Fact-checked by Grok 2 weeks ago

Orbit

An orbit is the gravitationally curved of an object, such as a , , or artificial , around a more massive central body, resulting from the balance between gravitational attraction and the object's inertial motion. This path is typically regular and repeating for bound systems, with the orbiting body referred to as a of the central mass. Orbits are fundamental to , governing the motion of bodies in the solar system and enabling and technology. The mathematical description of orbits originated with Johannes Kepler's three empirical laws of planetary motion, derived from precise observations by Tycho Brahe in the late 16th and early 17th centuries, which established that planets follow elliptical paths with the Sun at one focus, sweep equal areas in equal times, and have periods squared proportional to the semi-major axis cubed. Isaac Newton unified these laws in 1687 through his law of universal gravitation, which posits that every mass attracts every other with a force proportional to the product of their masses and inversely proportional to the square of the distance between them, explaining why objects "fall" into curved paths around massive bodies. This framework predicts that orbits are conic sections: elliptical (including circular) for bound, closed paths with total energy less than zero; parabolic for marginally unbound trajectories with zero energy; and hyperbolic for unbound escapes with positive energy, characterized by eccentricity values of less than 1, exactly 1, and greater than 1, respectively. In contemporary astronomy and engineering, orbits encompass natural phenomena like Earth's 365.25-day elliptical orbit around the Sun and artificial ones, such as low orbits (160–2,000 km altitude) for imaging satellites, medium orbits (2,000–35,500 km) for navigation systems like GPS, and geostationary orbits (about 35,786 km) where satellites appear fixed over a point on equator. Polar and Sun-synchronous orbits, often low types, allow global coverage and consistent lighting for . These applications rely on precise —semi-major axis, , inclination, and others—to predict and control trajectories, with perturbations from atmospheric , solar radiation, and other bodies requiring ongoing adjustments.

Fundamentals

Definition and Basic Concepts

An orbit is the curved followed by an object in space as it moves under the influence of gravitational attraction from a more massive central body, such as a or . These paths are typically described as conic sections, including ellipses for closed orbits, parabolas for marginally escaping trajectories, and hyperbolas for paths that do not return to the central body. In this geometric framework, the central body occupies one of the foci of the conic section, defining the orbit's shape relative to the attracting mass. Key geometric features include the periapsis, the point of closest approach to the central body, and the apoapsis, the point of greatest distance. These distances vary depending on the orbit's , with circular orbits having equal periapsis and apoapsis radii. Orbits are classified qualitatively as bound or unbound: bound orbits, such as ellipses, keep the object perpetually near the central body due to insufficient energy for , while unbound orbits, like parabolas and hyperbolas, allow the object to depart indefinitely after a close approach. Common examples of orbits include natural satellites like Earth's , which follows an elliptical path around our planet, and artificial satellites in used for communication and observation. Maintaining such an orbit requires orbital velocity, the tangential speed at which the object's centripetal acceleration balances the gravitational pull, preventing either a crash into the central body or escape to infinity. This velocity arises from Newtonian gravity as the fundamental force governing the motion.

Types of Orbits

Orbits in celestial mechanics are classified primarily by their shape, which follows the geometry of conic sections, determined by the eccentricity e of the trajectory relative to a central gravitational body. Elliptical orbits, with $0 \leq e < 1, represent bound, closed paths where the orbiting body repeatedly returns to the same points, maintaining a stable configuration under gravitational influence; these are the most common for natural satellites and planets. A circular orbit is a special case of an elliptical orbit where e = 0, resulting in a constant radius and uniform speed, simplifying mission planning for Earth-orbiting satellites like those in low Earth orbit. Parabolic orbits, with e = 1, describe marginally unbound trajectories where the body approaches from infinity, reaches a closest point, and escapes back to infinity with zero velocity at infinite distance, often idealized for escape scenarios. Hyperbolic orbits, with e > 1, are unbound and open, characteristic of high-speed flybys where the body enters from infinity, swings around the central body, and departs without returning, as seen in some comet paths or interplanetary probes. Practical applications extend these classifications into specialized types tailored to mission needs. The , an elliptical path with the initial and final circular orbits as its apsides, enables fuel-efficient transitions between concentric orbits by requiring only two impulsive burns, widely used for planetary missions such as those to Mars. Geostationary orbits exemplify a circular, equatorial configuration at approximately 35,786 km altitude, where the matches Earth's rotation (about 23 hours 56 minutes), allowing satellites to remain fixed over a single point on the equator for continuous communication or weather monitoring. Orbits are further distinguished by their orientation relative to the primary body's and equatorial . Prograde orbits proceed in the same as the primary's , aligning with the natural for , while retrograde orbits move in the opposite , often resulting from captures or collisions and requiring more delta-v for insertion. Equatorial orbits have an inclination of 0° (or 180° for retrograde), lying in the plane of the , whereas polar orbits feature a 90° inclination, passing over the poles to provide global coverage. Orbit inclination plays a key role in : low-inclination equatorial orbits suit geostationary applications for regional focus, while higher inclinations, especially near-polar, enable sun-synchronous passes for consistent lighting in , allowing comprehensive scanning of latitudes without gaps. For instance, polar orbits facilitate full planetary imaging as the satellite's shifts daily due to .

