P}\right), Y2(t)=aM1Msin(2πtP),\displaystyle Y_{2}(t)=\frac{aM_{1}}{M}\sin\left(\frac{2\pi t}{P}\right), (2)

where a>0a>0 is the distance between the two primaries, M=M1+M2M=M_{1}+M_{2} is the total mass of the primaries, and P>0P>0 is the period of the primaries’ circular orbit. The three parameters aa, MM, and PP are related by Kepler’s Third Law,

P2a3=4π2M.\displaystyle\frac{P^{2}}{a^{3}}=\frac{4\pi^{2}}{M}. (3)

Throughout this paper, we allow aa and MM to be arbitrary but fixed, so that PP is always determined by (3).

We introduce two more bodies with masses m1,m2>0m_{1},m_{2}>0 (hereafter called asteroids) at respective positions q1=(x1,y1)q_{1}=(x_{1},y_{1}), q2=(x2,y2)q_{2}=(x_{2},y_{2}). These asteroids have kinetic energies

K1=p122m1=12m1(px12+py12),\displaystyle K_{1}=\frac{p_{1}^{2}}{2m_{1}}=\frac{1}{2m_{1}}(p_{x_{1}}^{2}+p_{y_{1}}^{2}), K2=p222m2=12m2(px22+py12),\displaystyle K_{2}=\frac{p_{2}^{2}}{2m_{2}}=\frac{1}{2m_{2}}(p_{x_{2}}^{2}+p_{y_{1}}^{2}),

where p1=(px1,py1)p_{1}=(p_{x_{1}},p_{y_{1}}) and p2=(px2,py2)p_{2}=(p_{x_{2}},p_{y_{2}}) are the corresponding momenta. We assume that

max{m1,m2}min{M1,M2},\max\{m_{1},m_{2}\}\ll\min\{M_{1},M_{2}\},

so the asteroids do not appreciably affect the motion of the primaries but do affect each other. Therefore, to a reasonable approximation, the potential energy of asteroids is given by

U\displaystyle U =m1m2q1q2m1M1𝔛1(t)q1m2M1𝔛1(t)q2\displaystyle=-\frac{m_{1}m_{2}}{\|q_{1}-q_{2}\|}-\frac{m_{1}M_{1}}{\|\mathfrak{X}_{1}(t)-q_{1}\|}-\frac{m_{2}M_{1}}{\|\mathfrak{X}_{1}(t)-q_{2}\|}
m1M2𝔛2(t)q1m2M2𝔛2(t)q2.\displaystyle\ \ \ \ -\frac{m_{1}M_{2}}{\|\mathfrak{X}_{2}(t)-q_{1}\|}-\frac{m_{2}M_{2}}{\|\mathfrak{X}_{2}(t)-q_{2}\|}.

The time-dependent Hamiltonian for this binary asteroid model is

H(qi,pi,t)\displaystyle H(q_{i},p_{i},t) =K1+K2+U.\displaystyle=K_{1}+K_{2}+U. (4)

2.2 Rotating Coordinates

To eliminate the time dependence in (4), we use the rotation matrix

exp(2πtKP)=[cos(2πt/P)sin(2πt/P)sin(2πt/P)cos(2πt/P)], where K=[0110].\exp\left(\frac{2\pi tK}{P}\right)=\begin{bmatrix}\cos(2\pi t/P)&\sin(2\pi t/P)\\ -\sin(2\pi t/P)&\cos(2\pi t/P)\end{bmatrix},\text{ where }K=\begin{bmatrix}0&1\\ -1&0\end{bmatrix}.

We make the linear symplectic transformation

qiexp(2πtKP)qi,\displaystyle q_{i}\mapsto\exp\left(\frac{2\pi tK}{P}\right)q_{i}, piexp(2πtKP)pi.\displaystyle p_{i}\mapsto\exp\left(\frac{2\pi tK}{P}\right)p_{i}.

The associated remainder in the rotating coordinates is

R(qi,pi)=2πP(q1TKp1+q2TKp2).R(q_{i},p_{i})=-\frac{2\pi}{P}\left(q_{1}^{T}Kp_{1}+q_{2}^{T}Kp_{2}\right).

Since exp(2πtK/P)\exp\left(2\pi tK/P\right) is an orthogonal matrix, the norms in the denominators of the potential term are unchanged when we apply the rotation to the distance vectors. The Hamiltonian in the new coordinates is now the time-independent

HBA\displaystyle H_{BA} =p122m1+p222m22πP(q1TKp1+q2TKp2)m1m2q1q2\displaystyle=\frac{p_{1}^{2}}{2m_{1}}+\frac{p_{2}^{2}}{2m_{2}}-\frac{2\pi}{P}\left(q_{1}^{T}Kp_{1}+q_{2}^{T}Kp_{2}\right)-\frac{m_{1}m_{2}}{\|q_{1}-q_{2}\|}
m1M1q1+(aM2/M,0)m2M1q2+(aM2/M,0)\displaystyle\ \ \ \ -\frac{m_{1}M_{1}}{\|q_{1}+(aM_{2}/M,0)\|}-\frac{m_{2}M_{1}}{\|q_{2}+(aM_{2}/M,0)\|}
m1M2q1(aM1/M,0)m2M2q2(aM1/M,0).\displaystyle\ \ \ \ -\frac{m_{1}M_{2}}{\|q_{1}-(aM_{1}/M,0)\|}-\frac{m_{2}M_{2}}{\|q_{2}-(aM_{1}/M,0)\|}. (5)

In terms of the components of the positions and momenta, (5) becomes

HBA\displaystyle H_{BA} =12m1(px12+py12)+12m2(px22+py22)2πP(x1py1y1px1)\displaystyle=\frac{1}{2m_{1}}\left(p_{x_{1}}^{2}+p_{y_{1}}^{2}\right)+\frac{1}{2m_{2}}\left(p_{x_{2}}^{2}+p_{y_{2}}^{2}\right)-\frac{2\pi}{P}\left(x_{1}p_{y_{1}}-y_{1}p_{x_{1}}\right)
2πP(x2py2y2px2)m1m2(x1x2)2+(y1y2)2\displaystyle\ \ \ \ -\frac{2\pi}{P}\left(x_{2}p_{y_{2}}-y_{2}p_{x_{2}}\right)-\frac{m_{1}m_{2}}{\sqrt{(x_{1}-x_{2})^{2}+(y_{1}-y_{2})^{2}}}
m1M1(x1+aM2/M)2+y12m2M1(x2+aM2/M)2+y22\displaystyle\ \ \ \ -\frac{m_{1}M_{1}}{\sqrt{(x_{1}+aM_{2}/M)^{2}+y_{1}^{2}}}-\frac{m_{2}M_{1}}{\sqrt{(x_{2}+aM_{2}/M)^{2}+y_{2}^{2}}}
m1M2(x1aM1/M)2+y12m2M2(x2aM1/M)2+y22.\displaystyle\ \ \ \ -\frac{m_{1}M_{2}}{\sqrt{(x_{1}-aM_{1}/M)^{2}+y_{1}^{2}}}-\frac{m_{2}M_{2}}{\sqrt{(x_{2}-aM_{1}/M)^{2}+y_{2}^{2}}}. (6)

The Hamiltonian HBAH_{BA} in (6) is called the (planar) Binary Asteroid Hamiltonian and the associated system of ODEs is called the (planar) Binary Asteroid Problem.

2.3 COM Coordinates

We will reduce the number of degrees of freedom in (6) from 44 to 22 by passing to symmetric configurations for the binary asteroids under the assumption of equal masses

m1=m2=mm_{1}=m_{2}=m

for the binary asteroids, and under the assumption of equal masses

M1=M2=M2M_{1}=M_{2}=\frac{M}{2}

for the primaries. The quantity

ν=M1M22\nu=\frac{M_{1}-M_{2}}{2}

measures the deviation between the masses of the primaries and will be use as a perturbation parameter for an expansion of the Hamiltonian. With the matrix

=12[1010010110100101]\mathcal{M}=\frac{1}{2}\begin{bmatrix}1&0&-1&0\\ 0&1&0&-1\\ 1&0&1&0\\ 0&1&0&1\end{bmatrix}

we define the linear symplectic change of coordinates

(x1,y1,x2,y2)(x,y,α,β),\displaystyle(x_{1},y_{1},x_{2},y_{2})\mapsto(x,y,\alpha,\beta), (px1,py1,px2,py2)(px,py,pα,pβ),\displaystyle(p_{x_{1}},p_{y_{1}},p_{x_{2}},p_{y_{2}})\mapsto(p_{x},p_{y},p_{\alpha},p_{\beta}),

by

(x,y,α,β)T\displaystyle(x,y,\alpha,\beta)^{T} =(x1,y1,x2,y2)T,\displaystyle=\mathcal{M}(x_{1},y_{1},x_{2},y_{2})^{T},
(px,py,pα,pβ)T\displaystyle(p_{x},p_{y},p_{\alpha},p_{\beta})^{T} =(T)1(px1,py1,px2,py2)T.\displaystyle=(\mathcal{M}^{T})^{-1}(p_{x_{1}},p_{y_{1}},p_{x_{2}},p_{y_{2}})^{T}.

Explicitly, the change of coordinates is

x=x1x22,y=y1y22,α=x1+x22,β=y1+y22,\displaystyle x=\frac{x_{1}-x_{2}}{2},\ y=\frac{y_{1}-y_{2}}{2},\ \alpha=\frac{x_{1}+x_{2}}{2},\ \beta=\frac{y_{1}+y_{2}}{2}, (7)
px=px1px2,py=py1py2,pα=px1+px2,pβ=py1+py2.\displaystyle p_{x}=p_{x_{1}}-p_{x_{2}},\ p_{y}=p_{y_{1}}-p_{y_{2}},\ p_{\alpha}=p_{x_{1}}+p_{x_{2}},\ p_{\beta}=p_{y_{1}}+p_{y_{2}}.

We call the new coordinates the COM coordinates (where COM means “Center Of Mass”). Physically, α\alpha and β\beta are the components of the asteroids’ barycenter, while xx and yy are the components of half of the position of the asteroid with mass m1m_{1} relative to asteroid with mass m2m_{2}. The change of coordinates in (7) is similar to Jacobi coordinates for the two-body problem.

In the COM coordinates the Hamiltonian HBAH_{BA} is analytic in the parameter ν\nu. By expanding HBAH_{BA} in a Taylor series in ν\nu about ν=0\nu=0, we obtain the Hamiltonian

HCOM\displaystyle H_{COM} =14m(px2+py2)+14m(pα2+pβ2)\displaystyle=\frac{1}{4m}\left(p_{x}^{2}+p_{y}^{2}\right)+\frac{1}{4m}\left(p_{\alpha}^{2}+p_{\beta}^{2}\right) (8)
2πP(xpyypx+αpββpα)m22x2+y2\displaystyle\ \ \ \ -\frac{2\pi}{P}\left(xp_{y}-yp_{x}+\alpha p_{\beta}-\beta p_{\alpha}\right)-\frac{m^{2}}{2\sqrt{x^{2}+y^{2}}}
mM/2(x+α+a2)2+(y+β)2mM/2(xαa2)2+(yβ)2\displaystyle\ \ \ \ -\frac{mM/2}{\sqrt{(x+\alpha+\frac{a}{2})^{2}+(y+\beta)^{2}}}-\frac{mM/2}{\sqrt{(x-\alpha-\frac{a}{2})^{2}+(y-\beta)^{2}}}
mM/2(x+αa2)2+(y+β)2mM/2(xα+a2)2+(yβ)2\displaystyle\ \ \ \ -\frac{mM/2}{\sqrt{(x+\alpha-\frac{a}{2})^{2}+(y+\beta)^{2}}}-\frac{mM/2}{\sqrt{(x-\alpha+\frac{a}{2})^{2}+(y-\beta)^{2}}}
+O(ν).\displaystyle\ \ \ \ +O(\nu).

We consider the problem when ν=0\nu=0. We will comment on the problem when ν0\nu\neq 0 in Subsection 5.4.

2.4 Reduction

The equations of motion in the COM coordinates (α,β,pα,pβ)(\alpha,\beta,p_{\alpha},p_{\beta}) defined by (8) are

α˙\displaystyle\dot{\alpha} =pα2m+2πβP\displaystyle=\frac{p_{\alpha}}{2m}+\frac{2\pi\beta}{P} β˙\displaystyle\dot{\beta} =pβ2m2παP\displaystyle=\frac{p_{\beta}}{2m}-\frac{2\pi\alpha}{P}
p˙α\displaystyle\dot{p}_{\alpha} =2πpβPmM(x+α+a/2)/2((x+α+a2)2+(y+β)2)3/2+mM(xαa/2)/2((xαa2)2+(yβ)2)3/2\displaystyle=\frac{2\pi p_{\beta}}{P}-\frac{mM(x+\alpha+a/2)/2}{({(x+\alpha+\frac{a}{2})^{2}+(y+\beta)^{2}})^{3/2}}+\frac{mM(x-\alpha-a/2)/2}{((x-\alpha-\frac{a}{2})^{2}+(y-\beta)^{2})^{3/2}}
mM(x+αa/2)/2((x+αa2)2+(y+β)2)3/2+mM(xα+a/2)/2((xα+a2)2+(yβ)2)3/2\displaystyle\ \ \ \ -\frac{mM(x+\alpha-a/2)/2}{((x+\alpha-\frac{a}{2})^{2}+(y+\beta)^{2})^{3/2}}+\frac{mM(x-\alpha+a/2)/2}{((x-\alpha+\frac{a}{2})^{2}+(y-\beta)^{2})^{3/2}}
p˙β\displaystyle\dot{p}_{\beta} =2πpαPmM(y+β)/2((x+α+a2)2+(y+β)2)3/2+mM(yβ)/2((xαa2)2+(yβ)2)3/2\displaystyle=-\frac{2\pi p_{\alpha}}{P}-\frac{mM(y+\beta)/2}{((x+\alpha+\frac{a}{2})^{2}+(y+\beta)^{2})^{3/2}}+\frac{mM(y-\beta)/2}{((x-\alpha-\frac{a}{2})^{2}+(y-\beta)^{2})^{3/2}}
mM(y+β)/2((x+αa2)2+(y+β)2)3/2+mM(yβ)/2((xα+a2)2+(yβ)2)3/2.\displaystyle\ \ \ \ -\frac{mM(y+\beta)/2}{((x+\alpha-\frac{a}{2})^{2}+(y+\beta)^{2})^{3/2}}+\frac{mM(y-\beta)/2}{((x-\alpha+\frac{a}{2})^{2}+(y-\beta)^{2})^{3/2}}. (9)

The zero functions α=β=pα=pβ=0\alpha=\beta=p_{\alpha}=p_{\beta}=0 satisfy (9). Setting α=β=pα=pβ=0\alpha=\beta=p_{\alpha}=p_{\beta}=0 corresponds to the two asteroids having a fixed barycenter at the origin, where the position and the momentum of the asteroid of mass m1=mm_{1}=m satisfy

(x1,y1)=(x,y),(px1,py1)=12(px,py),(x_{1},y_{1})=(x,y),\ (p_{x_{1}},p_{y_{1}})=\frac{1}{2}(p_{x},p_{y}),

and the position and the momentum of asteroid of mass m2=mm_{2}=m satisfy

(x2,y2)=(x,y),(px2,py2)=12(px,py).(x_{2},y_{2})=-(x,y),\ (p_{x_{2}},p_{y_{2}})=-\frac{1}{2}(p_{x},p_{y}).

