Fact-checked by Grok 2 weeks ago

n-body problem

The n-body problem is the challenge in of determining the motions of n point masses interacting solely through , typically formulated as a system of coupled second-order ordinary differential equations describing their positions and velocities over time. For n = 1 or n = 2, the problem admits closed-form analytic solutions: the one-body case is trivial, consisting of uniform rectilinear motion, while the reduces to an equivalent one-body problem in a central field, yielding conic-section orbits as described by Kepler's laws. However, for n ≥ 3, no general analytic solution exists in terms of elementary functions, rendering the system generally non-integrable and prone to chaotic behavior, though special solutions like periodic orbits and central configurations have been identified. The problem originated with Newton's Philosophiæ Naturalis Principia Mathematica (1687), where he first posited universal gravitation and solved the two-body case, but recognized the complexity of planetary perturbations in the solar system. In the , Euler and Lagrange advanced the field by discovering equilibrium configurations (central configurations) and relative equilibria, such as the Lagrangian points in the restricted . The saw efforts by Laplace to prove solar system stability via , but Henri Poincaré's work in the late 1880s, particularly in his prize essay for King Oscar II, revealed the inherent instability and non-integrability for n ≥ 3, laying foundations for . Mathematically, the problem is expressed in Hamiltonian form, with the H = T + V, where T is the and V is the V(q) = -∑_{i<j} m_i m_j / |q_i - q_j| (with gravitational constant G = 1), leading to equations of motion \ddot{q}i = ∑{j ≠ i} m_j (q_j - q_i) / |q_j - q_i|^3 for positions q_i ∈ ℝ³. Conserved quantities include total energy, linear momentum, angular momentum, and the center-of-mass motion, allowing reduction to relative coordinates, but the reduced system remains high-dimensional and nonlinear for n > 2. In modern applications, the n-body problem underpins simulations of solar system dynamics, stability, star cluster evolution, and formation, relying on numerical integrators like methods to approximate long-term behavior despite . Breakthroughs include the discovery of non-trivial periodic solutions, such as the figure-eight orbit for the in 1993 by Chenciner and Montgomery, and KAM theory results ensuring for nearly integrable cases with small perturbations. These insights highlight the problem's enduring role in understanding gravitational systems, from trajectory planning to cosmological models.

Formulation

General Equations

The n-body problem is formulated within , where n point masses interact solely through pairwise gravitational forces governed by Newton's . The arise directly from Newton's second law and the law of universal gravitation. For the i-th body with mass m_i and position vector \mathbf{r}_i(t) \in \mathbb{R}^3, the acceleration is given by the gravitational attractions from all other bodies: m_i \frac{d^2 \mathbf{r}_i}{dt^2} = G \sum_{j=1, j \neq i}^n m_i m_j \frac{\mathbf{r}_j - \mathbf{r}_i}{|\mathbf{r}_j - \mathbf{r}_i|^3}, where G is the gravitational constant. This yields a system of $3n second-order ordinary differential equations (ODEs) in the $3n-dimensional configuration space of the positions \{\mathbf{r}_i(t)\}_{i=1}^n, or equivalently $6n first-order ODEs in the full phase space including velocities. Due to the translation and rotation invariance of the gravitational potential, as well as time-independence, the system admits exactly 10 classical integrals of motion in three dimensions: the total energy (1 scalar), the total linear momentum vector (3 components), the total angular momentum vector (3 components), and the center-of-mass position vector (3 components).

Assumptions and Boundary Conditions

The n-body problem in models interacting as point masses possessing no physical size or internal structure, thereby simplifying the dynamics to purely gravitational interactions without considerations of tidal deformations or extended body effects. These point masses obey , where the force between any two is inversely proportional to the square of their separation distance and acts instantaneously without retardation effects, distinguishing the classical formulation from relativistic treatments that incorporate finite propagation speeds for gravitational influences. The is strictly Newtonian, scaling as the inverse of the distance between masses, and the system excludes non-gravitational forces such as electromagnetic interactions or relativistic corrections unless explicitly modified. Initial conditions for the n-body problem are specified by assigning initial positions \mathbf{r}_i(0) and velocities \mathbf{v}_i(0) to each of the n at time t=0, which uniquely determine the subsequent trajectory in the 6n-dimensional comprising all positions and momenta. These conditions define an (IVP), where the evolve the system forward (or backward) in time from the given state, contrasting with boundary value problems that might constrain states at multiple times. The is equipped with a structure arising from the formulation, ensuring that the flow preserves the symplectic form and thus maintains key invariants like and in numerical simulations. Boundary conditions in the standard n-body problem assume an evolving in infinite with no external potentials or confining boundaries, allowing unbounded motions such as hyperbolic escapes. In restricted variants, such as planetary problems, one or more central may be fixed at specified positions to approximate dominant influences, effectively imposing Dirichlet-like boundaries on their motion while permitting perturbations among lighter . This setup highlights a conceptual hierarchy: the full n-body IVP lacks closed-form boundary constraints beyond conservation laws, relying instead on the of to constrain long-term behavior without resolving the general integrability issue.

Historical Development

Early Formulations

The roots of the n-body problem lie in models of celestial motion, which sought to explain the observed paths of heavenly bodies through geometric frameworks rather than gravitational interactions. (384–322 BCE) proposed a geocentric divided into terrestrial and celestial realms, with the latter consisting of unchanging, spherical motions driven by "unmoved movers" and composed of a , ; this model assumed uniform circular orbits for the , Sun, planets, and , laying a qualitative precursor to multi-body by implying harmonious interactions among . Building on Aristotelian cosmology, Claudius Ptolemy (c. 90–168 CE) developed a more predictive geocentric system in his , using epicycles—smaller circles whose centers moved along larger deferent circles—to account for the irregular apparent motions of planets relative to ; this introduced mathematical complexity to model interactions among multiple bodies, such as the retrograde motion of Mars, though it remained kinematic without a unifying force law. Ptolemy's framework dominated for over a millennium, influencing later recognition of perturbations in planetary paths as arising from mutual influences beyond simple two-body approximations. Johannes Kepler advanced the empirical foundation in the early by deriving three laws of planetary motion from Brahe's observations, providing a precise description of two-body (Sun-planet) dynamics while implicitly revealing multi-body challenges. The first law (elliptical orbits with the Sun at one focus) and second law (equal areas swept in equal times) appeared in (1609), while the third law (period squared proportional to semi-major axis cubed) was published in (1619); these laws idealized Sun-planet isolation but Kepler noted deviations, such as in the Moon's , hinting at perturbations from additional bodies like . Isaac Newton's (1687) marked the first explicit mathematical formulation of the n-body problem through his law of universal gravitation, stating that every particle attracts every other with a force proportional to the product of their masses and inversely proportional to the square of their distance; this generalized Kepler's laws to arbitrary numbers of interacting bodies, enabling qualitative analysis of planetary systems. In Book 1, Section 11, Newton discussed perturbations qualitatively, such as how a third body (e.g., ) disturbs the elliptical orbit of a primary pair (Earth-Moon), deriving tendencies like without full solutions due to the problem's inherent complexity. A pivotal of these ideas was Newton's "moon test," a comparing the of a falling apple (about 9.8 m/s² at 's surface) to the Moon's centripetal (derived from its orbital and , yielding roughly 0.0027 m/s², consistent with inverse-square scaling over 60 radii); this , recounted in Newton's later writings, demonstrated universal gravity's reach from terrestrial to celestial scales and highlighted three-body effects, where the Sun's pull on 's oceans—analogous to deforming a fluid body like an —causes bulges and .

