1. Introduction
The study of the n-body problem on curved spaces, especially on spheres, introduces unique challenges and phenomena not present in Euclidean spaces. The work of Borisov et al. [
1] provides essential insights into the dynamics of bodies on spaces of constant curvature, foundational for our investigation.
In the realm of celestial mechanics, understanding the intricacies of motion in non-Euclidean geometries is crucial. The research conducted by Ortega-Palencia and Reyes-Victoria [
2] and further expanded by Ortega Palencia et al. [
3] delves into the n-body problem in spaces of constant positive and negative curvature, offering valuable perspectives that inform our approach.
Our study is also informed by the analysis presented by Diacu et al. [
4], which explores the n-body problem in spaces of constant curvature. Their findings provide a broader context for our work, emphasizing the diversity of dynamics that different geometric settings can induce.
The foundational principles laid out by Abraham and Marsden [
5] in their seminal work on mechanics provide the theoretical underpinnings for analyzing dynamical systems in a geometric context, essential for our study.
Additionally, the exploration of antipodal equilibria in the two-dimensional sphere by Ortega-Palencia et al. [
6] and their further discussion in the arXiv preprint [
7] offer important precedents for our study. These works highlight the peculiarities of motion in curved spaces and the significance of the chosen potential in determining the system’s behavior.
Through this research, we aim to build on these foundational studies, extending the understanding of the n-body problem in curved spaces and exploring new facets of relative equilibria and their stability. Our work is particularly inspired by the recent developments in the field and seeks to contribute to the ongoing dialogue within the scientific community regarding the dynamics of celestial bodies in non-traditional settings.
We want to point out the possibility of having a relative equilibrium in antipodal points, as stated in [
8].
In celestial mechanics with curvature, specifically in the case of a sphere, there are two types of singularities related to the classical cotangent potential [
8]. One refers to collisions, while the other refers to antipodal points.
Various researchers, including those cited as [
9,
10], have explored the behavior of collisions in spaces with non-zero constant curvature by applying classical regularization methods from Newtonian mechanics. Their findings align with those previously obtained in this domain, marking the beginning of investigations into dynamic types.
Additionally, in classical celestial mechanics, an alternative system that helps understand certain movements is studied, such as the planar three-body problem with an attractive potential of
[
11,
12]. According to the Lagrange–Jacobi identity, where
, for any solution that is bounded, it is required to possess zero energy and maintain a constant moment of inertia
I. This condition holds when the energy at the initial state is zero and its derivative is zero, the solution is bounded, as observed in the Newtonian potential
.
The work on how to carry out the study of behavior analysis around a geometric singularity, that is, by antipodal points, is just beginning. In this work, a geometric method is proposed to be able to study this type of singularity in the antipodal points for the problem of two bodies in a conformal sphere of dimension two, with one alternative potential that satisfies the Laplace–Beltrami equality [
13] and preserves the periodic orbits as the classic cotangent potential in the curved problem.
Our study focuses on the motion of two interacting point particles with masses
and
on the two-dimensional conformal sphere
. Consider the complex variable
w and its conjugate
, upon which we define a conformal metric of the form
where
R is a constant parameter determining the conformal properties of the metric.
We define the corresponding differentiable structure for this space in these coordinates (for more details, refer to [
14,
15]).
By aligning the vector field within the Lie algebra that corresponds to the associated subgroup with the gravitational field in the cotangent space, we determine the time-dependent algebraic criteria (t) required for solutions to achieve a state of relative equilibrium.
The techniques used in [
8,
16] are employed in this analysis.
In celestial mechanics on , it is common to use the cotangent potential as an extension of the Newtonian potential. This paper presents arguments justifying the introduction of a suitable variant of this potential.
This document is organized as follows:
For an overview of the equations of motion, see
Section 2, we have developed a revised potential to tackle the issue, successfully overcoming the challenge of singularities at antipodal points.
We express the equations of motion for the problem in complex coordinates in
, following the approach used in [
8,
16].