Thus, the two asteroids move symmetrically opposite each other through the origin, where the coordinates (x,y,px,py)(x,y,p_{x},p_{y}) describe the motion of the asteroid with mass m1=mm_{1}=m. By the symmetry, the motion of the asteroid with mass m2=mm_{2}=m is described by the coordinates (x,y,px,py)(-x,-y,-p_{x},-p_{y}).

The Hamiltonian for motion of the asteroid with mass m=m1m=m_{1} is obtained by substitution of solution α=0\alpha=0, β=0\beta=0, pα=0p_{\alpha}=0, and pβ=0p_{\beta}=0 in the Hamiltonian HCOMH_{COM}. This gives the Hamiltonian

Hxy(x,y,px,py)\displaystyle H_{xy}(x,y,p_{x},p_{y}) =14m(px2+py2)2πP(xpyypx)m22x2+y2\displaystyle=\frac{1}{4m}\left(p_{x}^{2}+p_{y}^{2}\right)-\frac{2\pi}{P}\left(xp_{y}-yp_{x}\right)-\frac{m^{2}}{2\sqrt{x^{2}+y^{2}}}
mM(x+a/2)2+y2mM(xa/2)2+y2.\displaystyle\ \ \ \ -\frac{mM}{\sqrt{(x+a/2)^{2}+y^{2}}}-\frac{mM}{\sqrt{(x-a/2)^{2}+y^{2}}}. (10)

It is straight-forward to verify that the four equations of the Hamiltonian system associated with HxyH_{xy} agree with the equations x˙=HCOM/px\dot{x}=\partial H_{\rm COM}/\partial p_{x}, y˙=HCOM/py\dot{y}=\partial H_{\rm COM}/\partial p_{y}, p˙x=HCOM/x\dot{p}_{x}=-\partial H_{\rm COM}/\partial x, and p˙y=HCOM/y\dot{p}_{y}=-\partial H_{\rm COM}/\partial y with α=β=pα=pβ=0\alpha=\beta=p_{\alpha}=p_{\beta}=0 substituted into them. This completes the reduction to symmetric configurations when m1=m2=mm_{1}=m_{2}=m and M1=M2=M/2M_{1}=M_{2}=M/2 and gives the following result.

Theorem 1 (COM Reduction).

Each equilibrium or TT-periodic solution

(x(t),y(t),px(t),py(t))(x(t),y(t),p_{x}(t),p_{y}(t))

of the Hamiltonian system for HxyH_{xy} in (10) defines an equilibrium or TT-periodic solution of the Hamiltonian system for HBAH_{BA} in (6) ((the planar Binary Asteroid Problem)) when m1=m2=mm_{1}=m_{2}=m and M1=M2=M/2M_{1}=M_{2}=M/2, given in the COM coordinates (7) by

(x(t),y(t),0,0,px(t),py(t),0,0).(x(t),y(t),0,0,p_{x}(t),p_{y}(t),0,0).
Remark 1.

The form of the Hamiltonian HxyH_{xy} in (10) is similar to that in [11]. The benefit of the COM reduction is that we may apply analysis typical of restricted problems, despite that the binary asteroid problem is not a restricted problem.

3 Relative Equilibria

We determine which of the relative equilibria in the planar Four-Body Problem [20] the COM reduced Binary Asteroid Problem inherits. By an analysis of the amended potential associated to the COM reduced Binary Asteroid Problem, we prove there are exactly six relative equilibria of HxyH_{xy}. Four of these six relative equilibria determine two distinct symmetric collinear configurations of the four bodies. The remaining two relative equilibria determine one doubly-symmetric convex kite configuration of the four bodies, which is a rhombus. A stability analysis shows that one of the two symmetric collinear configurations is always a saddle center, independent of the relationship between mm and MM, while the remaining symmetric collinear configuration is a saddle center when mMm\ll M. A stability analysis of the doubly symmetric convex kite configuration shows that it is hyperbolic when mMm\ll M.

3.1 Amended Potential; Symmetries

The Hamiltonian HxyH_{xy} has three parts: the kinetic term,

K(px,py)=14m(px2+py2),K(p_{x},p_{y})=\frac{1}{4m}\left(p_{x}^{2}+p_{y}^{2}\right),

the Coriolis term,

Ω(x,y,px,py)=2πP(xpyypx)=2πP(ypxxpy),\Omega(x,y,p_{x},p_{y})=-\frac{2\pi}{P}(xp_{y}-yp_{x})=\frac{2\pi}{P}(yp_{x}-xp_{y}),

and the potential term

U(x,y)=m22x2+y2+mM(x+a/2)2+y2+mM(xa/2)2+y2.U(x,y)=\frac{m^{2}}{2\sqrt{x^{2}+y^{2}}}+\frac{mM}{\sqrt{(x+a/2)^{2}+y^{2}}}+\frac{mM}{\sqrt{(x-a/2)^{2}+y^{2}}}.

Thus we can write

Hxy=K(px,py)+Ω(x,y,px,py)U(x,y).H_{xy}=K(p_{x},p_{y})+\Omega(x,y,p_{x},p_{y})-U(x,y).

The Hamiltonian equations associated with HxyH_{xy} then take the form

x˙\displaystyle\dot{x} =Kpx+Ωpx=px2m+2πyP,\displaystyle=\frac{\partial K}{\partial p_{x}}+\frac{\partial\Omega}{\partial p_{x}}=\frac{p_{x}}{2m}+\frac{2\pi y}{P}, (11)
y˙\displaystyle\dot{y} =Kpy+Ωpy=py2m2πxP,\displaystyle=\frac{\partial K}{\partial p_{y}}+\frac{\partial\Omega}{\partial p_{y}}=\frac{p_{y}}{2m}-\frac{2\pi x}{P},
p˙x\displaystyle\dot{p}_{x} =Ωx+Ux=2πpyP+Ux,\displaystyle=-\frac{\partial\Omega}{\partial x}+\frac{\partial U}{\partial x}=\frac{2\pi p_{y}}{P}+\frac{\partial U}{\partial x},
p˙y\displaystyle\dot{p}_{y} =Ωy+Uy=2πpxP+Uy.\displaystyle=-\frac{\partial\Omega}{\partial y}+\frac{\partial U}{\partial y}=-\frac{2\pi p_{x}}{P}+\frac{\partial U}{\partial y}.

The Newtonian equations of the motion (the system of second-order ODEs) is obtained by eliminating the momenta from the first order system of ODEs in (11). For this we have

x¨\displaystyle\ddot{x} =4πy˙P+4π2xP2+12mUx,\displaystyle=\frac{4\pi\dot{y}}{P}+\frac{4\pi^{2}x}{P^{2}}+\frac{1}{2m}\frac{\partial U}{\partial x}, (12)
y¨\displaystyle\ddot{y} =4πx˙P+4π2yP2+12mUy.\displaystyle=-\frac{4\pi\dot{x}}{P}+\frac{4\pi^{2}y}{P^{2}}+\frac{1}{2m}\frac{\partial U}{\partial y}.

Define the amended potential

V(x,y)=2π2P2(x2+y2)+12mU(x,y).V(x,y)=\frac{2\pi^{2}}{P^{2}}(x^{2}+y^{2})+\frac{1}{2m}U(x,y). (13)

The second-order system of ODEs in (12) have the form

x¨4πy˙P\displaystyle\ddot{x}-\frac{4\pi\dot{y}}{P} =4π2xP2+12mUx=Vx,\displaystyle=\frac{4\pi^{2}x}{P^{2}}+\frac{1}{2m}\frac{\partial U}{\partial x}=\frac{\partial V}{\partial x}, (14)
y¨+4πx˙P\displaystyle\ddot{y}+\frac{4\pi\dot{x}}{P} =4π2yP2+12mUy=Vy.\displaystyle=\frac{4\pi^{2}y}{P^{2}}+\frac{1}{2m}\frac{\partial U}{\partial y}=\frac{\partial V}{\partial y}.

The Jacobi integral for the second-order system (14) is

C=2V(x˙2+y˙2).C=2V-(\dot{x}^{2}+\dot{y}^{2}).

Since by Kepler’s Third Law (3) there holds

2π2P2=12(Ma3),\frac{2\pi^{2}}{P^{2}}=\frac{1}{2}\left(\frac{M}{a^{3}}\right),

the amended potential becomes

V(x,y)\displaystyle V(x,y) =M2a3(x2+y2)+m4x2+y2\displaystyle=\frac{M}{2a^{3}}(x^{2}+y^{2})+\frac{m}{4\sqrt{x^{2}+y^{2}}}
+M2(x+a/2)2+y2+M2(xa/2)2+y2.\displaystyle\ \ \ \ +\frac{M}{2\sqrt{(x+a/2)^{2}+y^{2}}}+\frac{M}{2\sqrt{(x-a/2)^{2}+y^{2}}}.

The amended potential is expressed solely in terms of the three parameters mm, MM, and aa. The amended potential VV has the symmetries

VR1=V,VR2=V,V\circ R_{1}=V,\ V\circ R_{2}=V, (15)

for the involutions

R1(x,y)=(x,y),R2(x,y)=(x,y).R_{1}(x,y)=(-x,y),\ R_{2}(x,y)=(x,-y).

See Figure 1 for a graph of amended potential. See Figure 2 for the zero velocity curves V=C/2V=C/2 associated with the amended potential. For both figures the value of mm is exaggerated, possibly beyond mMm\ll M, to overcome the graphing resolution issue associated with the singularity of VV at the origin.

Refer to caption
Figure 1: Graph of amended potential function VV when m=0.5m=0.5, M=10M=10, and a=2a=2.
Refer to caption
Figure 2: Zero velocity curves V=C/2V=C/2 when m=0.5m=0.5, M=10M=10, and a=2a=2, where in the legend on the right is a selection of values of C/2.C/2.

The Hamiltonian HxyH_{xy} has the symmetries

HxyR^1=Hxy,HxyR^2=Hxy,H_{xy}\circ\hat{R}_{1}=H_{xy},\ H_{xy}\circ\hat{R}_{2}=H_{xy},

for the involutions

R^1(x,y,px,py)=(x,y,px,py),R^2(x,y,px,py)=(x,y,px,py),\hat{R}_{1}(x,y,p_{x},p_{y})=(-x,y,p_{x},-p_{y}),\ \hat{R}_{2}(x,y,p_{x},p_{y})=(x,-y,-p_{x},p_{y}),

where R^i\hat{R}_{i} is an extension of RiR_{i}, i=1,2i=1,2. Associated to each R^i\hat{R}_{i} is a time-reversing symmetry: if (x(t),y(t),px(t),py(t))(x(t),y(t),p_{x}(t),p_{y}(t)) is a solution of system of equations (11) then so is

R^i(x(t),y(t),px(t),py(t)),i=1,2.\hat{R}_{i}(x(-t),y(-t),p_{x}(-t),p_{y}(-t)),\ i=1,2. (16)

3.2 Critical Points

Equilibria of the second-order system in (14) are given by the critical points of the amended potential VV. The partial derivatives of VV are

Vx\displaystyle\frac{\partial V}{\partial x} =Mxa3mx4(x2+y2)3/2M(x+a/2)2((x+a/2)2+y2)3/2\displaystyle=\frac{Mx}{a^{3}}-\frac{mx}{4(x^{2}+y^{2})^{3/2}}-\frac{M(x+a/2)}{2((x+a/2)^{2}+y^{2})^{3/2}} (17)
M(xa/2)2((xa/2)2+y2)3/2,\displaystyle\ \ \ \ -\frac{M(x-a/2)}{2((x-a/2)^{2}+y^{2})^{3/2}},
Vy\displaystyle\frac{\partial V}{\partial y} =y[Ma3m4(x2+y2)3/2M2((x+a/2)2+y2)3/2\displaystyle=y\bigg{[}\frac{M}{a^{3}}-\frac{m}{4(x^{2}+y^{2})^{3/2}}-\frac{M}{2((x+a/2)^{2}+y^{2})^{3/2}}
M2((xa/2)2+y2)3/2].\displaystyle\ \ \ \ \ \ -\frac{M}{2((x-a/2)^{2}+y^{2})^{3/2}}\bigg{]}.

Setting m=0m=0, M=1/2M=1/2, and a=1a=1 in (17) gives precisely the equations that determine the five equilibria in the Circular Restricted Three-Body Problem with equal primary masses, otherwise known as the classical Copenhagen Problem [14, 21]. For m=0m=0, M>0M>0, and a>0a>0 in (17), we refer to the five critical points

(1,0),(±2,0),(0,±4)({\mathcal{L}}_{1},0),(\pm{\mathcal{L}}_{2},0),(0,\pm{\mathcal{L}}_{4})

of VV as the equilibria of the Generalized Copenhagen Problem. Here 1=0{\mathcal{L}}_{1}=0, 2>a/2{\mathcal{L}}_{2}>a/2, and 4=3a/2{\mathcal{L}}_{4}=\sqrt{3}a/2. We will prove the existence of six equilibria

(±L1,0),(±L2,0),(0,±L4)(\pm L_{1},0),(\pm L_{2},0),(0,\pm L_{4})

of the binary asteroid problem (m>0m>0). The four equilibria (±L1,0)(\pm L_{1},0) and (±L2,0)(\pm L_{2},0) form collinear symmetric configurations and the two configurations (0,±L4)(0,\pm L_{4}) form a doubly symmetric convex kite (and rhombus) configuration. We relate the six equilibria of the binary asteroid problem to the five equilibria of the Generalized Copenhagen Problem as m0m\to 0.

The partial derivatives of VV imply there are two cases in which to search for critical points. Setting the partial derivative V/y\partial V/\partial y to zero gives y=0y=0 and

0=Ma3m4(x2+y2)3/2M2((x+a/2)2+y2)3/2M2((xa/2)2+y2)3/2.0=\frac{M}{a^{3}}-\frac{m}{4(x^{2}+y^{2})^{3/2}}-\frac{M}{2((x+a/2)^{2}+y^{2})^{3/2}}-\frac{M}{2((x-a/2)^{2}+y^{2})^{3/2}}. (18)

Substitution of y=0y=0 into V/x=0\partial V/\partial x=0 gives the first case,

0=Mxa3mx4|x|3M(x+a/2)2|x+a/2|3M(xa/2)2|xa/2|3.0=\frac{Mx}{a^{3}}-\frac{mx}{4|x|^{3}}-\frac{M(x+a/2)}{2|x+a/2|^{3}}-\frac{M(x-a/2)}{2|x-a/2|^{3}}. (19)

To obtain the second case, we multiply (18) by x-x and add it to V/x=0\partial V/\partial x=0. This gives an equation that implies x=0x=0. Substitution of x=0x=0 into (18) gives the second case,

0=Ma3m4|y|3M((a/2)2+y2)3/2.0=\frac{M}{a^{3}}-\frac{m}{4|y|^{3}}-\frac{M}{((a/2)^{2}+y^{2})^{3/2}}. (20)

Further restrictions on the domains in the two cases follow from the symmetries of VV given in (15) . We need only search for critical points in first case on x>0x>0, and in the second case on y>0y>0.

Theorem 2.