Key Theoretical Advances

In the 1760s, Leonhard Euler made significant early advances in the by identifying collinear central configurations, where the three bodies remain aligned throughout their motion, and introducing the first integrals of motion, such as and conservation, to constrain the system's . These contributions, detailed in his memoirs on , provided analytical tools for special cases but highlighted the challenge of general solutions. Joseph-Louis Lagrange built upon this in 1772 with his "Essai sur le problème des trois corps," where he discovered equilateral triangular central configurations, now known as Lagrangian points, in the restricted ; these are stable equilibrium points where a third body can remain relative to two orbiting primaries. Lagrange's formulation reduced the degrees of freedom through symmetries and introduced the Lagrangian equations of motion, laying groundwork for in . In the late 18th and early 19th centuries, further developed to analyze the mutual gravitational influences among planets. In his Traité de Mécanique Céleste (1799–1825), he used series expansions to calculate planetary orbits and sought to prove the long-term , demonstrating that the eccentricities and inclinations of planetary orbits remain small and nearly constant over time. In the late , revolutionized the understanding of the n-body problem for n ≥ 3, particularly through his prize essay for King Oscar II submitted in 1888 and revised in 1890, and in "Les méthodes nouvelles de la mécanique céleste" (1892–1899), by proving its non-integrability through qualitative analysis, demonstrating that no additional analytic integrals exist beyond the classical ten, and introducing the recurrence theorem that implies dense orbital filling in . His work on periodic solutions and in the restricted laid foundations for modern . Karl Fritiof Sundman achieved a milestone in 1912 by constructing a global analytic solution to the using a power series expansion that converges for all time, provided the total angular momentum is non-zero; this regularization handled collisions and extended solutions beyond finite times. Although the series converges too slowly for practical computation, Sundman's approach generalized to the n-body problem and affirmed the existence of uniform solutions. The mid-20th century saw the Kolmogorov-Arnold-Möser (KAM) theorem emerge as a key advance for perturbed n-body systems, with Andrey Kolmogorov's 1954 proof showing that most quasi-periodic orbits persist under small perturbations in nearly integrable systems, later refined by in 1963 and Jürgen Moser in 1962. Applied to , KAM theory explains long-term stability in planetary systems by preserving invariant tori, contrasting Poincaré's results for larger perturbations.

Special Cases

Two-Body Problem

The describes the motion of two point masses interacting solely through a central gravitational , serving as the only exactly solvable case within the broader n-body framework. By transforming to the center-of-mass frame, where the total momentum is zero, the system separates into the uniform motion of the center of mass and the relative motion of the two bodies. This reduction equates the two-body dynamics to an equivalent one-body problem, in which a fictitious particle of \mu = \frac{m_1 m_2}{m_1 + m_2} orbits a fixed central M = m_1 + m_2 under the . The equation of motion for the relative vector \mathbf{r} = \mathbf{r}_1 - \mathbf{r}_2 in this reduced system is given by \mu \frac{d^2 \mathbf{r}}{dt^2} = -\frac{[G](/page/Gravitational_constant) m_1 m_2}{|\mathbf{r}|^3} \mathbf{r}, where [G](/page/Gravitational_constant) is the . Due to the central nature of the force, both total energy E and \mathbf{L} are conserved, restricting the relative orbit to a plane perpendicular to \mathbf{L}. Solving the radial equation using these constants yields the general orbit as a conic section: for bound motion (E < 0), parabola for marginal escape (E = 0), or hyperbola for unbound trajectories (E > 0). The specific shape is parameterized by the e = \sqrt{1 + \frac{2 E L^2}{\mu (G m_1 m_2)^2}}, with the r(\theta) = \frac{L^2 / (\mu [G](/page/Gravitational_constant) m_1 m_2)}{1 + e \cos \theta}. This closed-form analytic solution enables exact prediction of trajectories without approximation, distinguishing the two-body case from higher-order problems. Kepler's laws emerge as direct consequences: the first, elliptical orbits with the central body at one focus, follows from the conic form for e < 1; the second, equal areas swept in equal times, arises from constant areal velocity \frac{dA}{dt} = \frac{L}{2\mu}; and the third, T^2 \propto a^3 where T is the period and a the semi-major axis, derives from energy conservation relating a = -\frac{G m_1 m_2}{2E} to the orbital period T = 2\pi \sqrt{\frac{a^3}{G(m_1 + m_2)}}.

Three-Body Problem

The three-body problem concerns the motion of three point masses under mutual gravitational attraction according to . While the two-body problem admits a closed-form solution, the general three-body problem does not, marking it as the simplest non-integrable case in classical mechanics. The system's phase space is 18-dimensional, corresponding to the three-dimensional positions and momenta of each body. There are exactly 10 classical integrals of motion—namely, the total energy, the three components of linear momentum, the three components of angular momentum, and the center-of-mass motion—which reduce the effective dimensionality but are insufficient for complete integrability. In 1887, proved that no additional algebraic integrals exist beyond these 10 for the n-body problem with n ≥ 3. extended this in 1890 by showing that no new analytic integrals are possible, establishing the non-integrability rigorously. A key simplification is the restricted three-body problem, where one body has negligible mass and does not influence the motion of the other two, assumed to orbit each other circularly in the circular restricted variant. This setup yields the Jacobi integral, a conserved quantity combining energy and angular momentum in the rotating frame. The zero-velocity curves, derived from setting the kinetic energy to zero in the Jacobi integral, delineate forbidden regions of motion and bound the accessible Hill's regions, named after George William Hill's 1878 analysis of lunar perturbations. These regions determine possible trajectories for the negligible-mass body, such as bounded orbits around or unbounded hyperbolic escapes. Despite non-integrability, specific exact solutions exist, including periodic orbits. Joseph-Louis Lagrange identified equilateral triangle configurations in 1772, where the three bodies maintain a rotating equilateral formation as relative equilibria, stable for equal masses under certain conditions. More broadly, collinear and non-collinear periodic orbits have been identified through analytical and numerical methods. A notable example is the figure-eight solution, discovered numerically by Cris Moore in 1993 for equal masses with zero angular momentum, where the bodies trace a symmetric figure-eight path; its existence was rigorously proved in 2000 by and . Recent numerical searches have uncovered thousands more periodic orbits, including 13 new ones in 2013 and over 10,000 three-dimensional solutions reported in 2025. These solutions highlight structured behaviors amid the general chaos. Poincaré's qualitative analysis in his 1892 work introduced groundbreaking insights into the dynamics, revealing homoclinic tangles—intersections of stable and unstable manifolds near periodic orbits—that generate exponential sensitivity to initial conditions and the onset of chaos. This tangled structure precludes simple predictability, influencing modern understanding of nonlinear dynamics in celestial mechanics.