Historical Development

Pre-Modern Observations

Early civilizations, particularly the Babylonians around 2000–500 BCE, maintained meticulous records of celestial phenomena from a geocentric perspective, viewing Earth as the fixed center of the universe. Babylonian astronomers documented the positions of the Sun, Moon, and planets, noting patterns such as the irregular wandering paths of planets against the fixed stars and the occurrence of retrogrades, where planets appeared to loop backward in the sky. These observations, inscribed on clay tablets, emphasized empirical cycles like the 18-year saros cycle for lunar eclipses and planetary periods, which were used for calendrical and astrological predictions rather than theoretical explanations. In , (384–322 BCE) developed a philosophical framework for these observations, positing a geocentric composed of concentric made of , a distinct from earthly matter. He argued that natural motion in the celestial realm was eternal and circular, with planets and stars embedded in these spheres rotating uniformly around due to their inherent , contrasting with the linear fall of sublunary bodies toward the center. This model integrated earlier Babylonian data but prioritized qualitative harmony over quantitative precision, influencing subsequent geocentric views. The Ptolemaic system, formalized by Claudius Ptolemy in his around 150 CE, refined geocentric models to better fit accumulated observations of planetary irregularities. Ptolemy introduced deferents—large circles centered near —upon which moved via smaller epicycles, allowing the model to replicate observed retrogrades and varying speeds; he further employed equants, points offset from the deferent's center, to ensure uniform angular motion as seen from . This geometric construct successfully predicted planetary positions, including Mars' prominent retrogrades and the Moon's anomalous cycles, drawing on centuries of and Babylonian data for validation. Medieval Islamic astronomers built upon Ptolemaic foundations through extensive observations, enhancing accuracy in tracking celestial patterns. (c. 858–929 CE) conducted over 40 years of precise measurements from , , refining the length of the solar year to 365 days, 5 hours, 46 minutes, 24 seconds, and cataloging fixed stars while noting planetary retrogrades and lunar phases with improved trigonometric methods. Later, (1201–1274 CE) devised the , a kinematic device using two circular motions to produce linear oscillation, which eliminated the need for Ptolemy's equant in planetary models while preserving geocentric predictions of observed irregularities like Mercury's tight cycles near . These contributions emphasized empirical refinement, compiling tables for predictive astronomy based on long-term records of retrogrades and eclipses. In the early 16th century, (1473–1543) proposed a heliocentric model in his 1543 work , placing the Sun at the center with and other orbiting it in circular paths, often with epicycles to account for irregularities. This system simplified explanations of retrogrades and seasonal variations compared to geocentric models, though it retained uniform circular motion assumptions, setting the stage for later refinements.