There exist 66 critical points (±L1,0)(\pm L_{1},0), (±L2,0)(\pm L_{2},0), and (0,±L4)(0,\pm L_{4}) of VV that are smoothly dependent on (m,M,a)(+)3(m,M,a)\in({\mathbb{R}}^{+})^{3} where L1(0,a/2)L_{1}\in(0,a/2), L2(a/2,)L_{2}\in(a/2,\infty), and L4(3a/2,)L_{4}\in(\sqrt{3}a/2,\infty). For fixed (M,a)(+)2(M,a)\in({\mathbb{R}}^{+})^{2} there holds L11L_{1}\to{\mathcal{L}}_{1}, L22L_{2}\to{\mathcal{L}}_{2}, and L44=3a/2L_{4}\to{\mathcal{L}_{4}}=\sqrt{3}a/2 as m0+m\to 0^{+}, and L4L_{4} is an increasing function of mm with L4L_{4}\to\infty as mm\to\infty. Furthermore, for fixed (M,a)(+)2(M,a)\in({\mathbb{R}}^{+})^{2}, the expansion of L4L_{4} in terms of mm near m=0m=0 is

L4(m)=3a2+4a27Mm+O(m2).L_{4}(m)=\frac{\sqrt{3}a}{2}+\frac{4a}{27M}m+O(m^{2}).
Proof.

We start with the first case of critical points of the form (x,0)(x,0), x>0x>0. Setting y=0y=0, the amended potential is

xV(x,0)=M2a3x2+m4x+M2(x+a/2)+M2|xa/2|,x>0.x\mapsto V(x,0)=\frac{M}{2a^{3}}x^{2}+\frac{m}{4x}+\frac{M}{2(x+a/2)}+\frac{M}{2|x-a/2|},\ x>0.

We use coercivity and convexity of xV(x,0)x\mapsto V(x,0), as is done in the Circular Restricted Three-Body Problem, to show that there is a unique critical point of V(x,0)V(x,0) in each of the intervals

(0,a/2),(a/2,).(0,a/2),(a/2,\infty).

The limits

limx0+V(x,0)=,limxa/2V(x,0)=,\displaystyle\lim_{x\to 0^{+}}V(x,0)=\infty,\ \lim_{x\to{a/2}^{-}}V(x,0)=\infty,
limxa/2+V(x,0)=,limxV(x,0)=,\displaystyle\lim_{x\to{a/2}^{+}}V(x,0)=\infty,\ \lim_{x\to\infty}V(x,0)=\infty,

imply the coercivity of V(x,0)V(x,0) on the intervals (0,a/2)(0,a/2) and (a/2,)(a/2,\infty). The second derivative of VV with respect to xx evaluated at (x,0)(x,0) for x(0,a/2)(a/2,)x\in(0,a/2)\cup(a/2,\infty) is

2Vx2(x,0)=Ma3+m2x3+M(x+a/2)3+M|xa/2|3>0.\frac{\partial^{2}V}{\partial x^{2}}(x,0)=\frac{M}{a^{3}}+\frac{m}{2x^{3}}+\frac{M}{(x+a/2)^{3}}+\frac{M}{|x-a/2|^{3}}>0. (21)

Thus xV(x,0)x\mapsto V(x,0) is convex on each of the intervals (0,a/2)(0,a/2) and (a/2,)(a/2,\infty). The unique critical point L1(0,a/2)L_{1}\in(0,a/2) of xV(x,0)x\mapsto V(x,0) satisfies (19), i.e.,

0=Vx(L1,0)=ML1a3m4L12M2(L1+a/2)2+M2(L1a/2)2,0=\frac{\partial V}{\partial x}(L_{1},0)=\frac{ML_{1}}{a^{3}}-\frac{m}{4L_{1}^{2}}-\frac{M}{2(L_{1}+a/2)^{2}}+\frac{M}{2(L_{1}-a/2)^{2}}, (22)

and the unique critical point L2(a/2,)L_{2}\in(a/2,\infty) of xV(x,0)x\mapsto V(x,0) satisfies (19), i.e,

0=Vx(L2,0)=ML2a3m4L22M2(L2+a/2)2M2(L2a/2)2.0=\frac{\partial V}{\partial x}(L_{2},0)=\frac{ML_{2}}{a^{3}}-\frac{m}{4L_{2}^{2}}-\frac{M}{2(L_{2}+a/2)^{2}}-\frac{M}{2(L_{2}-a/2)^{2}}. (23)

We show that L1L_{1} is a smooth function of (m,M,a)(+)3(m,M,a)\in({\mathbb{R}}^{+})^{3}, and for each fixed M>0M>0 and a>0a>0, as m0m\to 0, that L1L_{1} approaches the equilibrium 1=0{\mathcal{L}}_{1}=0 of the Generalized Copenhagen Problem. For x(0,a/2)x\in(0,a/2) set

F(x,m,M,a)=Vx(x,0)=Mxa3m4x2M2(x+a/2)2+M2(xa/2)2.F(x,m,M,a)=\frac{\partial V}{\partial x}(x,0)=\frac{Mx}{a^{3}}-\frac{m}{4x^{2}}-\frac{M}{2(x+a/2)^{2}}+\frac{M}{2(x-a/2)^{2}}.

Since F(L1,m,M,a)=0F(L_{1},m,M,a)=0 by (22) and

Fx(L1,0)=2Vx2(L1,0)>0,\frac{\partial F}{\partial x}(L_{1},0)=\frac{\partial^{2}V}{\partial x^{2}}(L_{1},0)>0,

it follows by the Implicit Function Theorem that L1L_{1} is a smooth function of (m,M,a)(m,M,a) on (+)3({\mathbb{R}}^{+})^{3}. Fixing the values of M>0M>0 and a>0a>0 shows that L1L_{1} is a smooth function mm on the interval m>0m>0. We cannot apply the Implicit Function Theorem to FF because it is not defined when x=0x=0. Instead we split FF, i.e., V/x(x,0)\partial V/\partial x(x,0), into two functions,

f(x,M,a)=Mxa3M2(x+a/2)2+M2(xa/2)2f(x,M,a)=\frac{Mx}{a^{3}}-\frac{M}{2(x+a/2)^{2}}+\frac{M}{2(x-a/2)^{2}}

and

g(x,m)=m4x2,g(x,m)=\frac{m}{4x^{2}},

so that

F(x,m,M,a)=f(x,M,a)g(x,m).F(x,m,M,a)=f(x,M,a)-g(x,m).

For fixed M>0M>0 and a>0a>0, the function ff has the limits

limx0+f(x,M,a)=0,limxa/2f(x,M,a)=.\lim_{x\to 0^{+}}f(x,M,a)=0,\ \lim_{x\to a/2^{-}}f(x,M,a)=\infty.

Since

fx=Ma3+M(x+a/2)3M(xa/2)3\frac{\partial f}{\partial x}=\frac{M}{a^{3}}+\frac{M}{(x+a/2)^{3}}-\frac{M}{(x-a/2)^{3}}

is positive on the interval x(0,a/2)x\in(0,a/2), it follows that the function ff is positive and increasing on x(0,a/2)x\in(0,a/2). The function gg is positive and decreasing on x(0,a/2)x\in(0,a/2). The graphs of ff and gg intersect at precisely one point, whose xx value is L1L_{1}. As m0+m\to 0^{+} the graph of ff remains fixed while the graph of gg shifts the unique intersection of the graphs of ff and gg towards the origin, i.e., L10+L_{1}\to 0^{+}. Setting m=0m=0 in FF gives

F(x,0,M,a)=Mxa3M2(x+a/2)2+M2(xa/2)2,F(x,0,M,a)=\frac{Mx}{a^{3}}-\frac{M}{2(x+a/2)^{2}}+\frac{M}{2(x-a/2)^{2}},

a solution of which is x=0x=0, the equilibrium 1{\mathcal{L}}_{1} of the Generalized Copenhagen Problem. Thus, the equilibrium L1L_{1} converges to the equilibrium 1{\mathcal{L}}_{1} of the Generalized Copenhagen Problem when m0+m\to 0^{+}.

We show that L2L_{2} is a smooth function of (m,M,a)(+)3(m,M,a)\in({\mathbb{R}}^{+})^{3}, and for fixed M>0M>0 and a>0a>0, that as m0+m\to 0^{+}, L2L_{2} approaches the equilibrium 2{\mathcal{L}}_{2} of the Generalized Copenhagen Problem. For x(a/2,)x\in(a/2,\infty) now set

F(x,m,M,a)=Vx(x,0)=Mxa3m4x2M2(x+a/2)2M2(xa/2)2.F(x,m,M,a)=\frac{\partial V}{\partial x}(x,0)=\frac{Mx}{a^{3}}-\frac{m}{4x^{2}}-\frac{M}{2(x+a/2)^{2}}-\frac{M}{2(x-a/2)^{2}}.

Since F(L2,m,M,a)=0F(L_{2},m,M,a)=0 from (23) and

Fx(L2,0)=2Vx2(L2,0)>0\frac{\partial F}{\partial x}(L_{2},0)=\frac{\partial^{2}V}{\partial x^{2}}(L_{2},0)>0

it follows from the Implicit Function Theorem that L2L_{2} is a smooth function of (m,M,a)(m,M,a) on (+)3({\mathbb{R}}^{+})^{3}. Fixing the values of M>0M>0 and a>0a>0 shows that L2L_{2} is a smooth function of mm on the interval m>0m>0. Setting m=0m=0 in F(x,m,M,a)=0F(x,m,M,a)=0 gives

0=Mxa3M2(x+a/2)2M2(xa/2)20=\frac{Mx}{a^{3}}-\frac{M}{2(x+a/2)^{2}}-\frac{M}{2(x-a/2)^{2}}

for which the unique solution x>a/2x>a/2 is the equilibrium 2{\mathcal{L}}_{2} of the generalized Copenhagen problem. Since

Fx(2,0,M,a)=Ma3+M(2+a/2)3+M(2a/2)3>0,\frac{\partial F}{\partial x}({\mathcal{L}}_{2},0,M,a)=\frac{M}{a^{3}}+\frac{M}{({\mathcal{L}}_{2}+a/2)^{3}}+\frac{M}{({\mathcal{L}}_{2}-a/2)^{3}}>0,

the uniqueness part of the Implicit Function Theorem implies that 2{\mathcal{L}}_{2} is the limit of L2L_{2} as m0+m\to 0^{+}.

Now we treat the second case of the critical point of the form (0,y)(0,y), y>0y>0. We show that L4L_{4} uniquely exists and smoothly depends on (m,M,a)(+)3(m,M,a)\in({\mathbb{R}}^{+})^{3}. Using equation (20) set

f(y,m)=m4y3andg(y,M,a)=Ma3M((a/2)2+y2)3/2,f(y,m)=\frac{m}{4y^{3}}{\rm\ and\ }g(y,M,a)=\frac{M}{a^{3}}-\frac{M}{((a/2)^{2}+y^{2})^{3/2}},

and define

F(y,m,M,a)=g(y)f(y).F(y,m,M,a)=g(y)-f(y).

The function ff is positive with limit \infty as x0+x\to 0^{+} and limit 0 as xx\to\infty, and is decreasing. The function gg has limits

limy0+g(y)7Ma3<0andlimyg(y)=Ma3>0,\lim_{y\to 0^{+}}g(y)-\frac{7M}{a^{3}}<0{\rm\ and\ }\lim_{y\to\infty}g(y)=\frac{M}{a^{3}}>0,

and is increasing because

g(y)=3My((a/2)2+y2)5/2>0.g^{\prime}(y)=\frac{3My}{((a/2)^{2}+y^{2})^{5/2}}>0.

For each choice of (m,M,a)(+)3(m,M,a)\in({\mathbb{R}}^{+})^{3}, the graphs of ff and gg intersect in a unique point with y=L4>0y=L_{4}>0 which satisfies F(L4,m,M,a)=0F(L_{4},m,M,a)=0. Since

Fy(L4,m,M,a)=3m4L44+3ML4((a/2)2+L42)5/2>0\frac{\partial F}{\partial y}(L_{4},m,M,a)=\frac{3m}{4L_{4}^{4}}+\frac{3ML_{4}}{((a/2)^{2}+L_{4}^{2})^{5/2}}>0

it follows by the Implicit Function Theorem that L4L_{4} is smooth function of (m,M,a)(+)3(m,M,a)\in({\mathbb{R}}^{+})^{3}. Fixing M>0M>0 and a>0a>0 shows that L4L_{4} is smooth function of mm on the interval m>0m>0.

For M>0M>0 and a>0a>0 fixed, the graph of gg is fixed, while the graph of ff, which is independent of MM and aa, shifts the point of intersection between the graphs of ff and gg to the right as mm increases. Thus for fixed M>0M>0 and a>0a>0 the value of L4L_{4} increases with increasing mm. The limit of gg is positive as yy\to\infty while the limit of ff is 0 as mm\to\infty. This implies that L4L_{4}\to\infty as mm\to\infty.

For fixed M>0M>0 and a>0a>0, and with m>0m>0 and close to 0, we show that the unique solution y=L4y=L_{4} of (20) satisfies L4>3a/2L_{4}>\sqrt{3}a/2 and has limit 3a/2\sqrt{3}a/2 as m0+m\to 0^{+}. For y>0y>0 we use Equation (20) to define

F(y,m,M,a)=Ma3m4y3M((a/2)2+y2)3/2=0.F(y,m,M,a)=\frac{M}{a^{3}}-\frac{m}{4y^{3}}-\frac{M}{((a/2)^{2}+y^{2})^{3/2}}=0. (24)

The equation F(y,0,M,a)=0F(y,0,M,a)=0 has the solution

y=3a2y=\frac{\sqrt{3}a}{2}

which corresponds to the equilateral triangle equilibrium 4{\mathcal{L}}_{4} in the generalized Copenhagen problem. Because

Fy(3a2,0,M,a)=33M2a4>0,\frac{\partial F}{\partial y}\left(\frac{\sqrt{3}a}{2},0,M,a\right)=\frac{3\sqrt{3}M}{2a^{4}}>0,

the Implicit Function Theorem implies there exists a smooth function L4L_{4} defined on an open neighbourhood OO of {(0,M,a):M>0,a>0}3\{(0,M,a):M>0,a>0\}\subset{\mathbb{R}}^{3} such that L4(0,M,a)=3a/2L_{4}(0,M,a)=\sqrt{3}a/2 and F(L4(m,M,a),m,M,a)=0F(L_{4}(m,M,a),m,M,a)=0 for all (m,M,a)O(m,M,a)\in O. In particular, for fixed M>0M>0 and a>0a>0, there is ϵ>0\epsilon>0 (depending on MM and aa) and a smooth function L4(m)=L4(m,M,a)L_{4}(m)=L_{4}(m,M,a) such that L4(0)=L4(0,M,a)=3a/2L_{4}(0)=L_{4}(0,M,a)=\sqrt{3}a/2 and F(L4(m),m,M,a)=0F(L_{4}(m),m,M,a)=0 for all m(ϵ,ϵ)m\in(-\epsilon,\epsilon). Differentiation of F(L4(m),m,M,a)=0F(L_{4}(m),m,M,a)=0 with respect to mm followed by evaluation at m=0m=0 gives

dL4dm(0)=4a27M.\frac{dL_{4}}{dm}(0)=\frac{4a}{27M}.

This gives the Taylor series expansion of L4L_{4} in mm about m=0m=0 as

L4(m)=3a2+4a27Mm+O(m2).L_{4}(m)=\frac{\sqrt{3}a}{2}+\frac{4a}{27M}m+O(m^{2}). (25)

Thus, for m>0m>0 and close to 0 the unique critical point L4(m)>3a/2L_{4}(m)>\sqrt{3}a/2 satisfies L44=3a/2L_{4}\to{\mathcal{L}}_{4}=\sqrt{3}a/2 as m0+m\to 0^{+}. ∎

Remark 2.