Restricted and Planetary Variants

The restricted three-body problem approximates scenarios in celestial mechanics where one body has negligible mass compared to the other two, treating it as a massless test particle influenced by the gravitational fields of the primaries without affecting their motion. In the circular restricted three-body problem (CRTBP), the two primaries orbit each other in circular paths around their common center of mass, providing a simplified framework for analyzing the test particle's dynamics in a rotating reference frame. This variant, foundational to understanding satellite orbits and asteroid motion, was first systematically explored by in his 1772 essay on the three-body problem. A key feature of the CRTBP is the existence of five equilibrium points, known as L1 through L5, where the gravitational and centrifugal forces balance, allowing the test particle to remain stationary relative to the primaries. The collinear points L1, L2, and L3 lie along the line joining the primaries, while L4 and L5 form equilateral triangles with them, offering stable configurations for long-term orbital placement, as utilized in modern space missions. These points emerge as solutions to the equations of motion in the synodic frame, highlighting the problem's utility in predicting stable relative positions. The dynamics in the CRTBP are constrained by the Jacobi integral, a conserved quantity that combines kinetic energy, gravitational potential, and centrifugal terms: J = (x^2 + y^2) + 2U - v^2, where v is the test particle's speed and U = \frac{1-\mu}{r_1} + \frac{\mu}{r_2} is the gravitational potential due to the primaries (in dimensionless units). This integral, derived from the system's symmetries, delineates allowed regions of motion and zero-velocity surfaces, aiding in the study of bounded orbits and escapes; it was introduced by in his 1842-1843 lectures on dynamics. Planetary variants extend these restricted approximations to hierarchical systems dominated by a central massive body, such as the , with smaller bodies following nearly perturbed by mutual interactions. In the planetary problem, epicyclic approximations model small deviations from circular orbits as harmonic oscillations in radius and azimuth, simplifying the analysis of eccentricities and inclinations under weak perturbations. This approach underpins perturbation theories for multi-planet systems, where the Sun's gravity sets the zeroth-order motion. A seminal application is Hill's lunar theory, developed in 1878, which addresses Earth-Moon-Sun perturbations by treating the Moon's orbit as an epicyclic variation around the , incorporating the Sun's disturbing function to compute long-term inequalities in lunar motion. This method revolutionized lunar ephemerides by avoiding divergent series in traditional expansions. Similar restricted formulations apply to n=4 systems, such as Sun-Jupiter-asteroid configurations, where the asteroid acts as a test particle in the defined by the Sun and Jupiter as primaries, revealing resonant structures that govern asteroid belt stability.

Central Configurations

Central configurations represent a class of equilibrium arrangements in the n-body problem where the gravitational acceleration of each body is directed toward the center of mass and proportional to its position vector relative to that center. This condition implies that the second time derivative of the position vector for each body i satisfies \ddot{\mathbf{r}}_i = \lambda \mathbf{r}_i for some constant \lambda, assuming the center of mass is at the origin without loss of generality. Equivalently, the configuration satisfies the equation \sum_{j \neq i} m_j \frac{\mathbf{r}_j - \mathbf{r}_i}{|\mathbf{r}_j - \mathbf{r}_i|^3} = -\lambda \mathbf{r}_i for each i, where \mathbf{r}_i are the position vectors and m_i the masses. These configurations are significant because they yield self-similar solutions to the equations of motion, known as homographic solutions, in which the entire system expands or contracts uniformly while possibly rotating. They also underpin relative equilibria, where the configuration rotates rigidly about the center of mass at a constant angular velocity, providing insight into stable or periodic orbits in gravitational systems. Additionally, central configurations play a role in analyzing near-collision dynamics, helping to characterize shapes that minimize collision risks in simulations of few-body interactions. For the three-body problem, the known central configurations include Euler's collinear solutions, discovered in 1767, in which the three masses align along a straight line with relative positions determined by the masses to satisfy the proportionality condition. Lagrange identified the equilateral triangular configuration in 1772, where the masses form the vertices of an equilateral triangle regardless of their individual values, representing a non-collinear equilibrium. In the four-body case, a notable central configuration is the regular tetrahedral (pyramidal) arrangement for equal masses, which extends the Lagrange solution into three dimensions and satisfies the central condition through symmetric force balance. For equal masses in the planar four-body problem, numerical studies have identified three distinct central configurations up to symmetry. Extending to five bodies with equal masses, computational enumerations reveal a larger but finite set, with 5 planar configurations, highlighting the increasing complexity as n grows. A key open question in the study of central configurations is Smale's conjecture, posed in 1998 as part of his list of problems for the 21st century, which asserts that for any choice of n positive masses, there exists only a finite number of central configurations up to similarity (scaling, rotation, and translation). This conjecture has been affirmatively resolved for n ≤ 4 through analytic and numerical methods, but remains unresolved for larger n, with numerical evidence supporting finiteness up to n=5.