Kepler's Contributions

In the early 1600s, utilized the unprecedentedly precise astronomical observations compiled by from the 1570s through 1601 to investigate planetary motions, with a particular emphasis on the challenging for Mars. Brahe, a Danish nobleman and leading , had amassed these records without the aid of telescopes, achieving accuracies far superior to prior efforts, which Kepler accessed after joining Brahe's team in in 1600 and inheriting the full archive following Brahe's death in 1601. Over several years of intensive computation, Kepler rejected the prevailing assumptions of circular orbits and uniform speeds, instead seeking mathematical patterns that fit Brahe's data empirically. Kepler formulated three empirical laws of planetary motion based on this analysis. The first law describes planetary paths as ellipses, with the Sun positioned at one focus rather than the center, fundamentally altering the geometric understanding of orbits from the perfect circles posited in ancient models. The second law, the principle of equal areas, states that an imaginary line connecting a planet to the Sun sweeps out equal areas in equal time intervals, implying variable orbital speeds—faster near the Sun and slower farther away—to conserve . The third law establishes a proportional relationship between a planet's T and the semi-major axis a of its elliptical path, expressed as T^2 \propto a^3, applicable across all planets and enabling predictions of periods from distances or vice versa. Kepler detailed his first two laws in the 1609 publication (New Astronomy), a comprehensive dedicated to explaining the motions of Mars using Brahe's observations as the foundation for a causal physical astronomy. This work spanned over five years of Kepler's efforts and marked a pivotal empirical breakthrough, as the elliptical model for Mars required abandoning epicycles and deferents from Ptolemaic theory. He introduced the third law a decade later in (The Harmony of the World, 1619), where it emerged from comparative analysis of multiple planetary datasets, completing his quantitative framework for heliocentric planetary dynamics. These laws catalyzed the transition from geocentric to heliocentric models by providing verifiable, data-driven descriptions that supported Copernicus's sun-centered universe without relying on philosophical ideals. By demonstrating elliptical orbits, Kepler's contributions dismantled the Aristotelian doctrine of celestial perfection through uniform , paving the way for a mechanistic view of the grounded in observation. Later, these empirical relations were theoretically unified under Isaac Newton's laws of motion and gravitation in the late .

Newton's Synthesis

In his 1687 work , synthesized the laws of motion with the concept of to establish the theoretical foundations of . In Proposition 1 of Book I, Newton demonstrated that bodies moving in orbits describe equal areas in equal times when radii are drawn to a fixed center of force, a property arising from the impulsive nature of centripetal acceleration in curvilinear paths. He initially applied this to circular orbits, where the force required to maintain motion is directed toward the center and proportional to the square of the speed divided by the radius, but extended the reasoning to elliptical paths by considering polygonal approximations that converge to conic sections under continuous central forces. Central to this synthesis was Newton's of gravitation, which posits that every particle attracts every other with a force proportional to the product of their masses and inversely proportional to the square of the distance between them: F = G \frac{m_1 m_2}{r^2}, where G is the . Newton showed that this generates conic-section orbits—ellipses for bound planetary motions, parabolas for marginally escaping trajectories, and hyperbolas for unbound paths—thus providing a physical for the elliptical orbits observed in the solar system. A pivotal was Newton's between terrestrial and celestial motion: the Moon's orbit around results from its continual "fall" toward the planet, deflected by tangential velocity into a curved rather than a straight descent, mirroring the fall of an apple but at a scale where inverse-square diminution allows stable revolution. Similarly, he interpreted cometary trajectories as parabolic or orbits under the same gravitational influence, explaining their transient appearances and high velocities as unbound excursions through the solar system. Newton's theory faced initial skepticism, particularly from Continental philosophers like and , who criticized its reliance on action-at-a-distance without a explanation for . By the early 18th century, however, empirical successes such as Edmund predictions fostered wider acceptance, profoundly influencing subsequent mathematicians like and Leonhard Euler, who built upon it to refine .

Newtonian Mechanics

Two-Body Problem

In Newtonian mechanics, the describes the motion of two point masses interacting through a central , such as the gravitational attraction between a and its . This system can be exactly solved by reducing it to an equivalent one-body problem. The relative motion of the two bodies is equivalent to the motion of a single fictitious particle with \mu = \frac{m_1 m_2}{m_1 + m_2}, where m_1 and m_2 are the masses of the two bodies, orbiting a fixed point at their common under the same central force law. This reduction simplifies the vector equations of motion into a scalar central force problem in the plane perpendicular to the , preserving the essential dynamics while eliminating the need to track both bodies separately. A key consequence of the central force in the reduced one-body problem is the conservation of angular momentum. The angular momentum vector \mathbf{L} = \mu \mathbf{r} \times \mathbf{v} remains constant in both magnitude and direction due to the absence of external torques, where \mathbf{r} is the position vector from the center of force and \mathbf{v} is the velocity. In scalar terms, the magnitude L = \mu v r \sin \theta is constant, where \theta is the angle between \mathbf{r} and \mathbf{v}. This conservation implies a constant areal velocity, given by \frac{dA}{dt} = \frac{L}{2\mu}, meaning the orbiting body sweeps out equal areas in equal times regardless of its position in the orbit. To determine the shape of the orbit, the equations of motion are transformed into polar coordinates using the substitution u = 1/r. This leads to the orbital differential equation: \frac{d^2 u}{d\theta^2} + u = -\frac{f(1/u)}{h^2 u^2}, where h = L/\mu is the specific angular momentum and f(r) is the central acceleration (force per unit reduced mass). For the inverse-square gravitational force, the acceleration f(r) = -\frac{G(m_1 + m_2)}{r^2}. Substituting yields: \frac{d^2 u}{d\theta^2} + u = \frac{G(m_1 + m_2)}{h^2}. This is the equation of a harmonic oscillator shifted by a constant, yielding solutions that are conic sections in polar form. The general solution for the inverse-square case is the polar equation of a conic section: r = \frac{h^2 / [G(m_1 + m_2)]}{1 + e \cos \theta}, where e is the determining the conic type (e < 1 for ellipses, e = 1 for parabolas, e > 1 for hyperbolas) and \theta is the measured from periapsis. This equation fully characterizes the ideal two-body orbit, providing the foundation from which empirical laws, such as Kepler's, emerge as geometric consequences.