Even though the six equilibria (±L1,0)(\pm L_{1},0), (±L2,0)(\pm L_{2},0), and (0,±L4)(0,\pm L_{4}) exist for all (m,M,a)(+)3(m,M,a)\in({\mathbb{R}}^{+})^{3}, they only are physically plausible for the model when mMm\ll M. In particular, Theorem 25 states for fixed M>0M>0 and a>0a>0 that the pair (0,±L4)(0,\pm L_{4}) has the property that L4L_{4}\to\infty as mm\to\infty. This implies in the case of the binary asteroids having the same mass as the primaries, m=Mm=M, the configuration of the four bodies is not a square. This stands in contrast to the known square configuration in the equal mass Four-Body Problem [16].

Remark 3.

The expansion of L4(m)L_{4}(m) in (25) is similar to, but differs from that in Theorem 2 part (c) in [6]. The difference is that here L4(m)L_{4}(m) increases from L4(0)L_{4}(0) as mm increases from 0, whereas in [6] the equivalent value of their L4(m)L_{4}(m) decreases from their L4(0)L_{4}(0) as mm increases from 0 (cf Case A2 in [16] when mm is small).

Although Theorem 2 gives six critical points, hence six equilibria of HxyH_{xy}, there are only 33 distinct configurations of the four bodies associated to the six equilibria. The primaries have the same mass MM and are located at the R1R_{1}-symmetric positions (±a/2,0)(\pm a/2,0). The COM coordinates imply that placing the asteroid with mass m1=mm_{1}=m at (x,y)(x,y) means placing the asteroid with mass m2=mm_{2}=m at (x,y)(-x,-y). The six equilibria (±L1,0)(\pm L_{1},0), (±L2,0)(\pm L_{2},0), and (0,±L4)(0,\pm L_{4}) are in the form of three symmetric pairs, with (±Li,0)(\pm L_{i},0), i=1,2i=1,2, being R1R_{1} symmetric, and (0,±L4)(0,\pm L_{4}) being R2R_{2} symmetric. So there are three distinct choices for the placement of the two asteroids in equilibrium.

Placing the two asteroids at (±L1,0)(\pm L_{1},0) gives the first of three distinct configurations of the four bodies. The masses M/2M/2, mm, mm, M/2M/2 are collinear and located at the R1R_{1}-symmetric positions (a/2,0)(-a/2,0), (L1,0)(-L_{1},0), (L1,0)(L_{1},0), (a/2,0)(a/2,0), where for fixed M>0M>0 and a>0a>0 the value of L1L_{1} depends smoothly on m>0m>0 by Theorem 2 (cf Section 2 in [19]). We call this collinear configuration non-separated because there is not a primary between the binary asteroids. It has the property that L11=0L_{1}\to{\mathcal{L}}_{1}=0 as m0m\to 0 by Theorem 2 (cf Case C2 in [16]). That the two asteroids in this collinear configuration are not separated means we could not use the Implicit Function Theorem to prove that L11L_{1}\to{\mathcal{L}}_{1} as m0+m\to 0^{+}.

Placing the two asteroids at (±L2,0)(\pm L_{2},0) gives the second of the distinct configuration of the four bodies. The masses mm, M/2M/2, M/2M/2, mm are collinear and located at the R1R_{1}-symmetric positions (L2,0)(-L_{2},0), (a/2,0)(-a/2,0), (a/2,0)(a/2,0), (L2,0)(L_{2},0), where for fixed M>0M>0 and a>0a>0 the value of L2L_{2} depends smoothly on m>0m>0 by Theorem 2 (cf Section 2 in [19]). We call this collinear configuration associated to (±L2,0)(\pm L_{2},0) separated because there is at least one primary (in this case both) between the binary asteroids. It has the property that L220L_{2}\to{\mathcal{L}}_{2}\neq 0 as m0m\to 0 by Theorem 2 (cf Case C1 in [16]). That the two asteroids are separated means we could use the Implicit Function Theorem to prove that L22L_{2}\to{\mathcal{L}}_{2} as m0+m\to 0^{+}.

Placing the two asteroids at (0,±L4)(0,\pm L_{4}) gives the third of three distinct configurations. The masses M/2M/2, mm, M/2M/2, mm are located at the R1R_{1}-symmetric and R2R_{2}-symmetric positions (a/2,0)(-a/2,0), (0,L4)(0,L_{4}), (a/2,0)(a/2,0), (0,L4)(0,-L_{4}), i.e., it is doubly-symmetric. This configuration is a convex kite (cf [6]), is a rhombus (cf [13]), and has an axes of symmetry (cf [1]).

3.3 Linearization Matrix

At a equilibrium, the linearization of the Hamiltonian system in (11) is

A=[02πP12m02πP0012m2Ux22Uxy02πP2Uxy2Uy22πP0].A=\begin{bmatrix}0&\displaystyle\frac{2\pi}{P}&\displaystyle\frac{1}{2m}&0\vspace{0.05in}\\ -\displaystyle\frac{2\pi}{P}&0&0&\displaystyle\frac{1}{2m}\vspace{0.05in}\\ \displaystyle\frac{\partial^{2}U}{\partial x^{2}}&\displaystyle\frac{\partial^{2}U}{\partial x\partial y}&0&\displaystyle\frac{2\pi}{P}\vspace{0.05in}\\ \displaystyle\frac{\partial^{2}U}{\partial x\partial y}&\displaystyle\frac{\partial^{2}U}{\partial y^{2}}&-\displaystyle\frac{2\pi}{P}&0\end{bmatrix}.

Using the relationship between the potential UU and the amended potential VV in (13), using Kepler’s Third Law in (3), and setting

ω2=Ma3,α=2Vx2,β=2Vxy,γ=2Vy2,\omega^{2}=\frac{M}{a^{3}},\ \alpha=\frac{\partial^{2}V}{\partial x^{2}},\ \beta=\frac{\partial^{2}V}{\partial x\partial y},\ \gamma=\frac{\partial^{2}V}{\partial y^{2}},

the linearization matrix becomes

A=[0ω12m0ω0012m2m(αω2)2mβ0ω2mβ2m(γω2)ω0].A=\begin{bmatrix}0&\omega&\displaystyle\frac{1}{2m}&0\vspace{0.05in}\\ -\omega&0&0&\displaystyle\frac{1}{2m}\vspace{0.05in}\\ 2m(\alpha-\omega^{2})&\displaystyle 2m\beta&0&\omega\vspace{0.05in}\\ 2m\beta&2m(\gamma-\omega^{2})&-\omega&0\end{bmatrix}.

The characteristic polynomial of AA is

det(zIA)=z4+(4ω2αγ)z2+αγβ2.{\rm det}(zI-A)=z^{4}+(4\omega^{2}-\alpha-\gamma)z^{2}+\alpha\gamma-\beta^{2}.

Setting w=z2w=z^{2}, the characteristic polynomial becomes

w2+(4ω2αγ)w+αγβ2.w^{2}+(4\omega^{2}-\alpha-\gamma)w+\alpha\gamma-\beta^{2}.

An equilibrium is an saddle-center if αγβ2<0\alpha\gamma-\beta^{2}<0, and an equilibrium is hyperbolic if the discriminant

(4ω2αγ)24(αγβ2)<0.(4\omega^{2}-\alpha-\gamma)^{2}-4(\alpha\gamma-\beta^{2})<0.

3.4 Spectral Stability

By the symmetry of VV, given in (15), we need only determine the eigenvalues of the linearizations for (L1,0)(L_{1},0), (L2,0)(L_{2},0), and (0,L4)(0,L_{4}).

Theorem 3.

Let (m,M,a)(+)3(m,M,a)\in({\mathbb{R}}^{+})^{3}. The equilibrium (L1,0)(L_{1},0) is a saddle-center, the equilibrium (L2,0)(L_{2},0) satifies L2(a/2,3a/2)L_{2}\in(a/2,3a/2) and is a saddle-center if mMm\ll M, and the equilibrium (0,L4)(0,L_{4}) is hyperbolic if mMm\ll M.

Proof.

To show that (L1,0)(L_{1},0) and (L2,0)(L_{2},0) are saddle-centers we show that αγβ2<0\alpha\gamma-\beta^{2}<0. Computing the sign of α\alpha and β\beta for L1L_{1} and L2L_{2} is relatively easy to do. By the convexity of xV(x,0)x\mapsto V(x,0), as given in (21), it follows that

α=2Vx2(Li,0)>0,i=1,2.\alpha=\frac{\partial^{2}V}{\partial x^{2}}(L_{i},0)>0,\ i=1,2.

Since

2Vxy=3mxy4(x2+y2)5/2+3M(x+a/2)y2((x+a/2)2+y2)5/2+3M(xa/2)y2((xa/2)2+y2)5/2,\frac{\partial^{2}V}{\partial x\partial y}=\frac{3mxy}{4(x^{2}+y^{2})^{5/2}}+\frac{3M(x+a/2)y}{2((x+a/2)^{2}+y^{2})^{5/2}}+\frac{3M(x-a/2)y}{2((x-a/2)^{2}+y^{2})^{5/2}}, (26)

it follows that

β=2Vxy(Li,0)=0,i=1,2.\beta=\frac{\partial^{2}V}{\partial x\partial y}(L_{i},0)=0,\ i=1,2.

To compute the sign of γ\gamma for LiL_{i}, i=1,2i=1,2, we start with the expression

γ=2Vy2(Li,0)=Ma3m4Li3M2d13M2d23,\gamma=\frac{\partial^{2}V}{\partial y^{2}}(L_{i},0)=\frac{M}{a^{3}}-\frac{m}{4L_{i}^{3}}-\frac{M}{2d_{1}^{3}}-\frac{M}{2d_{2}^{3}}, (27)

where

d1=|Li+a/2|andd2=|Lia/2|,i=1,2.d_{1}=|L_{i}+a/2|{\rm\ and\ }d_{2}=|L_{i}-a/2|,\ i=1,2.

When i=1i=1 the values d1d_{1} and d2d_{2} satisfy d1=L1+a/2d_{1}=L_{1}+a/2, d2=a/2L1d_{2}=a/2-L_{1}, and d1+d2=ad_{1}+d_{2}=a, and equation (22) that L1L_{1} satisfies becomes

0=ML1a3m4L12M2d12+M2d22.0=\frac{ML_{1}}{a^{3}}-\frac{m}{4L_{1}^{2}}-\frac{M}{2d_{1}^{2}}+\frac{M}{2d_{2}^{2}}. (28)

When i=1i=1 the values d1d_{1} and d2d_{2} satisfy d1=L2+a/2d_{1}=L_{2}+a/2, d2=L2a/2d_{2}=L_{2}-a/2, and d1d2=ad_{1}-d_{2}=a, and equation (23) that L2L_{2} satisfies becomes

0=ML2a3m4L22M2d12M2d22.0=\frac{ML_{2}}{a^{3}}-\frac{m}{4L_{2}^{2}}-\frac{M}{2d_{1}^{2}}-\frac{M}{2d_{2}^{2}}. (29)

Using (28) and the properties of d1d_{1} and d2d_{2} for L1L_{1} and then (29) and the properties of d1d_{1} and d2d_{2} for L2L_{2}, and some algebra in both cases, the expression for γ\gamma at LiL_{i} in (27) becomes

γ=m4Li2[d1LiLid1]M2d1[a3d23a2d23].\gamma=-\frac{m}{4L_{i}^{2}}\left[\frac{d_{1}-L_{i}}{L_{i}d_{1}}\right]-\frac{M}{2d_{1}}\left[\frac{a^{3}-d_{2}^{3}}{a^{2}d_{2}^{3}}\right].

Since d1Li=a/2>0d_{1}-L_{i}=a/2>0 for i=1,2i=1,2, there holds

m4Li2[d1LiLid1]<0.-\frac{m}{4L_{i}^{2}}\left[\frac{d_{1}-L_{i}}{L_{i}d_{1}}\right]<0.

To determine the sign of the remaining term in γ\gamma requires separate investigations for L1L_{1} and L2L_{2}.

To determine the value of γ\gamma for L1L_{1}, we note that d2=L1+a/2d_{2}=-L_{1}+a/2 and L1(0,a/2)L_{1}\in(0,a/2). So a>d2a>d_{2} which gives a3>d23a^{3}>d_{2}^{3} and hence

M2d1[a3d23a2d23]<0.-\frac{M}{2d_{1}}\left[\frac{a^{3}-d_{2}^{3}}{a^{2}d_{2}^{3}}\right]<0.

Thus γ<0\gamma<0 for L1L_{1} for all m>0m>0 and all M>0M>0.

To determine the value of γ\gamma for L2L_{2} we note that for

M2d1[a3d23a2d23]<0-\frac{M}{2d_{1}}\left[\frac{a^{3}-d_{2}^{3}}{a^{2}d_{2}^{3}}\right]<0

to hold requires that

a3d23>0a>d2.a^{3}-d_{2}^{3}>0\ \Leftrightarrow\ a>d_{2}.

Since d2=L2a/2d_{2}=L_{2}-a/2, the requirement a>d2a>d_{2} is equivalent to

L2<3a2.L_{2}<\frac{3a}{2}.

Since

Vx(3a2,0)=1a2[7M8m9],\frac{\partial V}{\partial x}\left(\frac{3a}{2},0\right)=\frac{1}{a^{2}}\left[\frac{7M}{8}-\frac{m}{9}\right],

and xV(x,0)x\mapsto V(x,0) is convex on (a/2,)(a/2,\infty), the requirement a>d2a>d_{2} is satisfied when 0<mM0<m\ll M, i.e., the unique critical point of VV on (a/2,)(a/2,\infty) occurs in interval (a/2,3a/2)(a/2,3a/2) because V/x(3a/2,0)>0\partial V/\partial x(3a/2,0)>0 when 0<mM0<m\ll M.

To show that (0,L4)(0,L_{4}) is hyperbolic we show that the discriminant (4ω2αγ)24(αγβ)<0(4\omega^{2}-\alpha-\gamma)^{2}-4(\alpha\gamma-\beta)<0. For L4L_{4} we compute explicit simplified expressions of α\alpha and γ\gamma and determine the value of β\beta. For the latter it follows easily from (26) that β=0\beta=0 because x=0x=0 for (0,L4)(0,L_{4}). Computing α\alpha we have

α=Ma3m4L43+3Ma24((a/2)2+L42)5/2M((a/2)2+L42)3/2.\alpha=\frac{M}{a^{3}}-\frac{m}{4L_{4}^{3}}+\frac{3Ma^{2}}{4((a/2)^{2}+L_{4}^{2})^{5/2}}-\frac{M}{((a/2)^{2}+L_{4}^{2})^{3/2}}. (30)

The point L4L_{4} satisfies (24), i.e.,

Ma3m4L43M((a/2)2+L42)3/2=0.\frac{M}{a^{3}}-\frac{m}{4L_{4}^{3}}-\frac{M}{((a/2)^{2}+L_{4}^{2})^{3/2}}=0. (31)

Making the obvious substitution of (31) into (30) gives

α=3Ma24((a/2)2+L42)5/2>0.\alpha=\frac{3Ma^{2}}{4((a/2)^{2}+L_{4}^{2})^{5/2}}>0. (32)

Evaluation of 2V/y2\partial^{2}V/\partial y^{2} at the point (0,L4)(0,L_{4}) gives

γ=Ma3+2m4L43+3ML42((a/2)2+L42)5/2M((a/2)2+L42)3/2.\gamma=\frac{M}{a^{3}}+\frac{2m}{4L_{4}^{3}}+\frac{3ML_{4}^{2}}{((a/2)^{2}+L_{4}^{2})^{5/2}}-\frac{M}{((a/2)^{2}+L_{4}^{2})^{3/2}}. (33)

Substitution of (31) into (33) gives

γ=3m4L43+3ML42((a/2)2+L42)5/2>0.\gamma=\frac{3m}{4L_{4}^{3}}+\frac{3ML_{4}^{2}}{((a/2)^{2}+L_{4}^{2})^{5/2}}>0. (34)

Substitution of the simplified expressions for α\alpha in (32) and γ\gamma in (34) into the discriminant, and substitution of the Taylor series expansion of L4L_{4} as a function of mm about m=0m=0 from (25), and then expanding the resulting expression as a Taylor series in mm about m=0m=0 gives

(4ω2αγ)24(αγβ2)=Ma6[23M4+53m3]+O(m2).(4\omega^{2}-\alpha-\gamma)^{2}-4(\alpha\gamma-\beta^{2})=\frac{M}{a^{6}}\left[-\frac{23M}{4}+\frac{5\sqrt{3}m}{3}\right]+O(m^{2}).