Analytical Methods

Exact Solvability

The n-body problem for n \geq 3 lacks additional algebraic first integrals beyond the classical ones—energy, linear momentum, and angular momentum—establishing its algebraic non-integrability, as proved by Ernst Heinrich Bruns in 1887 through an analysis of polynomial dependencies in the equations of motion. Henri Poincaré strengthened this result in 1889 by demonstrating that no further uniform transcendental first integrals exist for the three-body problem when one mass is sufficiently small relative to the others, using series expansions to show divergence in potential additional conserved quantities. These proofs imply that closed-form solutions via integration of additional invariants are impossible in general, though the absence applies specifically to algebraic and uniform transcendental forms, leaving open the possibility of non-uniform transcendental integrals that do not converge uniformly. Despite this non-integrability, exact closed-form solutions exist in restricted configurations where the system's symmetries allow explicit expressions for the trajectories. A prominent class comprises homographic solutions, in which the mutual positions of the bodies undergo uniform scaling (homothety) over time while rotating rigidly, preserving the shape of a central configuration; these solutions derive from Keplerian motions around the center of mass and admit analytic parametrization via elliptic functions or simpler forms for specific cases. In zero-angular-momentum scenarios, collinear configurations yield exact rectilinear solutions, such as Euler's homothetic collapses or expansions, where all bodies align along a line and move radially toward or away from the center of mass with velocities proportional to their distances, solvable explicitly as r_i(t) = r_i(0) (1 + \lambda t)^{2/3} for scaling parameter \lambda. For the three-body problem, relative equilibrium solutions—rigidly rotating central configurations—provide exact closed forms, with five distinct families: three collinear Euler configurations (differing by which body occupies the central position) and two equilateral triangular Lagrange configurations (distinguished by orientation). These generalize to higher n in zero-angular-momentum cases through homothetic expansions of arbitrary central configurations, where the entire system scales uniformly without rotation, yielding explicit solutions of the form \mathbf{q}_i(t) = \alpha(t) \mathbf{q}_i(0) with \alpha(t) satisfying a quadratic differential equation derived from the energy constraint. Such solutions highlight the problem's solvability in symmetric, low-dimensional subspaces despite its general intractability.

Series and Perturbation Solutions

Series solutions for the n-body problem, particularly the three-body case, provide approximate analytic expressions through infinite power series expansions that converge under specific conditions. In 1912, Karl Fritiof Sundman developed a global power series solution for the general three-body problem, expressing positions and velocities as series in fractional powers of time, such as r(t) = \sum_{k=0}^{\infty} a_k (t - t_0)^{k/3}, where the exponent $1/3 arises from regularization to handle the singularity at triple collisions. This series converges for all finite times except at collision points, offering a formal proof of the existence of analytic solutions despite the lack of closed-form expressions. Perturbation theory addresses the n-body problem by treating small interactions as corrections to integrable subsystems, such as the two-body Keplerian orbits in planetary systems. Secular perturbations, which describe long-term changes in orbital elements like eccentricity and inclination, are derived by averaging the disturbing function over fast orbital periods and expanding it in Legendre polynomials P_l(\cos \psi), where \psi is the angular separation between bodies. This expansion facilitates the computation of resonant and non-resonant effects in multi-planet configurations, enabling predictions of orbital stability over astronomical timescales. For non-resonant planetary perturbations, Édouard von Zeipel's method (1910) constructs a canonical transformation to eliminate short-period terms, yielding a secular Hamiltonian that isolates long-term dynamics. In cases involving close encounters, Levi-Civita regularization (1903) transforms the equations of motion in the plane, replacing the singular $1/r potential with a regular one by introducing a complex coordinate w = u + iv such that r = w^2, thereby extending series solutions through near-collision regimes. This approach improves convergence for binary interactions within larger n-body systems. Convergence of these series is limited by singularities corresponding to collisions or escapes, but analytic continuation techniques allow extension beyond the initial radius of convergence by re-expanding around branch points in the complex time plane. Such methods ensure the series remains valid for practical computations in non-collisional regimes, though numerical verification is often required near potential singularities.

Singularities and Breakdowns

In the n-body problem, singularities arise when the configuration of bodies leads to undefined accelerations due to the inverse-square gravitational force, typically when distances between bodies approach zero. Binary collisions represent a primary type of singularity, occurring when two bodies i and j satisfy | \mathbf{r}_i - \mathbf{r}_j | \to 0 in finite time, causing infinite mutual acceleration while the overall configuration remains bounded. These events are algebraic branch points in the complex plane and can be analyzed as elastic bounces upon regularization. Total collapses constitute another class, where all n bodies converge to a single point in finite time, resulting in a global singularity of the system; such collapses require zero total angular momentum and are rarer than binary collisions. In 1897, Paul Painlevé proved that for the three-body problem, all singularities are collision singularities, which can be removed through suitable variable changes without introducing non-removable branch points or essential singularities. He conjectured that for n ≥ 4, the n-body problem admits solutions with non-collision singularities, indicating topological instability. This conjecture was proved for n ≥ 5 in 1990 by C. Marchal, for the planar n=4 case in 2014 by A. Albouy et al., and for the general n=4 case in 2021 by J. Xue. In configurations with zero angular momentum, collisions can become inevitable under specific energy conditions, as demonstrated by Zhihong Xia in 1992, who proved that certain zero-angular-momentum n-body systems lead to unavoidable collisions through repeated ejection-collision mechanisms, where bodies are successively expelled to large distances before returning to collide. These mechanisms underscore the role of angular momentum in preventing singularities, as nonzero values generally avoid total collapses while permitting near-collisions without actual impact. To extend solutions beyond these singularities, regularization techniques transform the singular equations into regular ones. The Kustaanheimo-Stiefel (KS) transformation achieves this by embedding the three-dimensional position vectors into a four-dimensional quaternionic space, converting binary collision singularities into nonsingular harmonic oscillator dynamics and allowing analytic continuation past the breakdown points. This method, originally developed for the , effectively handles isolated binary collisions in the general n-body setting by rescaling time and coordinates proportionally to the inverse square root of the moment of inertia.

Numerical Approaches

Integration Techniques

Numerical integration of the n-body problem involves solving the system of ordinary differential equations (ODEs) derived from and gravitation. Direct integration methods, such as , are commonly employed for their simplicity and high local accuracy in approximating solutions to these ODEs. However, these explicit are not symplectic, leading to secular energy drift and poor long-term stability in Hamiltonian systems like the gravitational n-body problem, where phase-space volume preservation is crucial. To mitigate these issues, symplectic integrators are preferred, as they preserve the symplectic structure of the Hamiltonian formulation, ensuring bounded energy errors over extended simulations. The Verlet algorithm and its variant, the leapfrog integrator, are foundational symplectic methods widely used in n-body simulations for their time-reversibility and conservation of energy and momentum. In the leapfrog scheme, positions and velocities are updated in a staggered manner to maintain second-order accuracy with minimal computational cost; for instance, the half-step position update is given by \mathbf{r}_{n+1/2} = \mathbf{r}_n + \frac{h}{2} \mathbf{v}_n, followed by a full-step velocity update using accelerations at the half-steps, and then the next half-step position. This approach ensures excellent long-term behavior in conservative systems without artificial dissipation. For few-body problems involving close encounters, where singularities in the equations can cause numerical instability, regularized integrators transform the coordinates and time to eliminate these divergences, allowing stable integration through near-collisions. Democratic regularization, a global approach that treats all pairwise interactions symmetrically without favoring a central body, is particularly effective for handling multiple close encounters in few-body dynamics by reparametrizing the Hamiltonian to bound the step sizes. A notable advancement in symplectic integration for planetary systems is the Wisdom-Holman mapping, introduced in 1991, which exploits the hierarchical structure of such systems by separating fast Keplerian motions around a dominant central mass from slower perturbations. This kick-drift-kick formulation enables efficient long-term simulations with superior energy conservation compared to non-symplectic methods.