Gravitational Potential and Energy

In the under Newtonian , the gravitational potential energy V between a central M and an orbiting m at separation r is given by V(r) = -\frac{G M m}{r}, where G is the ; this expression assumes a point- central and derives from integrating the over distance. The total E of the orbiting body is the sum of its K = \frac{1}{2} m v^2, where v is the speed, and the gravitational potential V, yielding E = \frac{1}{2} m v^2 - \frac{G M m}{r}. This total energy remains constant throughout the orbit due to the conservative nature of the gravitational . For circular orbits, the balances the gravitational attraction with the required centripetal acceleration, resulting in v = \sqrt{G M / r}. Substituting this into the energy expression gives E = -\frac{G M m}{2 r}, illustrating that circular orbits are bound with negative total . The generalizes the speed-energy relation for any conic-section orbit, stating v^2 = G M \left( \frac{2}{r} - \frac{1}{a} \right), where a is the semi-major axis; this equation stems from and connects instantaneous velocity to orbital parameters. The sign of the total energy classifies orbits: E < 0 corresponds to bound elliptical orbits (including circles), E = 0 to marginally unbound parabolic trajectories, and E > 0 to unbound paths. The , defined as \varepsilon = E / m = -G M / (2 a), provides a per-unit-mass measure that is negative for bound orbits and scales inversely with the semi-major axis. in this context complements conservation as a key principle governing two-body motion.

Orbital Shapes and Kepler's Laws

In Newtonian mechanics, the shape of an orbit under a central inverse-square gravitational force is determined by solving the , where the trajectory follows a conic with the center of at one . For bound orbits with negative total energy, the conic section is an , characterized by its semi-major axis a and e < 1. This result emerges from the differential equation of motion derived from Newton's second law and the inverse-square law of gravitation, \mathbf{F} = -\frac{G M m}{r^2} \hat{r}, leading to the polar form of the orbit equation r = \frac{h^2 / \mu}{1 + e \cos \theta}, where h is the specific angular momentum, \mu = G(M + m) is the standard gravitational parameter, and \theta is the true anomaly. Kepler's first law states that the orbit is an ellipse with the central body at one focus, which follows directly from this conic section solution for the inverse-square force. To derive it, consider the acceleration in polar coordinates under a central force f(r) = -k / r^2, where k = G M m. The radial equation of motion is \ddot{r} - r \dot{\theta}^2 = -k / m r^2, and using the conservation of angular momentum h = r^2 \dot{\theta} (constant), substitute \dot{\theta} = h / r^2 to get \ddot{r} - h^2 / r^3 = - \mu / r^2, with \mu = k / m. Changing variables to u = 1/r and differentiating with respect to time yields the Binet equation \frac{d^2 u}{d\theta^2} + u = \mu / h^2, whose general solution is u = (\mu / h^2) + A \cos(\theta - \theta_0), or equivalently r = \frac{h^2 / \mu}{1 + e \cos \theta} with e = A h^2 / \mu. For bound orbits where the energy is negative, e < 1, confirming the elliptical shape. Kepler's second law, which asserts that a line joining the orbiting body to the central mass sweeps out equal areas in equal times, arises from the conservation of angular momentum in a central force field. The areal velocity is \frac{dA}{dt} = \frac{1}{2} r^2 \dot{\theta}. Since the torque \boldsymbol{\tau} = \mathbf{r} \times \mathbf{F} = 0 for a central force, angular momentum \mathbf{L} = m \mathbf{r} \times \mathbf{v} is conserved, so the specific angular momentum h = r^2 \dot{\theta} is constant. Thus, \frac{dA}{dt} = \frac{h}{2}, a constant, independent of the specific form of the central force—though the inverse-square law ensures closed orbits. This holds for the reduced mass in the two-body problem, with h per unit reduced mass. Kepler's third law relates the orbital period T to the semi-major axis a via T^2 = \frac{4\pi^2}{\mu} a^3, where \mu = G(M + m) \approx G M for m \ll M. This can be derived by combining the vis-viva equation, which expresses the specific energy \varepsilon = \frac{v^2}{2} - \frac{\mu}{r} = -\frac{\mu}{2a} for elliptical orbits, with the conservation of angular momentum. For a circular orbit approximation as a starting point, balance centripetal acceleration with gravity: \frac{v^2}{r} = \frac{\mu}{r^2}, so v = \sqrt{\mu / r}. Since v = 2\pi r / T, substitute to get T^2 = \frac{4\pi^2 r^3}{\mu}. Extending to ellipses, the time-averaged form uses the semi-major axis a in place of r, confirmed by integrating over the orbit or using the constant energy, yielding the same relation because the period depends on the total energy, which scales with $1/a. The eccentricity e, which quantifies the deviation from a circular orbit, ties directly to the specific energy \varepsilon and specific angular momentum h through the formula e = \sqrt{1 + \frac{2 \varepsilon h^2}{\mu^2}}. For bound elliptical orbits, \varepsilon < 0, ensuring e < 1; as |\varepsilon| decreases (less negative), e approaches 1 for highly elongated ellipses, while higher h for fixed \varepsilon reduces e toward circular. This relation follows from substituting the vis-viva energy and the orbit equation parameters, where the semi-latus rectum p = h^2 / \mu = a (1 - e^2), and solving for e using \varepsilon = -\mu / (2a).