This shows that the discriminant is negative when 0<mM0<m\ll M. ∎

The hyperbolic dynamics near (0,L4)(0,L_{4}) for mMm\ll M are related to corresponding hyperbolic dynamics near (0,L4)(0,-L_{4}) for mMm\ll M by time reversing symmetry R^2\hat{R}_{2} in (16).

Corollary 4.

Let (m,M,a)(+)3(m,M,a)\in({\mathbb{R}}^{+})^{3}. Emanating from each equilibrium (±L1,0)(\pm L_{1},0) is a one-parameter family of periodic orbits. Emanating from each equilibrium (±L2,0)(\pm L_{2},0) when mMm\ll M is a one-parameter family of periodic orbits.

Proof.

This follows by the Lyapunov Center Theorem (see [14]) for (Li,0)(L_{i},0), i=1,2i=1,2. Applying the time-reversing symmetry R^1\hat{R}_{1} in (16) gives the result for (Li,0)(-L_{i},0), i=1,2i=1,2. ∎

4 Periodic Orbits Near the Origin

The two asteroids in COM coordinates experience a binary collision at the origin. We show near the origin that there are two families of periodic orbits in the COM reduced Binary Asteroid Problem through symplectic scaling and the Poincaré Continuation Method as implemented in Meyer and Offin [14].

4.1 Scaling

For small δ>0\delta>0 the transformation (x,y,px,py)(ξ,η,pξ,pη)(x,y,p_{x},p_{y})\mapsto(\xi,\eta,p_{\xi},p_{\eta}) given by the inverse

x=δ2ξ,\displaystyle x=\delta^{2}\xi, y=δ2η,\displaystyle y=\delta^{2}\eta, px=δ1pξ,\displaystyle p_{x}=\delta^{-1}p_{\xi}, py=δ1pη\displaystyle p_{y}=\delta^{-1}p_{\eta} (35)

is symplectic with multiplier δ1\delta^{-1}. The Hamiltonian HxyH_{xy} in (10) becomes

Hξη\displaystyle H_{\xi\eta} =δ1Hxy\displaystyle=\delta^{-1}H_{xy} (36)
=14δ3m(pη2+pξ2)2πP(ξpηηpξ)m22δ3ξ2+η2\displaystyle=\frac{1}{4\delta^{3}m}\big{(}p_{\eta}^{2}+p_{\xi}^{2}\big{)}-\frac{2\pi}{P}\big{(}\xi p_{\eta}-\eta p_{\xi}\big{)}-\frac{m^{2}}{2\delta^{3}\sqrt{\xi^{2}+\eta^{2}}}
mMδ(δ2ξ+a/2)2+(δ2η)2mMδ(δ2ξa/2)2+(δ2η)2.\displaystyle\ \ \ \ -\frac{mM}{\delta\sqrt{(\delta^{2}\xi+a/2)^{2}+(\delta^{2}\eta)^{2}}}-\frac{mM}{\delta\sqrt{(\delta^{2}\xi-a/2)^{2}+(\delta^{2}\eta)^{2}}}.

Scaling time by tδ3tt\to\delta^{-3}t is the same as multiplying HξηH_{\xi\eta} by δ3\delta^{3}. Multiplying HξηH_{\xi\eta} in (36) by δ3\delta^{3} gives the Hamiltonian

Hξη\displaystyle H_{\xi\eta} =14m(pη2+pξ2)2πδ3P(ξpηηpξ)m22ξ2+η2\displaystyle=\frac{1}{4m}\big{(}p_{\eta}^{2}+p_{\xi}^{2}\big{)}-\frac{2\pi\delta^{3}}{P}\big{(}\xi p_{\eta}-\eta p_{\xi}\big{)}-\frac{m^{2}}{2\sqrt{\xi^{2}+\eta^{2}}} (37)
δ2mM(δ2ξ+a/2)2+(δ2η)2δ2mM(δ2ξa/2)2+(δ2η)2.\displaystyle\ \ \ \ -\frac{\delta^{2}mM}{\sqrt{(\delta^{2}\xi+a/2)^{2}+(\delta^{2}\eta)^{2}}}-\frac{\delta^{2}mM}{\sqrt{(\delta^{2}\xi-a/2)^{2}+(\delta^{2}\eta)^{2}}}.

The last two terms in (37) are real analytic in δ\delta at δ=0\delta=0. Their Taylor series expansions in δ\delta about δ=0\delta=0 are

δ2mM(δ2ξ+a/2)2+(δ2η)2\displaystyle-\frac{\delta^{2}mM}{\sqrt{(\delta^{2}\xi+a/2)^{2}+(\delta^{2}\eta)^{2}}} =2mMaδ3+4mMξa2δ5+O(δ7),\displaystyle=-\frac{2mM}{a}\delta^{3}+\frac{4mM\xi}{a^{2}}\delta^{5}+O(\delta^{7}),
δ2mM(δ2ξa/2)2+(δ2η)2\displaystyle-\frac{\delta^{2}mM}{\sqrt{(\delta^{2}\xi-a/2)^{2}+(\delta^{2}\eta)^{2}}} =2mMaδ34mMξa2δ5+O(δ7).\displaystyle=-\frac{2mM}{a}\delta^{3}-\frac{4mM\xi}{a^{2}}\delta^{5}+O(\delta^{7}).

In these two expansions the terms of order δ5\delta^{5} cancel and the terms of order δ3\delta^{3} are constants (which we ignore), so the expansion of the Hamiltonian is

Hξη=H0+δ3H3+O(δ7)H_{\xi\eta}=H_{0}+\delta^{3}H_{3}+O(\delta^{7})

where

H0\displaystyle H_{0} =14m(pξ2+pη2)m22ξ2+η2,\displaystyle=\frac{1}{4m}\big{(}p_{\xi}^{2}+p_{\eta}^{2}\big{)}-\frac{m^{2}}{2\sqrt{\xi^{2}+\eta^{2}}},
H3\displaystyle H_{3} =2πP(ξpηηpξ).\displaystyle=-\frac{2\pi}{P}\big{(}\xi p_{\eta}-\eta p_{\xi}\big{)}.

4.2 Periodic Solutions to O(δ3)O(\delta^{3})

Seeking circular periodic solutions, we transform H=H0+δ3H3+O(δ7)H=H_{0}+\delta^{3}H_{3}+O(\delta^{7}) into symplectic polar coordinates

(ξ,η,pξ,pη)(r,θ,pr,pθ)(\xi,\eta,p_{\xi},p_{\eta})\to(r,\theta,p_{r},p_{\theta})

defined by (the inverse transformation, see [14])

ξ\displaystyle\xi =rcosθ,η=rsinθ,\displaystyle=r\cos\theta,\ \eta=r\sin\theta,
pξ\displaystyle p_{\xi} =prcosθpθrsinθ,pη=prsinθ+pθrcosθ.\displaystyle=p_{r}\cos\theta-\frac{p_{\theta}}{r}\sin\theta,\ p_{\eta}=p_{r}\sin\theta+\frac{p_{\theta}}{r}\cos\theta.

It follows that

H0=14m[pr2+pθ2r2]m22r,H3=2πpθP.H_{0}=\frac{1}{4m}\left[p_{r}^{2}+\frac{p_{\theta}^{2}}{r^{2}}\right]-\frac{m^{2}}{2r},\ H_{3}=-\frac{2\pi p_{\theta}}{P}.

The corresponding system of ODEs for H0+δ3H3H_{0}+\delta^{3}H_{3} is

r˙=pr2m,\displaystyle\dot{r}=\frac{p_{r}}{2m}, θ˙=pθ2mr2δ32πP,\displaystyle\dot{\theta}=\frac{p_{\theta}}{2mr^{2}}-\delta^{3}\frac{2\pi}{P}, p˙r=pθ22mr3m22r2,\displaystyle\dot{p}_{r}=\frac{p_{\theta}^{2}}{2mr^{3}}-\frac{m^{2}}{2r^{2}}, p˙θ=0.\displaystyle\dot{p}_{\theta}=0. (38)

For c=m3c=\sqrt{m^{3}}, equations (38) have two circular periodic solutions

r(t)=1,θ(t)=θ(0)+ωt,pr(t)=0,pθ=±c,r(t)=1,\ \theta(t)=\theta(0)+\omega t,\ p_{r}(t)=0,\ p_{\theta}=\pm c, (39)

whose frequencies are

ω=±m2δ32πP\omega=\pm\frac{\sqrt{m}}{2}-\delta^{3}\frac{2\pi}{P}

and whose corresponding periods are

T=2π|ω|=4πPmP4πδ3.T=\frac{2\pi}{|\omega|}=\frac{4\pi P}{\sqrt{m}P\mp 4\pi\delta^{3}}.

4.3 Multipliers and Continuation

We linearize the ODEs for r˙\dot{r} and p˙r\dot{p}_{r} along the two circular periodic solutions (39) to obtain their multipliers. The linearization is

r˙=pr2m,p˙r=m22r.\dot{r}=\frac{p_{r}}{2m},\ \dot{p}_{r}=-\frac{m^{2}}{2}r.

The nontrivial multipliers of the two circular periodic solutions are

exp(±iTm2)=1±2πi(4πmP)δ3+O(δ6).\exp\left(\pm iT\frac{\sqrt{m}}{2}\right)=1\pm 2\pi i\left(\frac{4\pi}{\sqrt{m}P}\right)\delta^{3}+O(\delta^{6}).

Thus, for small positive δ\delta, the two circular periodic solutions are elementary. By the argument in Meyer and Offin [14] and Theorem 1 we have the following result.

Theorem 5.

There exist two one-parameter families of nearly circular periodic solutions to the Binary Asteroid Problem (6) in the case of equal asteroid (m1=m2=m)(m_{1}=m_{2}=m) and equal primary (M1=M2=M/2)(M_{1}=M_{2}=M/2) masses, where the two asteroids are near the origin.

Figure 3 shows a solution to (39) undergoing a numerical process of continuation into (10). Notice how the orbits maintain their near-circular shape and periodic nature. This is consistent with Theorem 5.

Refer to caption
Figure 3: Depiction of a numerical δ\delta continuation process for a periodic orbit near the origin, according to (39). The hues darken as δ\delta is increased, with the darkest orbit corresponding to δ=1\delta=1. (Left) A view of positions showing the primaries as black x’s. (Middle) The left panel zoomed in to see the detail.

5 Hill-Type Orbits

We show that there exist two one-parameter families of near-circular Hill-type orbits (cf. [12, 14]) in the COM reduced Binary Asteroid Problem where one asteroid orbits one primary and the other asteroid orbits the other primary. We use symplectic scaling to make the Hamiltonian HxyH_{xy} resemble that of a rotating 2-body problem, which has simple circular periodic solutions.

5.1 Scaling

Using the COM reduced Hamiltonian (10), we perform a symplectic shift which places the left primary, located at (x,y)=(a/2,0)(x,y)=(-a/2,0), at the origin. We must also shift the vertical momentum in order to mitigate the introduction of unfavorable terms. The shift is

xx+a2,\displaystyle x\mapsto x+\frac{a}{2}, yy,\displaystyle y\mapsto y, pxpx,\displaystyle p_{x}\mapsto p_{x}, pypy+2πamP.\displaystyle p_{y}\mapsto p_{y}+\frac{2\pi am}{P}. (40)

In the shifted coordinates, the Hamiltonian (10) becomes

Hxy\displaystyle H_{xy} =14m(px2+py2)2πP(xpyypx)+4π2amxP2π2a2mP2\displaystyle=\frac{1}{4m}\left(p_{x}^{2}+p_{y}^{2}\right)-\frac{2\pi}{P}\left(xp_{y}-yp_{x}\right)+\frac{4\pi^{2}amx}{P^{2}}-\frac{\pi^{2}a^{2}m}{P^{2}}
m22(xa/2)2+y2mMx2+y2mM(xa)2+y2.\displaystyle\ \ \ \ -\frac{m^{2}}{2\sqrt{(x-a/2)^{2}+y^{2}}}-\frac{mM}{\sqrt{x^{2}+y^{2}}}-\frac{mM}{\sqrt{(x-a)^{2}+y^{2}}}. (41)

We use Kepler’s Third Law (3) to simplify 4π2amx/P2=mMx/a24\pi^{2}amx/P^{2}=mMx/a^{2}. Additionally, since (41) is time-independent, HxyH_{xy} remains an integral, and so the constant π2a2m/P2-\pi^{2}a^{2}m/P^{2} term is dropped.