For insights into elliptic relative equilibria and their properties, consult
Section 3: utilizing the newly introduced potential, we formulate the algebraic equations that define the elliptic relative equilibria in the general problem context.
For a detailed exploration of the two-body problem, refer to
Section 4: we classify the relative equilibria for the two-body problem, following the approach of Borisov et al. in [
17], but considering the new potential.
Regarding the analysis of antipodal points, see
Section 5: for any antipodal pair of points in
, we successfully derive limiting solutions by applying a regularized version of the original equations of motion.
2. Dynamics Formulation and Conditions for Equilibrium
As discussed in [
8], utilizing the stereographic projection method, the authors formulate the motion equations pertinent to this problem. This involves projecting the sphere (with a radius of
R) from its embedding in
onto the complex plane
, which utilizes the specified metric (
1).
In Ref. [
5], the classical motion equations for particles with positive masses
,
are discussed, situated within a Riemannian or semi-Riemannian manifold characterized by coordinates
, endowed with a metric
and an associated connection
. The connection here, specifically the Levi-Civita connection, is not arbitrary. It is determined by the specified metric (
1) through the Christoffel symbols, which are uniquely defined by the metric to ensure the connection is torsion-free and metric-compatible.
Replacing the Levi-Civita connection with an arbitrary one could lead to a different set of Christoffel symbols, altering the motion equations and, consequently, the particles’ trajectories and equilibrium states. Such changes would reflect a fundamental shift in the geometric structure of the space in which the particles are moving.
These particles move under the influence of a pairwise-acting potential U.
Theorem 1. Consider a system of n particles with positive masses , situated within a Riemannian or semi-Riemannian manifold characterized by coordinates , and endowed with a metric as specified in (1) and an associated Levi-Civita connection . The equations of motion for these particles under the influence of a pairwise-acting potential U are given bywhere . If the potential U is constant across a connected domain, the particle trajectories align with the geodesics of the manifold defined by the metric (1). Proof. Given the manifold’s metric
specified in (
1) and the associated Levi-Civita connection
, the covariant derivative
accounts for the curvature dictated by (
1) and ensures that the acceleration
is defined in a coordinate-independent manner. The equation of motion incorporates both the intrinsic geometry of the manifold, represented by the Christoffel symbols
, and the external force derived from the potential
U. When
U is constant, the term
vanishes, indicating that the particles’ acceleration is solely dictated by the manifold’s geometry as defined by (
1), thus aligning their trajectories with geodesics. □
Remark 1. It is noted that in Equation (2) the covariant derivative of is represented on the left-hand side, whereas the gradient of the potential within the specified metric is depicted on the right-hand side. Should the potential remain constant, particle trajectories align with geodesics. If a set of particles moves along a geodesic solution curve, then the right-hand side of Equation (2) vanishes, such that the solution of the potential is constant on one connected domain. In this section, we introduce the motion equations for the n-body problem within the conformal sphere denoted as .
2.1. Introducing the Novel Potential
Exploring the
n-body problem on the sphere, we reference a potential frequently encountered in contemporary research, expressed as follows:
in which
is the angle at the center of the sphere, delineated by the position vectors of the particles. This potential is attractive because
,
.
Here, we introduce a slight modification to the potential (
3) and define it as the new potential:
which remains attractive over the entire interval
.
Theorem 2. Consider a pair of particles with masses and positioned at locations and on the sphere . If and are their respective stereographic projections onto , then the potential experienced by the particles is given by Proof. Given the particles at
and
on
, the geodesic distance
between them is related to the angle
at the origin by
. From the trigonometric identity on the sphere, we have
Applying the law of cosines for spherical trigonometry gives us
Using the stereographic projection, the dot product
in terms of
and
is
Substituting this into the cotangent expression and simplifying yields the potential experienced by the particles as
which completes the proof. □
2.2. Equations of Motion
Designate as the collective position vector for n particles, each with mass , situated at points for , within the space .
The set of singularities within
for the
n-body problem, as defined by the cotangent relation, comprises the solutions to the equation
. Proceeding from this point, the singular set is identified as
where
represents instances of mutual collisions between particles having masses
and
.