Few-Body Simulations

Few-body simulations address the numerical integration of the n-body equations of motion for small systems, typically with fewer than 10 interacting bodies, where high precision is paramount due to the absence of statistical approximations and the potential for chaotic behavior. These simulations prioritize adaptive algorithms that maintain accuracy over long timescales, particularly in scenarios involving close gravitational encounters that can lead to numerical instabilities. Unlike larger-scale methods, they leverage direct computation of all pairwise interactions without softening or cutoffs, enabling detailed analysis of individual trajectories and energy conservation. The Bulirsch-Stoer extrapolation method exemplifies a high-accuracy approach suited for three-body scattering problems, where it employs variable-step integration combined with polynomial extrapolation to achieve machine precision in solutions. This predictor-corrector technique minimizes truncation errors by iteratively refining estimates from smaller substeps, proving effective for resolving the sensitive dynamics of unbound encounters in few-body systems. In applications to three-body interactions, it ensures converged results up to many significant digits, crucial for studying outcomes like binary formation or ejections. Event-driven methods enhance precision by detecting and resolving critical events such as close approaches or ejections, which pose risks of singularity in standard integrators. These techniques monitor inter-body distances during propagation, triggering specialized subroutines—often regularization transformations—to handle near-collisions without excessive time-step reduction. In few-body contexts, such as three-body disintegrations, they identify ejections when a body's energy exceeds the system's binding energy, allowing seamless continuation of the simulation while conserving total energy and angular momentum to high order. Few-body simulations play a key role in generating solar system ephemerides, as exemplified by JPL's DE430 model, which integrates the full n-body dynamics of the Sun, planets, Moon, and over 300 asteroids but emphasizes precise treatment of the dominant planetary interactions. This approach fits observed data to numerically integrated orbits, achieving sub-kilometer accuracy for inner planets over millennia, by focusing computational effort on the few primary gravitational perturbations rather than uniform treatment of all bodies. Visualization techniques like provide insights into chaotic structures in three-body simulations, reducing the continuous phase space to discrete maps at fixed energy surfaces or plane crossings. These sections reveal invariant tori for regular motion and tangled structures for chaos, aiding qualitative analysis of stability and homoclinic tangles in restricted three-body problems. In numerical studies, they highlight the onset of irregular behavior near resonances, complementing quantitative integrators like .

Many-Body and Astrophysical Applications

In astrophysical simulations of the n-body problem, direct summation of gravitational forces becomes computationally prohibitive for large particle numbers n > 10^3, necessitating approximate methods that trade precision for scalability to model cosmic , clustering, and evolution. These techniques enable the study of gravitational dynamics over vast scales, from individual to the large-scale , by exploiting spatial hierarchies or grid-based decompositions to reduce the complexity from O(N^2) to more efficient forms. Tree codes, such as the Barnes-Hut algorithm introduced in , achieve O(N \log N) scaling through hierarchical spatial partitioning, where the simulation volume is recursively subdivided into an or similar structure, allowing distant particle groups to be approximated as single multipole expansions for force calculations on nearby particles. This method balances accuracy and efficiency by using a criterion, often the opening angle \theta, to decide when to approximate clusters, making it suitable for collisionless systems like star clusters and galaxy mergers. Particle-mesh methods complement tree codes in cosmological applications by leveraging fast Fourier transforms (FFT) to solve Poisson's equation on a uniform grid with periodic boundary conditions, assigning particles to mesh points via cloud-in-cell interpolation before computing long-range forces in Fourier space at O(N_g \log N_g) cost, where N_g is the grid size typically comparable to N. These are particularly effective for uniform-density backgrounds in expanding universes, though they sacrifice short-range resolution, often requiring hybridization with direct or tree summations for close encounters. A landmark application is the Millennium Simulation of 2005, which evolved $10^{10} dark matter particles in a comoving volume of (500 \, \mathrm{Mpc}/h)^3 using a hybrid tree-particle-mesh approach in the GADGET code, resolving halo formation and large-scale structure in a \LambdaCDM cosmology over cosmic time from z=127 to the present. Recent advances as of 2024 include GPU-accelerated hybrid hydro/N-body codes like Enzo-N for simulating star clusters within galaxies and machine learning-assisted frameworks such as COmoving Computer Acceleration (COCA) for efficient N-body evolution in emulated reference frames. Key challenges in these large-n simulations include two-body relaxation, where cumulative small-angle deflections alter particle orbits on a timescale t_{\rm relax} \sim N / \log N times the dynamical time, potentially introducing artificial diffusion unless N is sufficiently large to exceed the Hubble time. Strong encounters are mitigated using softened potentials, such as the Plummer form \Phi(r) = -GM / \sqrt{r^2 + \epsilon^2}, where the softening length \epsilon suppresses unphysical force divergences and shot-noise fluctuations while preserving large-scale dynamics.