For small δ>0\delta>0 the transformation (x,y,px,py)(ξ,η,pξ,pη)(x,y,p_{x},p_{y})\mapsto(\xi,\eta,p_{\xi},p_{\eta}) given by

x=δ2ξ,\displaystyle x=\delta^{2}\xi, y=δ2η,\displaystyle y=\delta^{2}\eta, px=δ1pξ,\displaystyle p_{x}=\delta^{-1}p_{\xi}, py=δ1pη\displaystyle p_{y}=\delta^{-1}p_{\eta} (42)

is symplectic with multiplier δ1\delta^{-1}. The Hamiltonian (41) becomes

Hξη=δ1Hxy\displaystyle H_{\xi\eta}=\delta^{-1}H_{xy} =δ34m(pξ2+pη2)2πP(ξpηηpξ)+δmMξa2δ3mMξ2+η2\displaystyle=\frac{\delta^{-3}}{4m}\left(p_{\xi}^{2}+p_{\eta}^{2}\right)-\frac{2\pi}{P}\left(\xi p_{\eta}-\eta p_{\xi}\right)+\frac{\delta mM\xi}{a^{2}}-\frac{\delta^{-3}mM}{\sqrt{\xi^{2}+\eta^{2}}}
δ1m22(δ2ξa/2)2+(δ2η)2δ1mM(δ2ξa)2+(δ2η)2.\displaystyle\ \ \ \ -\frac{\delta^{-1}m^{2}}{2\sqrt{(\delta^{2}\xi-a/2)^{2}+(\delta^{2}\eta)^{2}}}-\frac{\delta^{-1}mM}{\sqrt{(\delta^{2}\xi-a)^{2}+(\delta^{2}\eta)^{2}}}. (43)

Scaling time via tδ3tt\mapsto\delta^{-3}t is the same as multiplying the Hamiltonian HξηH_{\xi\eta} in (43) by δ3\delta^{3}. After multiplying HξηH_{\xi\eta} by δ3\delta^{3} and grouping terms of like powers of δ\delta, we find

Hξη\displaystyle H_{\xi\eta} =[14m(pξ2+pη2)mMξ2+η2]δ3[2πP(ξpηηpξ)]+δ4[mMξa2]\displaystyle=\left[\frac{1}{4m}\left(p_{\xi}^{2}+p_{\eta}^{2}\right)-\frac{mM}{\sqrt{\xi^{2}+\eta^{2}}}\right]-\delta^{3}\left[\frac{2\pi}{P}\left(\xi p_{\eta}-\eta p_{\xi}\right)\right]+\delta^{4}\left[\frac{mM\xi}{a^{2}}\right]
δ2[m22(δ2ξa/2)2+(δ2η)2+mM(δ2ξa)2+(δ2η)2].\displaystyle\ \ \ \ -\delta^{2}\left[\frac{m^{2}}{2\sqrt{(\delta^{2}\xi-a/2)^{2}+(\delta^{2}\eta)^{2}}}+\frac{mM}{\sqrt{(\delta^{2}\xi-a)^{2}+(\delta^{2}\eta)^{2}}}\right]. (44)

The Hamiltonian in (44) is analytic in δ\delta about 0. Hence we expand HξηH_{\xi\eta} in a Taylor series in δ\delta about 0, yielding

Hξη=H0δ2m(m+M)a+δ3H3+δ4H4+O(δ6),H_{\xi\eta}=H_{0}-\frac{\delta^{2}m(m+M)}{a}+\delta^{3}H_{3}+\delta^{4}H_{4}+O(\delta^{6}), (45)

where

H0=14m(pξ2+pη2)mMξ2+η2,\displaystyle H_{0}=\frac{1}{4m}\left(p_{\xi}^{2}+p_{\eta}^{2}\right)-\frac{mM}{\sqrt{\xi^{2}+\eta^{2}}}, H3=2πP(ξpηηpξ),\displaystyle H_{3}=-\frac{2\pi}{P}\left(\xi p_{\eta}-\eta p_{\xi}\right), H4=2m2ξa2.\displaystyle H_{4}=-\frac{2m^{2}\xi}{a^{2}}. (46)

Once again, we ignore the constant δ2m(m+M)/a-\delta^{2}m(m+M)/a by time-independence.

5.2 Periodic Solutions to O(δ3)O(\delta^{3})

The Hamiltonian system defined by H0+δ3H3H_{0}+\delta^{3}H_{3} in (45) has circular solutions. To obtain them, we first make the change (ξ,η,pξ,pη)(r,θ,pr,pθ)(\xi,\eta,p_{\xi},p_{\eta})\mapsto(r,\theta,p_{r},p_{\theta}) to symplectic polar coordinates:

ξ\displaystyle\xi =rcosθ,η=rsinθ,\displaystyle=r\cos\theta,\ \eta=r\sin\theta,
pξ\displaystyle p_{\xi} =prcosθpθrsinθ,pη=prsinθ+pθrcosθ.\displaystyle=p_{r}\cos\theta-\frac{p_{\theta}}{r}\sin\theta,\ p_{\eta}=p_{r}\sin\theta+\frac{p_{\theta}}{r}\cos\theta.

The Hamiltonian becomes

H0+δ3H3=14m(pr2+pθ2r2)mMrδ32πpθP.H_{0}+\delta^{3}H_{3}=\frac{1}{4m}\left(p_{r}^{2}+\frac{p_{\theta}^{2}}{r^{2}}\right)-\frac{mM}{r}-\delta^{3}\frac{2\pi p_{\theta}}{P}. (47)

The equations of motion defined by (47) are

r˙=pr2m,\displaystyle\dot{r}=\frac{p_{r}}{2m}, θ˙=pθ2mr22πδ3P,\displaystyle\dot{\theta}=\frac{p_{\theta}}{2mr^{2}}-\frac{2\pi\delta^{3}}{P}, p˙r=pθ22mr3mMr2,\displaystyle\dot{p}_{r}=\frac{p_{\theta}^{2}}{2mr^{3}}-\frac{mM}{r^{2}}, p˙θ=0.\displaystyle\dot{p}_{\theta}=0. (48)

For c=2m2Mc=\sqrt{2m^{2}M}, equations (48) have two circular periodic solutions

r=1,θ(t)=θ(0)+ωt,pr=0,pθ=±c,r=1,\ \theta(t)=\theta(0)+\omega t,\ p_{r}=0,\ p_{\theta}=\pm c, (49)

whose frequencies are

ω=±cP4πmδ32mP=±m2MP4πmδ32mP\omega=\frac{\pm cP-4\pi m\delta^{3}}{2mP}=\frac{\pm m\sqrt{2M}P-4\pi m\delta^{3}}{2mP}

and corresponding periods are

T=2π|ω|=4πmPcP4πmδ3=4πmPm2MP4πmδ3.T=\frac{2\pi}{|\omega|}=\frac{4\pi mP}{cP\mp 4\pi m\delta^{3}}=\frac{4\pi mP}{m\sqrt{2M}P\mp 4\pi m\delta^{3}}. (50)

5.3 Multipliers and Continuation

We linearize the ODEs in rr and prp_{r} in (48) about the two circular periodic solutions (49) to obtain their multipliers. The linearization gives

r˙=pr2m,p˙r=mMr,\dot{r}=\frac{p_{r}}{2m},\ \dot{p}_{r}=-mMr,

whose solutions have the form exp(±itM/2)\exp(\pm it\sqrt{M/2}). The nontrivial multipliers of the two circular periodic solutions are

exp(±iTM/2)\displaystyle\exp(\pm iT\sqrt{M/2}) =exp(±iM24πmPm2MP4πmδ3)\displaystyle=\exp\left(\pm i\sqrt{\frac{M}{2}}\frac{4\pi mP}{m\sqrt{2M}P\mp 4\pi m\delta^{3}}\right)
=1±2πi(2π2PM)δ3+O(δ6).\displaystyle=1\pm 2\pi i\left(\frac{2\pi\sqrt{2}}{P\sqrt{M}}\right)\delta^{3}+O(\delta^{6}).

Thus, for small positive δ\delta, the two circular periodic solutions are elementary. By the Poincaré continuation argument in Meyer and Offin [14], and Theorem 1 we have the following result.

Theorem 6 (Hill Orbits).

There exist two one-parameter families of near-circular Hill-type periodic solutions to the Binary Asteroid Problem (6) in the case of equal asteroid (m1=m2=m)(m_{1}=m_{2}=m) and equal primary (M1=M2=M/2)(M_{1}=M_{2}=M/2) masses, where each asteroid orbits near one primary.

Figure 4 shows a solution to (49) undergoing a numerical process of continuation into (10), similar to the previous section. Notice the geometric similarities with the orbits depicted in Figure 3.

Refer to caption
Figure 4: Depiction of a numerical δ\delta continuation process of a Hill-type orbit from the circular orbit in (49).

5.4 Numerical Continuation

We remark that we have some numerical evidence that the orbits of Theorem 6 persist as quasi-periodic orbits, as we perturb the primary mass deviation parameter ν\nu from 0 (see Figure 5). This suggests that the COM approach can be adapted to perform the reduction in the arbitrary primary mass case. However, we have not yet determined what that adaptation is.

Refer to caption
Figure 5: Depiction of a numerical ν\nu continuation process for the darkest Hill-type orbit in Figure 4. The x’s in the left graph mark the primaries and are shaded to match their associated orbits.

6 Comet-Type Orbits

We show that there are two one-parameter families of near-circular comet-orbits in the Binary Asteroid Problem (cf. [14, 12]) when the two ratios m1/M1m_{1}/M_{1} and m2/M2m_{2}/M_{2} are small enough. We could have applied a symplectic scaling to the Hamiltonian HxyH_{xy} of the COM reduced Binary Asteroid Problem, but we found a more general approach for the Hamiltonian HBAH_{BA} of the Binary Asteroid Problem that does not requires the asteroids nor the primaries to have equal mass.

6.1 Scaling

Following the comet approach in [14], let δ>0\delta>0 be a small scaling parameter and define the transformation (xi,yi,pxi,pyi)(ξi,ηi,pξi,pηi)(x_{i},y_{i},p_{x_{i}},p_{y_{i}})\mapsto(\xi_{i},\eta_{i},p_{\xi_{i}},p_{\eta_{i}}), i=1,2i=1,2, by

xi=δ2ξi,\displaystyle x_{i}=\delta^{-2}\xi_{i}, yi=δ2ηi,\displaystyle y_{i}=\delta^{-2}\eta_{i}, pxi=δpξi,\displaystyle p_{x_{i}}=\delta p_{\xi_{i}}, pyi=δpηi.\displaystyle p_{y_{i}}=\delta p_{\eta_{i}}. (51)

This is symplectic with multiplier δ\delta and the Binary Asteroid Hamiltonian HBAH_{BA} in (6) becomes

Hξη\displaystyle H_{\xi\eta} =δHBA\displaystyle=\delta H_{BA} (52)
=δ32m1(pξ12+pη12)+δ32m2(pξ22+pη22)2πP(ξ1pη1η1pξ1)\displaystyle=\frac{\delta^{3}}{2m_{1}}\left(p_{\xi_{1}}^{2}+p_{\eta_{1}}^{2}\right)+\frac{\delta^{3}}{2m_{2}}\left(p_{\xi_{2}}^{2}+p_{\eta_{2}}^{2}\right)-\frac{2\pi}{P}\left(\xi_{1}p_{\eta_{1}}-\eta_{1}p_{\xi_{1}}\right)
2πP(ξ2pη2η2pξ2)δ3m1m2(ξ1ξ2)2+(η1η2)2\displaystyle\ \ \ \ -\frac{2\pi}{P}\left(\xi_{2}p_{\eta_{2}}-\eta_{2}p_{\xi_{2}}\right)-\frac{\delta^{3}m_{1}m_{2}}{\sqrt{(\xi_{1}-\xi_{2})^{2}+(\eta_{1}-\eta_{2})^{2}}}
δ3m1M1(ξ1+aδ2M2/M)2+η12δ3m2M1(ξ2+aδ2M2/M)2+η22\displaystyle\ \ \ \ -\frac{\delta^{3}m_{1}M_{1}}{\sqrt{(\xi_{1}+a\delta^{2}M_{2}/M)^{2}+\eta_{1}^{2}}}-\frac{\delta^{3}m_{2}M_{1}}{\sqrt{(\xi_{2}+a\delta^{2}M_{2}/M)^{2}+\eta_{2}^{2}}}
δ3m1M2(ξ1aδ2M1/M)2+η12δ3m2M2(ξ2aδ2M1/M)2+η22.\displaystyle\ \ \ \ -\frac{\delta^{3}m_{1}M_{2}}{\sqrt{(\xi_{1}-a\delta^{2}M_{1}/M)^{2}+\eta_{1}^{2}}}-\frac{\delta^{3}m_{2}M_{2}}{\sqrt{(\xi_{2}-a\delta^{2}M_{1}/M)^{2}+\eta_{2}^{2}}}. (53)

The Hamiltonian HξηH_{\xi\eta} in (53) is analytic in δ\delta at zero, hence we expand the last four terms in HξηH_{\xi\eta} in a Taylor series in δ\delta about δ=0\delta=0 and then group terms with like powers of δ\delta. The terms from the expansions with δ5\delta^{5} cancel and the terms with δ3\delta^{3} combine to give

Hξη=H0+δ3H3+O(δ7),H_{\xi\eta}=H_{0}+\delta^{3}H_{3}+O(\delta^{7}), (54)

where

H0\displaystyle H_{0} =2πP(ξ1pη1η1pξ1+ξ2pη2η2pξ2),\displaystyle=-\frac{2\pi}{P}\left(\xi_{1}p_{\eta_{1}}-\eta_{1}p_{\xi_{1}}+\xi_{2}p_{\eta_{2}}-\eta_{2}p_{\xi_{2}}\right),
H3\displaystyle H_{3} =12m1(pξ12+pη12)+12m2(pξ22+pη22)m1Mξ12+η12m2Mξ22+η22\displaystyle=\frac{1}{2m_{1}}\left(p_{\xi_{1}}^{2}+p_{\eta_{1}}^{2}\right)+\frac{1}{2m_{2}}\left(p_{\xi_{2}}^{2}+p_{\eta_{2}}^{2}\right)-\frac{m_{1}M}{\sqrt{\xi_{1}^{2}+\eta_{1}^{2}}}-\frac{m_{2}M}{\sqrt{\xi_{2}^{2}+\eta_{2}^{2}}}
m1m2(ξ1ξ2)2+(η1η2)2.\displaystyle\ \ \ \ -\frac{m_{1}m_{2}}{\sqrt{(\xi_{1}-\xi_{2})^{2}+(\eta_{1}-\eta_{2})^{2}}}. (55)

6.2 Phase Condition in O(δ3)O(\delta^{3})

To find circular periodic solutions of the system with Hamiltonian H0+δ3H3H_{0}+\delta^{3}H_{3} we change to double symplectic polar coordinates (ξi,ηi,pξi,pηi)(ri,θi,pri,pθi)(\xi_{i},\eta_{i},p_{\xi_{i}},p_{\eta_{i}})\mapsto(r_{i},\theta_{i},p_{r_{i}},p_{\theta_{i}}). This transformation is symplectic since the new coordinates of a single asteroid (i.e., fixed ii) depend only on that asteroid’s old coordinates, via a symplectic map (see [14]). The Hamiltonian becomes

H0+δ3H3\displaystyle H_{0}+\delta^{3}H_{3} =2πP(pθ1+pθ2)+δ32m1(pr12+pθ12r12)+δ32m2(pr22+pθ22r22)\displaystyle=-\frac{2\pi}{P}\left(p_{\theta_{1}}+p_{\theta_{2}}\right)+\frac{\delta^{3}}{2m_{1}}\left(p_{r_{1}}^{2}+\frac{p_{\theta_{1}}^{2}}{r_{1}^{2}}\right)+\frac{\delta^{3}}{2m_{2}}\left(p_{r_{2}}^{2}+\frac{p_{\theta_{2}}^{2}}{r_{2}^{2}}\right)
δ3m1Mr1δ3m2Mr2δ3m1m2r12+r222r1r2cos(θ1θ2).\displaystyle\ \ \ \ -\frac{\delta^{3}m_{1}M}{r_{1}}-\frac{\delta^{3}m_{2}M}{r_{2}}-\frac{\delta^{3}m_{1}m_{2}}{\sqrt{r_{1}^{2}+r_{2}^{2}-2r_{1}r_{2}\cos(\theta_{1}-\theta_{2})}}. (56)