Theorem 3 (Dynamics of n-Body Problem in
)
. Consider a set of n point particles with masses situated at points within the Riemannian manifold , where represents the collective position vector of these particles. The dynamics of each particle is governed by the following equation:for . Proof. The proof follows from integrating the geodesic equations for and the gradients computed from the potential , as detailed in the equations provided, into the dynamics defined by the Vlasov–Poisson equations. The resulting second-order complex ordinary differential equations dictate the motion of the particles, ensuring that the trajectories remain within the specified domain, avoiding the singular set . □
Corollary 1 (Singular Set and Binary Collisions in
)
. In the space , the singular set for the n-body problem consists solely of points satisfying for any pair of particles where , indicating binary collisions. Antipodal points satisfying are not considered part of the singular set in terms of the equations of motion.
Proof. The characterization of the singular set stems from the definition of mutual collisions between particles, which are the only points where the potential becomes undefined or singular. The exclusion of antipodal points from the singular set is due to the specific structure of the potential and the forces it dictates, which remain well defined for antipodal configurations. □
3. Elliptic Relative Equilibria
The group is given by the Lie algebra generated by three matrices; for our purposes we only work with the complex matrix:
Theorem 4. In the two-dimensional conformal sphere , the one-parameter subgroup generated by , where , induces a family of elliptic Möbius transformations. For any point w in the disk of radius R in the complex plane, the trajectory under this transformation is a circular path given by , representing a rotation around the z-axis in . This circular motion corresponds to the differential equation , signifying the dynamical system’s relative equilibrium state.
Proof. The exponential mapping applied to the line yields the one-parameter subgroup . The modulus of the off-diagonal elements, , is equal to 1, characterizing elliptic transformations since for . The action of this subgroup on a point w in is described by the Möbius transformation , which geometrically corresponds to a rotation around the z-axis. The trajectory of w under this transformation is a circular path in , aligning with the differential equation . This demonstrates that the system is in a state of relative equilibrium, as the trajectories are circular paths dictated by the subgroup’s rotational action. □
The trajectories generated by the one-parameter subgroup
in
, along with the circular paths on the two-dimensional sphere situated in
, are illustrated in
Figure 1 and
Figure 2.
Now, we can start our analysis of the so-called
solutions of elliptic relative equilibria, derived from the influence of the canonical one-dimensional parametric subgroup within
, corresponding to the differential equation
. There is a study of the action of the subgroup in [
2,
8,
17] for the classic cotangent potential.
Definition 1. An elliptic relative equilibrium for the n-body problem in is a solution of the equations of motion (8) that is invariant under the Killing vector field . We now present the following lemma:
Theorem 5. In the case of n point particles, each with positive masses , moving within , the requisite condition for to qualify as an elliptic relative equilibrium solution under (8) is encapsulated by the subsequent rational complex functional equations, which are time-dependent:where the velocity at each point is given by , where represents the value of the k-th component of the vector . Proof. Through direct calculations, we find that from equation
, we have
. Substituting this into Equation (
8) yields Equality (
9). □
The subsequent finding outlines prerequisites for the particles’ initial placements to yield an elliptic relative equilibrium solution for Equation (
9). These solutions depend on the fixed points and their velocities.
Corollary 2. In line with Theorem 5, the initial positions fulfill a necessary and sufficient criterion to produce an elliptic solution for the system (8), which remains invariant under the Killing vector field , through this set of algebraic equations: Furthermore, the required velocity for each particle is determined by the equation , where k ranges from 1 to n, where represents the initial value of the k-th component of the vector .
Proof. Take
to represent the impact of the Killing vector field
at the initial position
, corresponding to a velocity of
. By applying a multiplication of Equation (
10) with
and incorporating the identity
, the resultant system is derived:
This demonstrates that
serves as a solution to (
9). To establish the reverse argument, one simply sets
within the framework of (
9), thereby concluding the corollary’s proof. □
References [
2,
8] provide illustrative examples of the two- and three-body problems situated on the conformal sphere
with the classical cotangent potential.