Extensions

Non-Gravitational Systems

The n-body problem generalizes to non-gravitational systems where particles interact via central forces other than the inverse-square gravitational law, such as electrostatic or intermolecular potentials, leading to diverse applications in plasma physics, atomic structure, and molecular simulations. These extensions retain the core challenge of predicting individual motions under mutual interactions but introduce variations in force signs, ranges, and forms that can enable repulsive dynamics or short-range effects absent in purely gravitational cases. Unlike the gravitational formulation, which assumes universal attraction proportional to masses, non-gravitational interactions depend on particle properties like charges or sizes, often resulting in integrable or screened behaviors for specific configurations. A prominent example is the Coulomb n-body problem, which governs the classical dynamics of charged particles such as electrons and nuclei in atoms or ions in plasmas. The equations of motion are given by m_i \frac{d^2 \mathbf{r}_i}{dt^2} = \sum_{j \neq i} \frac{q_i q_j (\mathbf{r}_j - \mathbf{r}_i)}{|\mathbf{r}_j - \mathbf{r}_i|^3}, where m_i and q_i are the mass and charge of the i-th particle, and \mathbf{r}_i its position vector; this form mirrors the gravitational case but allows both attractive (opposite charges) and repulsive (like charges) forces. In atomic physics, the Born-Oppenheimer approximation separates electronic and nuclear motions by treating the lighter electrons quantum mechanically while reducing the heavier nuclei to a classical n-body problem on an effective potential surface derived from the electronic energy, enabling simulations of molecular vibrations and rotations. This classical nuclear treatment simplifies the full quantum many-body challenge while capturing key dynamical features like bond stretching in diatomic molecules. In plasma physics, the Coulomb n-body framework models collisionless dynamics of ionized gases, but real systems exhibit screening effects that alter the long-range inverse-square behavior. Debye screening, arising from the collective response of mobile charges, effectively shortens the interaction range by introducing a cloud of opposite charges around each particle, modifying the bare Coulomb potential to a screened Yukawa form V(r) \propto \frac{e^{-r/\lambda_D}}{r}, where \lambda_D is the Debye length depending on plasma temperature and density; this prevents divergences in many-body sums and stabilizes simulations of phenomena like wave propagation. Such screening distinguishes plasma n-body problems from unscreened gravitational ones, as it introduces a natural cutoff that enhances computational tractability for large particle numbers. Another key extension appears in molecular dynamics simulations, where the models non-bonded interactions between atoms or molecules, particularly van der Waals forces in liquids and solids. This pairwise potential combines a steep repulsive term for atomic overlaps with a weaker attractive dispersion, expressed as V(r) = 4\epsilon \left[ \left( \frac{\sigma}{r} \right)^{12} - \left( \frac{\sigma}{r} \right)^6 \right], with \epsilon setting the interaction strength and \sigma the finite atomic size; the r^{-12} repulsion provides a soft-core barrier mimicking Pauli exclusion without quantum details. Widely used in n-body simulations of biomolecular systems or materials, it facilitates studies of phase transitions and diffusion, though its empirical parameters must be tuned for accuracy in specific substances. These non-gravitational formulations highlight how force-law modifications can yield richer dynamical regimes, from stable equilibria in screened plasmas to fluid-like behaviors in Lennard-Jones gases.

Choreographies and Periodic Orbits

In the n-body problem, choreographies represent a class of symmetric periodic solutions where all bodies, assumed to have equal , trace out the same closed curve in the plane, positioned at equal angular intervals and phase-shifted by \frac{2\pi}{n} along the orbit. These solutions generalize relative equilibrium configurations, such as Lagrange's for n=3, by allowing more complex, non-circular paths while maintaining the choreographic symmetry. The existence of such orbits relies on the of the center of mass and the rotational invariance of the , ensuring the bodies "chase" each other without collision. A seminal example is the figure-eight for the , first identified numerically in 1993 and rigorously proven to exist in 2000 using variational methods on the reduced action functional. In this orbit, the three bodies follow a symmetric path with period T, starting from a central and evolving under mutual gravitational attraction. For n=3, numerical explorations have uncovered 345 distinct planar choreographies, many exhibiting intricate loops and symmetries beyond the figure-eight. These include families with varying topological complexities, such as those with multiple windings around the center of mass. For higher n, thousands of choreographies have been discovered through numerical continuation from lower-dimensional cases, revealing a rich variety of periodic orbits up to n in the hundreds. In the early , Carles Simó systematically computed such solutions for n up to 20 and beyond, using techniques that branch from known central configurations as initial guesses. These findings have implications for approximations in Hill's lunar problem, where choreographic symmetries inform analyses of restricted dynamics near the Earth-Moon system. Computational discovery of choreographies typically employs variational principles, minimizing the Maupertuis subject to fixed and symmetry constraints, often combined with methods to refine initial conditions. Numerical continuation tracks families of solutions as parameters like or n vary, leveraging the of the configuration to generate new orbits from bifurcations. Central configurations serve as natural starting points for these searches, providing collision-free equilibria that evolve into full periodic motions under time-dependent perturbations.