The associated equations of motion are

r˙1\displaystyle\dot{r}_{1} =δ3pr1m1,θ˙1=2πP+δ3pθ1m1r12,\displaystyle=\frac{\delta^{3}p_{r_{1}}}{m_{1}},\ \dot{\theta}_{1}=-\frac{2\pi}{P}+\frac{\delta^{3}p_{\theta_{1}}}{m_{1}r_{1}^{2}}, (57)
r˙2\displaystyle\dot{r}_{2} =δ3pr2m2,θ2˙=2πP+δ3pθ2m2r22,\displaystyle=\frac{\delta^{3}p_{r_{2}}}{m_{2}},\ \dot{\theta_{2}}=-\frac{2\pi}{P}+\frac{\delta^{3}p_{\theta_{2}}}{m_{2}r_{2}^{2}},
p˙r1\displaystyle\dot{p}_{r_{1}} =δ3[pθ12m1r13m1Mr12m1m2(r1r2cos(θ1θ2))r12+r222r1r2cos(θ1θ2)3],\displaystyle=\delta^{3}\left[\frac{p_{\theta_{1}}^{2}}{m_{1}r_{1}^{3}}-\frac{m_{1}M}{r_{1}^{2}}-\frac{m_{1}m_{2}(r_{1}-r_{2}\cos(\theta_{1}-\theta_{2}))}{\sqrt{r_{1}^{2}+r_{2}^{2}-2r_{1}r_{2}\cos(\theta_{1}-\theta_{2})}^{3}}\right],
p˙θ1\displaystyle\dot{p}_{\theta_{1}} =δ3m1m2r1r2sin(θ1θ2)r12+r222r1r2cos(θ1θ2)3,\displaystyle=-\frac{\delta^{3}m_{1}m_{2}r_{1}r_{2}\sin(\theta_{1}-\theta_{2})}{\sqrt{r_{1}^{2}+r_{2}^{2}-2r_{1}r_{2}\cos(\theta_{1}-\theta_{2})}^{3}},
p˙r2\displaystyle\dot{p}_{r_{2}} =δ3[pθ22m2r23m2Mr22m1m2(r2r1cos(θ1θ2))r12+r222r1r2cos(θ1θ2)3],\displaystyle=\delta^{3}\left[\frac{p_{\theta_{2}}^{2}}{m_{2}r_{2}^{3}}-\frac{m_{2}M}{r_{2}^{2}}-\frac{m_{1}m_{2}(r_{2}-r_{1}\cos(\theta_{1}-\theta_{2}))}{\sqrt{r_{1}^{2}+r_{2}^{2}-2r_{1}r_{2}\cos(\theta_{1}-\theta_{2})}^{3}}\right],
p˙θ2\displaystyle\dot{p}_{\theta_{2}} =δ3m1m2r1r2sin(θ1θ2)r12+r222r1r2cos(θ1θ2)3.\displaystyle=\frac{\delta^{3}m_{1}m_{2}r_{1}r_{2}\sin(\theta_{1}-\theta_{2})}{\sqrt{r_{1}^{2}+r_{2}^{2}-2r_{1}r_{2}\cos(\theta_{1}-\theta_{2})}^{3}}.

Any solution of system (57) which has sin(θ1θ2)0\sin(\theta_{1}-\theta_{2})\equiv 0 will necessarily have pθi=constp_{\theta_{i}}=const, i=1,2i=1,2. Requiring this restriction implies θ1θ2=nπ\theta_{1}-\theta_{2}=n\pi for some nn\in\mathbb{Z} and for all time. When nn is even, the asteroids have equal phase relative to the barycenter. We determined numerically that this behavior leads to a collision between the asteroids. However, when nn is odd, the asteroids have opposite phase, and there exist non-collision solutions.

In the odd case it is sufficient to study n=1n=1; we impose θ1θ2=π\theta_{1}-\theta_{2}=\pi. This implies through the pθip_{\theta_{i}} ODEs that pθi=cip_{\theta_{i}}=c_{i} for some constants cic_{i}\in\mathbb{R}. The phase condition also implies that θ˙1=θ2˙\dot{\theta}_{1}=\dot{\theta_{2}} for all time. That is,

2πP+δ3pθ1m1r12=2πP+δ3pθ2m2r22(r1r2)2=c1/m1c2/m2=(c1c2)(m2m1)>0.\displaystyle-\frac{2\pi}{P}+\frac{\delta^{3}p_{\theta_{1}}}{m_{1}r_{1}^{2}}=-\frac{2\pi}{P}+\frac{\delta^{3}p_{\theta_{2}}}{m_{2}r_{2}^{2}}\implies\left(\frac{r_{1}}{r_{2}}\right)^{2}=\frac{c_{1}/m_{1}}{c_{2}/m_{2}}=\left(\frac{c_{1}}{c_{2}}\right)\left(\frac{m_{2}}{m_{1}}\right)>0.

Immediately this implies r1/r2=const.r_{1}/r_{2}=const., and c1c_{1} and c2c_{2} must have the same sign. Define

b=c1/m1c2/m2>0,b=\sqrt{\frac{c_{1}/m_{1}}{c_{2}/m_{2}}}>0, (58)

so that we have r1(t)=br2(t)r_{1}(t)=br_{2}(t) for all time. We can therefore solve the ODEs (57) in terms of the second asteroid’s coordinates (r2,θ2,pr2,pθ2)(r_{2},\theta_{2},p_{r_{2}},p_{\theta_{2}}).

We investigate circular periodic solutions in the regime wherein ri=constr_{i}=const. In this case, the rir_{i} ODEs force pri0p_{r_{i}}\equiv 0. Since the phase condition gives us that cos(θ1θ2)=cos(π)=1\cos(\theta_{1}-\theta_{2})=\cos(\pi)=-1, this in turn gives (after some algebra)

0=p˙r1=δ3m1r22[b(c2/m2)2r2Mb2m2(1+b)2].\displaystyle 0=\dot{p}_{r_{1}}=\frac{\delta^{3}m_{1}}{r_{2}^{2}}\left[\frac{b(c_{2}/m_{2})^{2}}{r_{2}}-\frac{M}{b^{2}}-\frac{m_{2}}{(1+b)^{2}}\right].

Similarly,

0=p˙r2=δ3m2r22[(c2/m2)2r2Mm1(1+b)2].\displaystyle 0=\dot{p}_{r_{2}}=\frac{\delta^{3}m_{2}}{r_{2}^{2}}\left[\frac{(c_{2}/m_{2})^{2}}{r_{2}}-M-\frac{m_{1}}{(1+b)^{2}}\right].

We can eliminate r2r_{2} and the angular momentum c2c_{2} from these equations if we divide each equation by δ3mi/r22\delta^{3}m_{i}/r_{2}^{2} and multiply the p˙r2\dot{p}_{r_{2}} equation by bb. Doing so gives

0\displaystyle 0 =[b(c2/m2)2r2Mb2m2(1+b)2][b(c2/m2)2r2Mbm1b(1+b)2]\displaystyle=\left[\frac{b(c_{2}/m_{2})^{2}}{r_{2}}-\frac{M}{b^{2}}-\frac{m_{2}}{(1+b)^{2}}\right]-\left[\frac{b(c_{2}/m_{2})^{2}}{r_{2}}-Mb-\frac{m_{1}b}{(1+b)^{2}}\right]
=Mb+m1b(1+b)2Mb2m2(1+b)2\displaystyle=Mb+\frac{m_{1}b}{(1+b)^{2}}-\frac{M}{b^{2}}-\frac{m_{2}}{(1+b)^{2}}
=Mb3(1+b)2+m1b3M(1+b)2m2b2b2(1+b)2.\displaystyle=\frac{Mb^{3}(1+b)^{2}+m_{1}b^{3}-M(1+b)^{2}-m_{2}b^{2}}{b^{2}(1+b)^{2}}.

By setting the numerator to zero, dividing by MM, and collecting like powers of bb, we find that

q(b)=0, where q(z)=z5+2z4+(1+m1M)z3(1+m2M)z22z1.q(b)=0,\text{ where }q(z)=z^{5}+2z^{4}+\left(1+\frac{m_{1}}{M}\right)z^{3}-\left(1+\frac{m_{2}}{M}\right)z^{2}-2z-1. (59)

6.3 Analysis of bb

For most values of m1m_{1}, m2m_{2}, and MM, we cannot factor the real polynomial q(z)q(z) since it is a quintic. However, since the masses are all positive, we can analyze its roots. Descartes’ Rule of Signs says that in this case, there is exactly one positive real root, for there is one sign change in the sequence of coefficients of q(z)q(z). Since b>0b>0, it must be the one positive root.

In addition, we see that q(0)=1q(0)=-1, q(z)q(z)\to\infty as zz\to\infty, and q(1)=(m1m2)/Mq(1)=(m_{1}-m_{2})/M. By continuity of qq, we may therefore invoke the Intermediate Value Theorem to place bb within (0,)(0,\infty) according to the relative sizes of m1m_{1} and m2m_{2}:

b{(0,1)if m1>m2{1}if m1=m2(1,)if m1<m2.b\in\begin{cases}(0,1)&\text{if }m_{1}>m_{2}\\ \{1\}&\text{if }m_{1}=m_{2}\\ (1,\infty)&\text{if }m_{1}<m_{2}\end{cases}. (60)

Note that (60) means the case m1=m2m_{1}=m_{2} forces r1=r2r_{1}=r_{2}, in accordance with the COM approach.

In practice, we must obtain bb numerically once values of m1m_{1}, m2m_{2}, and MM are fixed. We found that for all realistic masses (e.g. when m1,m2Mm_{1},m_{2}\ll M), bb remains close to 1 (see Figure 6). To get b>2b>2 or b<0.5b<0.5, we find that the mass ratios μi=mi/M\mu_{i}=m_{i}/M need to be on the order of 20, which is extremely unrealistic.

Figure 7 shows the extreme case of a comet orbit with b=2b=2, to better demonstrate the effect on the asteroid trajectories. For more realistic asteroid masses, the effect is too small to observe on these graph scales.

Refer to caption
Figure 6: Contours showing the value of bb as a function of the input mass ratios m1/Mm_{1}/M and m2/Mm_{2}/M.
Refer to caption
Figure 7: A solution of (62) with b=2b=2 and δ=1\delta=1. To obtain this bb, we had to use extremely large (and thus unrealistic) asteroid masses.

6.4 Periodic Solution to O(δ3)O(\delta^{3})

Once bb is determined, we can solve for r2r_{2} in the pr2p_{r_{2}} ODE:

0=(c2/m2)2r2Mm1(1+b)2r2=(1+b)2(c2/m2)2M(1+b)2+m1.\displaystyle 0=\frac{(c_{2}/m_{2})^{2}}{r_{2}}-M-\frac{m_{1}}{(1+b)^{2}}\implies r_{2}=\frac{(1+b)^{2}(c_{2}/m_{2})^{2}}{M(1+b)^{2}+m_{1}}.

Some algebraic manipulation involving the use of q(b)=0q(b)=0 reveals

r1=br2=(1+b)2(c1/m1)2M(1+b)2+m2b2.r_{1}=br_{2}=\frac{(1+b)^{2}(c_{1}/m_{1})^{2}}{M(1+b)^{2}+m_{2}b^{2}}.

We may also obtain θ˙i\dot{\theta}_{i}:

θ˙i=ω=2πP+δ3[M(1+b)2+m1]2(1+b)4(c2/m2)3=2πP+δ3[M(1+b)2+m2b2]2(1+b)4(c1/m1)3.\dot{\theta}_{i}=\omega=-\frac{2\pi}{P}+\frac{\delta^{3}[M(1+b)^{2}+m_{1}]^{2}}{(1+b)^{4}(c_{2}/m_{2})^{3}}=-\frac{2\pi}{P}+\frac{\delta^{3}[M(1+b)^{2}+m_{2}b^{2}]^{2}}{(1+b)^{4}(c_{1}/m_{1})^{3}}. (61)

Hence we have obtained two periodic circular solutions to (57):

r1(t)\displaystyle r_{1}(t) =br2=(1+b)2(c1/m1)2M(1+b)2+m2b2,\displaystyle=br_{2}=\frac{(1+b)^{2}(c_{1}/m_{1})^{2}}{M(1+b)^{2}+m_{2}b^{2}}, r2(t)\displaystyle r_{2}(t) =(1+b)2(c2/m2)2M(1+b)2+m1,\displaystyle=\frac{(1+b)^{2}(c_{2}/m_{2})^{2}}{M(1+b)^{2}+m_{1}},
θ1(t)\displaystyle\theta_{1}(t) =π+θ0+ωt,\displaystyle=\pi+\theta_{0}+\omega t, θ2(t)\displaystyle\theta_{2}(t) =θ0+ωt,\displaystyle=\theta_{0}+\omega t,
pr1(t)\displaystyle p_{r_{1}}(t) =0,\displaystyle=0, pr2(t)\displaystyle p_{r_{2}}(t) =0,\displaystyle=0,
pθ1(t)\displaystyle p_{\theta_{1}}(t) =c1=m1m2b2c2,\displaystyle=c_{1}=\frac{m_{1}}{m_{2}}b^{2}c_{2}, pθ2(t)\displaystyle p_{\theta_{2}}(t) =c2.\displaystyle=c_{2}. (62)

The period is

T=2π|ω|=2πP(1+b)4(c2/m2)32π(1+b)4(c2/m2)32πPδ3[M(1+b)2+m1]2.T=\frac{2\pi}{|\omega|}=\frac{2\pi P(1+b)^{4}(c_{2}/m_{2})^{3}}{2\pi(1+b)^{4}(c_{2}/m_{2})^{3}-2\pi P\delta^{3}[M(1+b)^{2}+m_{1}]^{2}}.

6.5 Multipliers and Continuation

Since there are more degrees of freedom here in the comet case than the Hill case, we shall work with the full 8×88\times 8 matrix variational equation for the monodromy matrix Ξ(T)\Xi(T). The variational equation is Ξ˙=SΞ,Ξ(0)=I\dot{\Xi}=S\Xi,\Xi(0)=I, where S=J8D2(H0+δ3H3)S=J_{8}D^{2}(H_{0}+\delta^{3}H_{3}) along the solution (62). We found via Mathematica and verified by hand that SS has the following block structure:

S=[ATBCA],S=\begin{bmatrix}-A^{T}&B\\ C&A\end{bmatrix},

where

A=[0A1/b000000000A10000],\displaystyle A=\begin{bmatrix}0&A_{1}/b&0&0\\ 0&0&0&0\\ 0&0&0&A_{1}\\ 0&0&0&0\end{bmatrix}, B=[B10000B20000B30000B4],\displaystyle B=\begin{bmatrix}B_{1}&0&0&0\\ 0&B_{2}&0&0\\ 0&0&B_{3}&0\\ 0&0&0&B_{4}\end{bmatrix}, C=[C110C1300C220C22C130C3300C220C22],\displaystyle C=\begin{bmatrix}C_{11}&0&C_{13}&0\\ 0&C_{22}&0&-C_{22}\\ C_{13}&0&C_{33}&0\\ 0&-C_{22}&0&C_{22}\end{bmatrix},

for the time-independent entries given by

A1\displaystyle A_{1} =2δ3(c2/m2)r23,\displaystyle=\frac{2\delta^{3}(c_{2}/m_{2})}{r_{2}^{3}}, (63)
B1\displaystyle B_{1} =δ3m1,B2=δ3m1b2r22,B3=δ3m2,B4=δ3m2r22,\displaystyle=\frac{\delta^{3}}{m_{1}},\ B_{2}=\frac{\delta^{3}}{m_{1}b^{2}r_{2}^{2}},\ B_{3}=\frac{\delta^{3}}{m_{2}},\ B_{4}=\frac{\delta^{3}}{m_{2}r_{2}^{2}},
C11\displaystyle C_{11} =δ3m1[M(1+b)3+m2b2(3+b)]b3(1+b)3r23,C13=2δ3m1m2(1+b)3r23,\displaystyle=-\frac{\delta^{3}m_{1}[M(1+b)^{3}+m_{2}b^{2}(3+b)]}{b^{3}(1+b)^{3}r_{2}^{3}},\ C_{13}=\frac{2\delta^{3}m_{1}m_{2}}{(1+b)^{3}r_{2}^{3}},
C33\displaystyle C_{33} =δ3m2[M(1+b)3+m1(1+3b)](1+b)3r23,C22=δ3m1m2b(1+b)3r2.\displaystyle=-\frac{\delta^{3}m_{2}[M(1+b)^{3}+m_{1}(1+3b)]}{(1+b)^{3}r_{2}^{3}},\ C_{22}=\frac{\delta^{3}m_{1}m_{2}b}{(1+b)^{3}r_{2}}.