4. Equilibria States in the Two-Body Problem Context
In the following segment, it is shown that relative equilibria exist in the context of the two-body problem, specifically when employing potential (
5), where the bodies are in motion on the same circle or on two different circles. These results are consistent with the findings in [
1,
2,
8] for the classical cotangent potential.
Theorem 6. For the values of initial condition positions and (with ), for the two-body problem with equal masses on the conformal sphere , the system (10) yields only two types of relative equilibrium solutions: - 1.
Under the condition , the particles position themselves diametrically opposite on the same circle, with , termed as isosceles solutions (refer to Figure 1). - 2.
Given the condition , the particles occupy positions on separate circles, with , creating a right angle, known as right-angled solutions (refer to Figure 2). - 3.
Both types of relative equilibria coincide for the value of .
The solutions for β are in the interval .
Proof. Firstly, in positions
and
, we note that the system (
10) for the two-body problem can be expressed as the following algebraic system:
From Corollary 2, by performing a suitable rotation, a condition both necessary and sufficient for the presence of invariant elliptic solutions influenced by the Killing vector field
dictates that the initial positions
and
(where
) must conform to system (
11). Since
, when substituted into the system, it becomes
By equalizing the right- and left-hand sides of (
12) and replacing
,
, the following relation is derived:
For
the solutions to the system outlined in (
12), derived from Equation (
13), are as follows:
which is readily apparent.
For the initial scenario where
, the result is the isosceles configurations, and by substituting it into any of the equations in system (
12), we obtain the relation
We consider the function
defined in the interval
. It has a maximum value of
at the critical point
.
A simple analysis shows that there are isosceles solutions for this problem if , which proves the first item.
In the subsequent case, consider ℓ as the geodesic distance between and .
First, let us establish that is always smaller than within the interval . To determine which is smaller in the interval , we compare with .
If , then , which is clearly smaller than since .
If , then . This implies that is negative, and therefore, smaller than , which is positive.
We analyze the difference
:
Since and for , the difference is positive, indicating that is greater than .
Therefore, in all cases within the interval , the point is always smaller than .
By defining the arc
that connects points
and
through a parametrization given by
over the interval
, the geodesic distance
ℓ can be calculated as follows:
Now, let us consider the asymptotic behavior of ℓ as approaches R and 0:
1. As approaches R, the geodesic distance ℓ should theoretically approach 0 as the two points converge.
2. As approaches 0, the point approaches , and the geodesic distance ℓ should approach the maximum possible value on the circle, which is , corresponding to half the circumference of the circle of radius R.
This analysis provides a deeper understanding of the geometric configuration of the system and the behavior of the geodesic distance under different conditions.
This delineates the process and calculation of length ℓ for the specified path within the two-dimensional manifold .
Conversely, it is established that , with denoting the angle between the specified points in . Consequently, , classifying the solution as right-angled.
In this case, by substituting
into any of the equations in the system (
12), we obtain the relation
We consider the function
defined in the interval
. It has a maximum value of
at the critical point
. Once again, a simple analysis shows that there are right-angled solutions for this problem if
, which proves the second item.
This concludes the proof. □
In [
1], in theorem 4.3 there is a discussion about the stability of the solutions obtained.
In this theorem, we establish a connection between the relative equilibria for the two-body problem and Snell’s law of geometric optics.
Theorem 7. If two particles in the two-body problem are in relative equilibrium and the given substitution conditions are met, then the relationship between their masses and positions is analogous to Snell’s law, with the particularity that the “indices of refraction” are the masses and the “angles of refraction” are related to the positions of the particles.
Proof. Consider the following equation derived from the polynomial equation for the two-body problem (see Equation (
13)):
Here,
and
. Let us substitute
and
into Equation (
20):
Simplifying the equation, we obtain
Using the trigonometric identity
, we can rewrite the equation as
Now, using the double-angle formula
, we can express
in terms of
:
This relationship is analogous to Snell’s law, where the masses act as indices of refraction and the angles are related to the positions of the particles, thus proving the theorem. □