References

  1. [1]
    [PDF] The N-body problem - Ceremade
    Summary. We introduce the N-body problem of mathematical celestial mechanics, and dis- cuss its astronomical relevance, its simplest solutions inherited ...
  2. [2]
    [PDF] On the N-body Problem - BYU ScholarsArchive
    These equations are called the N-body problem of celestial mechanics. There are many problems associated with the dynamics of such a system, for example, exis- ...
  3. [3]
    [PDF] TOPICS IN CELESTIAL MECHANICS 1. The Newtonian n-body ...
    Celestial mechanics can be defined as the study of the solution of Newton's ... For example, consider the planar n-body problem (d = 2, n ≥ 2). The ...
  4. [4]
    [PDF] N-BODY PROBLEMS Math118, O. Knill
    We could for example look at the natural n body problem on a torus or the sphere. • One would need (6n − 1) integrals of motion to solve the n-body problem ...
  5. [5]
    [PDF] N-body Problem in - DSpace@MIT
    Feb 5, 2018 · N-body problem. The N-body ODE comes from a combination of Newton's second law, which relates the change in momentum to a force dependent on ...
  6. [6]
    [PDF] arXiv:2407.02358v3 [physics.class-ph] 27 May 2025
    May 27, 2025 · The general N-body problem has long been a challenge in analytical mechanics. Newto- nian gravity has been of particular interest, where the ...
  7. [7]
    [PDF] Four Open Questions for the N-Body Problem Richard Montgomery
    We've been working on the N-body problem for more than 330 years. You might think we'd be done with it. But the problem remains very much alive. Substantial.<|control11|><|separator|>
  8. [8]
    [PDF] N-Body Methods
    This yields the standard Newtonian equations of motion for N point masses. ... the singular potential wells associated with point masses create awkward numerical ...
  9. [9]
    Ancient Greek Astronomy and Cosmology | Modeling the Cosmos
    Ptolemy and Aristotle's Cosmic Legacy. Ptolemy came to represent a mathematical tradition, one focused on developing mathematical models with predictive power.
  10. [10]
    Kepler's Laws - Planetary Orbits - NAAP - UNL Astronomy
    Johannes Kepler published three laws of planetary motion, the first two in 1609 and the third in 1619. The laws were made possible by planetary data of ...Missing: empirical basis body multi- complexities
  11. [11]
    [PDF] Kepler's Laws of Planetary Motion: 1609-1666 JL Russell
    Feb 12, 2008 · Kepler's laws of planetary motion were largely ignored between the time of their first publication (1609, 1619) and the publication of Newton's.
  12. [12]
    Newton's Philosophiae Naturalis Principia Mathematica
    Dec 20, 2007 · The view is commonplace that what Newton did was to put forward his theory of gravity to explain Kepler's already established “laws” of orbital ...
  13. [13]
    Newton, Principia, 1687 - Hanover College History Department
    Isaac Newton is probably most famous for having discovered the universal laws of gravity. (That is, he showed that gravity explains the behavior of stars and ...Missing: qualitative | Show results with:qualitative
  14. [14]
    How a falling apple could have helped Newton discover universal ...
    Jan 12, 2024 · The article delves into the intriguing disagreement among historians of science regarding the origin of the idea of universal gravity.Missing: analogy | Show results with:analogy
  15. [15]
    Three body problem - Scholarpedia
    Oct 3, 2007 · The problem is to determine the possible motions of three point masses m_1\ , m_2\ , and m_3\ , which attract each other according to Newton's law of inverse ...
  16. [16]
    [PDF] Euler's Three Body Problem - OSU Math
    Euler discussed it in memoirs published in 1760. The problem is analytically solvable but requires the evaluation of elliptic integrals. Numerical methods may.
  17. [17]
    Solar and planetary destabilization of the Earth–Moon triangular ...
    However, Lagrange (1772) proved that there are five stationary positions in the restricted circular three-body problem. At these stationary points, a ...
  18. [18]
    [PDF] Poincaré and the Three-Body Problem
    The Three-Body Problem has been a recurrent theme of Poincaré's thought. Having understood very early the need for a qualitative study of “non- integrable” ...
  19. [19]
  20. [20]
    The dramatic episode of Sundman - ScienceDirect.com
    In 1912 the Finnish mathematical astronomer Karl Sundman published a remarkable solution to the three-body problem.
  21. [21]
    [PDF] karl fritiof sundman – the man who solved the three-body problem
    Karl Fritiof Sundman, a Finnish mathematician-astronomer, solved the three-body problem, which was thought to be unsolvable by Henri Poincaré.
  22. [22]
    Kolmogorov-Arnold-Moser theory - Scholarpedia
    Sep 23, 2010 · Kolmogorov-Arnold-Moser (KAM) theory deals with persistence, under perturbation, of quasi-periodic motions in Hamiltonian dynamical systems.Classical KAM theory · Small divisors and classical... · Applications and extensions
  23. [23]
    [PDF] An Introduction to KAM Theory
    Jan 22, 2008 · Remark 3.1 The three-body (or N-body) problem, in which we ignore the mu- tual interaction between the planets is an integrable system. Now ...
  24. [24]
    [PDF] The Two-Body Problem - UCSB Physics
    Thus, the angular momentum we previously defined is nothing other than the full two-body angular momentum, evaluated in the center-of-mass frame (re- member ...
  25. [25]
    [PDF] 8.01SC S22 Chapter 25: Celestial Mechanics - MIT OpenCourseWare
    Jun 25, 2013 · The parameter r in the one-body problem is the distance between the reduced mass and the central point, and also the relative distance between ...
  26. [26]
    [PDF] Kepler's Laws for the 2-Body Problem - Robert Vanderbei
    ABSTRACT. Kepler's three laws of planetary motion describe the dynamics of the 2-body problem where one body is the Sun and the other body is a planet.Missing: 1609-1619 empirical basis multi- complexities
  27. [27]
    Bruns' Theorem: The Proof and Some Generalizations - ResearchGate
    Aug 9, 2025 · There is a long history around the existence/non-existence of first integrals for the three-body problem as well as for general Hamiltonian ...
  28. [28]
    [PDF] Non-integrability of the n-body problem - arXiv
    May 24, 2025 · H. Bruns [4] showed the non-existence of additional algebraic first integrals, except energy, linear, and angular momentum. His proof was ...
  29. [29]
    [PDF] exploring the influence of a three-body interaction
    George. Hill then used this knowledge in 1878 to prove the existence of the zero velocity curves that bound regions of allowable motion for the small, third ...<|control11|><|separator|>
  30. [30]
    Zero velocity curves and orbits in the restricted problem of three bodies
    A GENERAL CRITERION IN this section we derive the condition for having zero velocity curves as orbits in the restricted problem of three bodies.Missing: source | Show results with:source
  31. [31]
    [PDF] On the Homoclinic Tangles of Henri Poincaré - Arizona Math
    In a letter addressed to Mittag-Leffler, Poincaré claimed to have proven a stability result for the restricted three-body problem. He wrote ([14], page. 44) ...
  32. [32]
    [PDF] Essai sur le Problème des trois Corps
    Essai sur le Problème des trois Corps by: Joseph Louis Lagrange, Joseph Alfred Serret in: Oeuvres de Lagrange, volume: Tome 6 pp. 229 - 332. Terms and ...
  33. [33]
    C. G. J. Jacobi's Vorlesungen über Dynamik - Internet Archive
    Jun 9, 2008 · Jacobi's Vorlesungen über Dynamik: Gehalten an der Universität zu Königsberg im ... ... PDF download · download 1 file · SINGLE PAGE PROCESSED ...
  34. [34]
    Researches in the Lunar Theory - jstor
    G. W. Hill, Researches in the Lunar Theory, American Journal of Mathematics, Vol. 1, No. 2 (1878), pp. 129-147.
  35. [35]
    Resonant Motion in the Restricted Three Body Problem - ADS
    The resonant structure of the restricted three body problem for the Sun- Jupiter asteroid system in the plane is studied, both for a circular and an elliptic ...