Mathematica gave the characteristic polynomial of SS as being of the form

det(SzI)=D1z2+D2z4+D3z6+z8.\det(S-zI)=D_{1}z^{2}+D_{2}z^{4}+D_{3}z^{6}+z^{8}. (64)

We are interested in D1D_{1} only, but for completeness the coefficients are:

D1\displaystyle D_{1} =B1B3C22[A12b2(b2C11+2bC13+C33)+(B2+B4)(C132C11C33)],\displaystyle=B_{1}B_{3}C_{22}\left[-\frac{A_{1}^{2}}{b^{2}}\left(b^{2}C_{11}+2bC_{13}+C_{33}\right)+(B_{2}+B_{4})\left(C_{13}^{2}-C_{11}C_{33}\right)\right], (65)
D2\displaystyle D_{2} =C22[A12b2(B1+b2B3)+(B2+B4)(B1C11+B3C33)]\displaystyle=C_{22}\left[\frac{A_{1}^{2}}{b_{2}}\left(B_{1}+b^{2}B_{3}\right)+(B_{2}+B_{4})(B_{1}C_{11}+B_{3}C_{33})\right]
B1B3(C132C11C33),\displaystyle\ \ \ \ -B_{1}B_{3}\left(C_{13}^{2}-C_{11}C_{33}\right),
D3\displaystyle D_{3} =[B1C11+(B2+B4)C22+B3C33].\displaystyle=-\left[B_{1}C_{11}+(B_{2}+B_{4})C_{22}+B_{3}C_{33}\right].

Since SS is time-independent, the monodromy matrix is Ξ(T)=exp(ST)\Xi(T)=\exp(ST), like it is in the Hill case, and the multipliers will be ezTe^{zT}, where zz are the roots of (64). So, to show that +1+1 is a multiplier exactly twice, we must show that D1D_{1} is nonzero. Since B1,B3B_{1},B_{3}, and C22C_{22} are each positive, it is sufficient to examine D1/(B1B3C22)D_{1}/(B_{1}B_{3}C_{22}).

We write D1/(B1B3C22)D_{1}/(B_{1}B_{3}C_{22}) in terms of δ,M,b,r2\delta,M,b,r_{2}, and the asteroid mass ratios μi=mi/M\mu_{i}=m_{i}/M, i=1,2i=1,2. Using the expressions for A1A_{1}, B2B_{2}, B4B_{4}, C11C_{11}, C13C_{13}, and C33C_{33} in (63), replacing one of the r2r_{2} from the r29r_{2}^{9} in the common denominator of (A12/b2)(b2C11+2bC13+C33)-(A_{1}^{2}/b^{2})(b^{2}C_{11}+2bC_{13}+C_{33}) with its constant value in the circular periodic solution in (62), and using the quintic equation q(b)=0q(b)=0 in (59) to express each of b5b^{5}, b6b^{6}, and b7b^{7} as linear combinations of 11, bb, b2b^{2}, b3b^{3}, and b4b^{4}, we employed Mathematica, and Maple, and also verified by hand (a lengthy computation), so show that

D1B1B3C22=δ9M3b5(1+b)4r28[k0+k1b+k2b2+k3b3+k4b4],\frac{D_{1}}{B_{1}B_{3}C_{22}}=\frac{\delta^{9}M^{3}}{b^{5}(1+b)^{4}r_{2}^{8}}\left[k_{0}+k_{1}b+k_{2}b^{2}+k_{3}b^{3}+k_{4}b^{4}\right],

where

k0\displaystyle k_{0} =6μ1+3μ2+μ1μ2,\displaystyle=6\mu_{1}+3\mu_{2}+\mu_{1}\mu_{2},
k1\displaystyle k_{1} =15μ1+12μ2+3μ1μ2,\displaystyle=15\mu_{1}+12\mu_{2}+3\mu_{1}\mu_{2},
k2\displaystyle k_{2} =15μ1+18μ2+11μ1μ2+3μ12+μ22μ1μ22,\displaystyle=15\mu_{1}+18\mu_{2}+11\mu_{1}\mu_{2}+3\mu_{1}^{2}+\mu_{2}^{2}-\mu_{1}\mu_{2}^{2},
k3\displaystyle k_{3} =9μ1+12μ2+8μ1μ22μ12+4μ22+2μ12μ2+3μ1μ22,\displaystyle=9\mu_{1}+12\mu_{2}+8\mu_{1}\mu_{2}-2\mu_{1}^{2}+4\mu_{2}^{2}+2\mu_{1}^{2}\mu_{2}+3\mu_{1}\mu_{2}^{2},
k4\displaystyle k_{4} =3μ1+3μ2+2μ1μ22μ12+3μ222μ12μ2.\displaystyle=3\mu_{1}+3\mu_{2}+2\mu_{1}\mu_{2}-2\mu_{1}^{2}+3\mu_{2}^{2}-2\mu_{1}^{2}\mu_{2}. (66)

The quantities k0k_{0} and k1k_{1} in (66) are positive for positive μ1\mu_{1} and μ2\mu_{2}. There are values of μ1>1,μ2>1\mu_{1}>1,\mu_{2}>1 for which each of the quantities k2,k3k_{2},k_{3}, and k4k_{4} in (66) is negative, but for realistic masses (e.g., for μ1,μ2(0,1]\mu_{1},\mu_{2}\in(0,1]), we have

k2\displaystyle k_{2} 15μ1+18μ2+10μ1μ2+3μ12+μ22>0,\displaystyle\geq 15\mu_{1}+18\mu_{2}+10\mu_{1}\mu_{2}+3\mu_{1}^{2}+\mu_{2}^{2}>0,
k3\displaystyle k_{3} 7μ1+12μ2+8μ1μ2+4μ22+2μ12μ2+3μ1μ22>0,\displaystyle\geq 7\mu_{1}+12\mu_{2}+8\mu_{1}\mu_{2}+4\mu_{2}^{2}+2\mu_{1}^{2}\mu_{2}+3\mu_{1}\mu_{2}^{2}>0,
k4\displaystyle k_{4} μ1+3μ2+3μ22>0.\displaystyle\geq\mu_{1}+3\mu_{2}+3\mu_{2}^{2}>0.

Hence, D1>0D_{1}>0 for all μ1,μ2(0,1]\mu_{1},\mu_{2}\in(0,1].

All of the nonzero entries of SS are a multiple of δ3\delta^{3}, as follows from (63). The matrix S^=δ3S\hat{S}=\delta^{-3}S does not depend on δ\delta and the eigenvalues of S^\hat{S} are determined by roots of

D^1z2+D^2z4+D^3z6+z8=z2(D^1+D^2z2+D^3z4+z6),\hat{D}_{1}z^{2}+\hat{D}_{2}z^{4}+\hat{D}_{3}z^{6}+z^{8}=z^{2}(\hat{D}_{1}+\hat{D}_{2}z^{2}+\hat{D}_{3}z^{4}+z^{6}),

where coefficients D^i\hat{D}_{i}, i=1,2,3i=1,2,3 are expressions, independent of δ\delta, given by the formulas in (65) in which each of the scalars A1A_{1}, B1B_{1}, B2B_{2}, B3B_{3}, B4B_{4}, C11C_{11}, C13C_{13}, C22C_{22}, and C33C_{33} in (63) are each multiplied by δ3\delta^{-3} before being substituted. The quantities D1D_{1} and D^1\hat{D}_{1} are related by δ9D^1=D1\delta^{9}\hat{D}_{1}=D_{1}, so that D1>0D_{1}>0 implies D^1>0\hat{D}_{1}>0. Hence S^\hat{S} has eigenvalue 0 with algebraic multiplicity two. For two circular periodic solutions in (62), one with a positive and one with a negative choice of c2c_{2}, say c2=±1c_{2}=\pm 1 (where c1=±b2(m1/m2)c_{1}=\pm b^{2}(m_{1}/m_{2}) as determined by (58)), an adaptation of the argument used in Meyer and Offin [14] can now be applied to prove continuation for all small positive values of δ\delta.

Theorem 7 (Comet Orbits).

There exist two one-parameter families of near-circular comet-type periodic solutions in the Binary Asteroid Problem (6) for all realistic masses (m1/M,m2/M(0,1])(m_{1}/M,m_{2}/M\in(0,1]). These families tend to infinity.

Refer to caption
Figure 8: Depiction of a numerical δ\delta continuation process for a comet-type orbit (62) with b=1.0001b=1.0001, M1=95M_{1}=95, and M2=5M_{2}=5.
Remark 4.

This comet approach can be adapted to develop Hill-type equations of motion similar in form to (57). The main difference is that we would employ symplectic shifts to center each asteroid on one primary, then scale. Doing so produces circular Hill solutions out to δ3\delta^{3}. However, multiplier analysis of these solutions shows they have +1+1 as a multiplier four times, and hence they do not continue. We believe that the lack of continuation in this case is due to the coupling term of the potential not being present in the equations of motion until O(δ6)O(\delta^{6}).

7 Conclusion

We have presented a planar 4-body model, the Binary Asteroid Problem, with Hamiltonian HBAH_{BA} in (6), and its COM reduced Binary Asteroid Problem with Hamiltonian HxyH_{xy} in (10). In the COM reduced Binary Asteroid Problem we have the existence of six relative equilibria (Theorem 2), four of which are saddle-centers (Theorem 3) and collinear with the primaries, and two of which are hyperbolic (Theorem 3) and form convex kites (and each is a rhombus) with the primaries. Near each saddle-center relative equilibrium there is a one-parameter family of periodic orbits (Corollary 4). Each relative equilibrium limits to its corresponding relative equilibrium in the Generalized Copenhagen Problem as the common mass of the asteroids goes to 0 (Theorem 2). In the COM reduced Binary Asteroid Problem we have two one-parameter families of near-circular periodic orbits near the origin (Theorem 5) on which the two asteroids form a binary asteroid. Also in the COM reduced Binary Asteroid Problem we found two one-parameter families of near-circular Hill-type orbits (Theorem 6) where one asteroids orbits one primary and the other asteroid orbits the other primary. In the Binary Asteroid Problem we found two families of near circular periodic comet-type orbits where the two asteroids are approximately opposite each other through the origin (Theorem 7).

The approaches used in this paper can be adapted to produce results in the spatial Binary Asteroid Problem. We have numerical simulations showing spatial continuation of the Hill-type and comet-type orbits. However, multiplier analysis is much more difficult because the variational equation for each of these spatial orbits becomes time-dependent. In addition to questions (1), (2), (3), (4), and part of (5) posed in the Introduction, future research will also include the study of these Hill-type and comet-type orbits in the spatial Binary Asteroid Problem, and the continuation of planar and spatial periodic orbits in the Binary Asteroid Problem into the four-body problem.

References

  • [1] M. Alvarez-Ramírez and J. Llibre, The symmetric central configuration of the 44-body problem with masses m1=m2m3=m4m_{1}=m_{2}\neq m_{3}=m_{4}, Appl Math Comput 219 (2013), 5996–6001.
  • [2] M. Alvarez-Ramírez, J.E.F. Skea, and T.J. Stuchi, Nonlinear stability analysis in a equilateral restricted four-body problem, Astrophys Space Sci 358 (2015), no. 3.
  • [3] L. F. Bakker and S. Cochran, Periodic solutions in the planar 55-body problem with two free bodies, Undergraduate Research, 2021.
  • [4] L. F. Bakker, J. Murri, and S. Simmons, A model for the binary asteroid 2017 ye5, In preparation, 2023.
  • [5] C. Cofield and J. Wendel, Observatories team up to reveal rare double asteroid, NASA, 2018, Online article at https://www.nasa.gov/feature/jpl/observatories-team-up-to-reveal-rare-double-asteroid.
  • [6] M. Corbera and J. Llibre, Central configurations of the 44-body problem with masses m1=m2>m3=m4>0m_{1}=m_{2}>m_{3}=m_{4}>0 and small, Appl Math Comput 246 (2014), 121–147.
  • [7] J. Davis, Planetary society asteroid hunters help find rare type of double asteroid, The Planetary Society (2018), Online article at https://www.planetary.org/articles/shoemaker-winners-2017-ye5.
  • [8] D. Espitia, E. A. Quintero, and I. D. Arellano-Ramírez, Determination of orbital elements and ephemerides using the geometric laplace’s method, J Astron Space Sci 37 (2020), no. 3, 171–185.
  • [9] N. J. Freeman, Investigations of a binary asteroid dynamical model, April 2023, Senior Thesis, Department of Physics and Astronomy, Brigham Young University.
  • [10] W. R. Johnston, Asteroids with satellites, 2023, Online Database at https://www.johnstonsarchive.net/astro/asteroidmoons.html.
  • [11] J. Llibre, D. Paşca, and C. Valls, The circular restricted 4-body problem with three equal primaries in the collinear central configuration of the 3-body problem, Celestial Mechanics and Dynamical Astronomy 133 (2021), no. 53.
  • [12] J. Llibre and C. Stoica, Comet- and hill-type periodic orbits in restricted (n+1)(n+1)-body problems, J Differential Equations 250 (2011), 1747–1766.
  • [13] Y. Long and S. Sun, Four-body central configurations with some equal masses, Arch Rational Mech Anal 162 (2002), 25–44.
  • [14] K. R. Meyer and D. C. Offin, Introduction to hamiltonian dynamical systems and the n-body problem, 3rd ed., Applied Mathematical Sciences, vol. 90, Springer International Publishing, 2017.
  • [15] F. Monteiro, E. Rondón, D. Lazzaro, J. Oey, M Evangelista-Santana, P. Avcoverde, M DeCicco, J. S. Silva-Cabrera, T. Rodrigues, and L. B. Santos, Physical characterization of equal-mass binary near-earth asteroid 2017 ye5: a possible dormant jupiter-family comet, Monthly Notices of the Royal Astronomical Society 507 (2021), 5403–5414.
  • [16] A. E. Roy and B. A. Steves, Some special restricted four-body problems-ii: From caledonia to copenhagen, Planet Space Sci 46 (1998), no. 11-12, 1475–1486.
  • [17] D. J. Scheeres and J. Bellerose, The restricted hill full 44-body problem: application to spacecraft motion about binary asteroids, Dyn Syst 20 (2005), no. 1, 23–44.
  • [18] J. Lu H. Shang and B. Wei, Accelerating binary asteroid system propagation via nested interpolation method, Celest Mech Dyn Astron 135 (2023).
  • [19] M. Shoaib and I. Faye, Collinear equilibrium solutions of four-body problems, J Astrophys Astr 32 (2011), 411–423.
  • [20] C. Simó, Relative equilibrium solutions in the four body problem, Celestial Mechanics 18 (1978), 165–184.
  • [21] V. Szebehely, The theory of orbits, Academic Press, New York, 1967.
  • [22] H.S. Wang and X. Y. Hou, Forced hovering orbit above the primary in the binary asteroid system, Celest Mech Dyn Astron 134 (2022).