  36. [36]
    [PDF] Lectures on Central Configurations
    This approach was initiated by Smale [44] and developed by Pal- more [36] for the planar n-body problem and extended to three dimensions using ...
  37. [37]
    [PDF] Central Configurations—A Problem for the Twenty-first Century
    More precisely for each way the particles can be ordered along a line, there is a unique position that causes a central configuration.
  38. [38]
    [PDF] The symmetric central configurations of four equal masses - IMCCE
    for some real numbers γ and ν. 4. The case of a configuration of four equal masses. We know that such a configuration, if central, possesses an axis of symmetry.
  39. [39]
    ON A RESULT OF BRUNS
    Bruns' Theorem states that the classical integrals of the gravitational three-body problem generate all algebraic integrals. We show that the first step in his ...<|separator|>
  40. [40]
    [PDF] the solution of the n-body problem
    covered in Brun's proof, Poincaré had no doubt that the result was true. In his prized paper he proved an even stronger theorem: there are no other integrals.
  41. [41]
    [PDF] The zero angular momentum, three-body problem - UC Santa Cruz
    At zero angular momentum, these solutions evolve by homothety (scaling). The Euler solutions are collinear at each instant. The Lagrange homothety solution [2] ...Missing: exact | Show results with:exact
  42. [42]
    [PDF] Bifurcations of relative equilibria of the (1+3)-body problem
    Abstract. We study the relative equilibria of the limit case of the pla- nar Newtonian 4–body problem when three masses tend to zero, the so-called (1 + 3)– ...<|separator|>
  43. [43]
    [PDF] The dramatic episode of Sundman - Open Research Online
    Bruns, H. 1887. Über die Integrale des Vielkörper-Problems. Acta Mathematica, 11, 25–. 96. Buchanan, H.E. 1937. Some recent results in the problem of three ...
  44. [44]
    MPP01, a new solution for planetary perturbations in the orbital ...
    The Legendre polynomials are expanded with an anal- ogous approach to Brown's. ... One makes a separation of monomials depending on lunar coordinates solely from.
  45. [45]
    SECULAR EVOLUTION OF HIERARCHICAL PLANETARY SYSTEMS
    Pk is the Legendre polynomial of degree k, and Φ is the angle between r1 and r2. The first two terms in equations (10) and. (11) represent the independent ...Missing: source | Show results with:source
  46. [46]
    The Von Zeipel–Lidov–Kozai Effect inside Mean Motion ...
    Aug 2, 2022 · According to the independent works performed by von Zeipel (1910), Lidov (1962), and Kozai (1962), the coupled oscillations between eccentricity ...
  47. [47]
    A regularization of multiple encounters in gravitational n-body ...
    The regularization of two-body encounters in two dimensions was introduced by LeviCivita (1903). The Levi-Civita regularization was generalized to three ...
  48. [48]
    [PDF] On the real singularities of the N-body problem - SciSpace
    This bibliography is intended to be a complete listing of papers on the title subject; the major part of the papers deals with the classical N-body problem ( ...
  49. [49]
    [PDF] Understanding the Dynamics of Collision and Near-Collision ... - arXiv
    Aug 30, 2012 · By Theorem 8, a solution of the 2-Body Problem with nonzero angular momentum does not experience a collision or total collapse. A nonzero.Missing: Zhihong | Show results with:Zhihong
  50. [50]
    The Existence of Noncollision Singularities in Newtonian Systems
    This difference in rotation permits the total angular momentum of the system to be zero and also permits the two binaries to have arbitrarily small but nonzero ...
  51. [51]
    Some insights from total collapse in the N-body problem
    It was Hooke's question on the two-body problem and Halley's encouragement (and financial resources) that eventually led Newton to publish the Principia.2 ...
  52. [52]
    [PDF] ON PAINLEVÉ CONJECTURE - Harvard Math
    Jan 4, 2021 · The conjecture asserts the topological instability of the N-body problem and seems extremely hard. Moreover, Herman in [21] also gives a ...
  53. [53]
    [PDF] REGULARIZATION IN CELESTIAL MECHANICS
    Regularization in celestial mechanics introduces new variables to remove singularities in equations of motion during binary collisions, making them regular.
  54. [54]
    The Lissajous–Kustaanheimo–Stiefel transformation
    Feb 15, 2019 · Whenever the angular momentum vanishes (even temporarily), the angles become undetermined and equations of motion are singular. It turns out ...
  55. [55]
    Efficient Numerical Methods for N-Body Simulations with Modern ...
    Dec 21, 2024 · Runge-Kutta methods and symplectic integrators offer high accuracy in solving differential equations governing N -body dynamics. Runge-Kutta ...
  56. [56]
    [PDF] Should N-body integrators be symplectic everywhere in phase space?
    Mar 26, 2019 · Symplectic integrators are the preferred method of solving conservative N-body problems in cosmological, stellar cluster, and planetary ...
  57. [57]
    Symplectic variable step size integration for N-body problems
    Multiple time stepping can be applied to the leapfrog/Störmer/Verlet integrator so as to effect a variable step size algorithm. The strategy maintains the ...
  58. [58]
    [PDF] REGULARIZATION AND NUMERICAL METHODS IN CELESTIAL ...
    Dec 3, 2023 · It is known that the regularizing technique is very useful for n-body simulations to handle close encounters. Regularized equations of motion ...
  59. [59]
  60. [60]
    On the reliability of N-body simulations
    Mar 28, 2015 · The gravitational N-body problem aims to solve Newton's equations of motion under gravity for N stars (Newton 1687). A popular integrator to ...<|separator|>
  61. [61]
    An implementation of N-body chain regularization
    AN IMPLEMENTATION OF N-BODY CHAIN REGULARIZATION SEPPO MIKKOLA Turku ... 2.1 THE N-BODY HAMILTONIAN Consider a perturbed N-body system with inertial ...Missing: democratic | Show results with:democratic
  62. [62]
    [PDF] The Planetary and Lunar Ephemerides DE430 and DE431 - NASA
    Feb 15, 2014 · The asteroid orbits were iteratively integrated with the positions of the planets, the Sun, and the Moon. The set of 343 asteroids is identical ...
  63. [63]
    [PDF] Poincaré Sections and Resonant Orbits in the Restricted Three ...
    2.1 Definitions in the n-body Problem . ... position vectors requires 18 integrals, but only 10 are available. In the ...
  64. [64]
    The cosmological simulation code gadget-2 - Oxford Academic
    We discuss the cosmological simulation code gadget-2, a new massively parallel TreeSPH code, capable of following a collisionless fluid with the N-body method.
  65. [65]
    A hierarchical O(N log N) force-calculation algorithm - Nature
    Dec 4, 1986 · Barnes, J., Hut, P. A hierarchical O(N log N) force-calculation algorithm. Nature 324, 446–449 (1986). https://doi.org/10.1038/324446a0.
  66. [66]
  67. [67]
    Towards optimal softening in three-dimensional N-body codes
    In N-body simulations of collisionless stellar systems, the forces are softened to reduce the large fluctuations owing to shot noise.
  68. [68]
    Optimal softening for force calculations in collisionless N-body ...
    In N-body simulations the force calculated between particles representing a given mass distribution is usually softened, to diminish the effect of graininess.Introduction · Methods · Optimal smoothing for a... · The case of two Plummer...
  69. [69]
    N-body choreographies - Scholarpedia
    Oct 30, 2013 · Impetus for their study was born out of the Chenciner and Montgomery, 2000 rediscovery and existence proof of the (Moore 1994) figure eight ...
  70. [70]
    [PDF] New Families of Solutions in N-Body Problems
    I am based on an extensive numerical search with zero angular momentum and ˙z1 = ˙z2. • The eight lives on c = 0, the zero angular momentum level. When c = 0 it.Missing: exact | Show results with:exact<|separator|>