Advanced Dynamics

Many-Body Interactions

In the Newtonian framework, the n-body problem describes the motion of multiple gravitationally interacting bodies, but unlike the two-body case, no general closed-form analytical exists for n > 2. This absence stems from the nonlinear coupling of gravitational forces, leading to highly sensitive dependence on initial conditions and inherent chaos in systems like the . A notable example is the figure-eight orbit, a periodic for three equal-mass bodies discovered numerically and rigorously proven to exist, where the three equal-mass bodies chase each other along a symmetric figure-eight path. To address the lack of exact solutions, perturbation theory provides approximations by treating smaller masses as disturbances to a dominant two-body interaction, expanding the dynamics in series for small perturbations. A seminal application is Lagrange's lunar theory from the late 18th century, which uses first-order expansions to model the Moon's orbit perturbed by the Sun's gravity, accounting for variations in eccentricity and inclination without solving the full three-body equations. In the restricted three-body problem, where one body has negligible mass, equilibrium points known as Lagrange points L1 through L5 emerge as stable or unstable configurations; L1, L2, and L3 lie along the line connecting the two primaries, while L4 and L5 form equilateral triangles, enabling long-term spacecraft positioning. Relatedly, the Hill sphere defines the region around a secondary body (e.g., a planet) where its gravity dominates over the primary (e.g., a star), approximated as the volume within which satellites can be retained against perturbations, with radius scaling as the cube root of the mass ratio. For precise modeling of complex systems, is indispensable, employing methods like Runge-Kutta for general accuracy or integrators to preserve and structure over long times. integrators, such as the Wisdom-Holman mapping, are particularly vital for solar system simulations, enabling efficient computation of planetary orbits over billions of years by splitting the into integrable parts while minimizing secular errors. These techniques reveal the instability and chaotic evolution in many-body systems, underscoring the need for high-fidelity approximations in .

Relativistic Orbits

In general relativity, orbital motion around a central mass is described by geodesics in curved spacetime, leading to deviations from Newtonian predictions that become significant in strong gravitational fields or for high-precision measurements. The Schwarzschild metric, which models the spacetime exterior to a non-rotating, spherically symmetric mass M, provides the foundation for analyzing such orbits. For test particles, the radial equation of motion can be recast using an effective potential that incorporates relativistic corrections. In the post-Newtonian approximation, valid for weakly relativistic systems, this effective potential takes the form V_\text{eff} = -\frac{GM}{r} + \frac{L^2}{2 r^2} - \frac{GML^2}{c^2 r^3}, where L is the specific angular momentum (per unit mass) and V_\text{eff} is per unit mass; the final term arises from the first-order post-Newtonian expansion of the Schwarzschild geodesic equations. This correction modifies the orbital shape, causing the ellipse to precess rather than close after each revolution, unlike in the Newtonian two-body problem. The most prominent relativistic effect on bound orbits is the precession of the perihelion, quantified by the advance in argument of periapsis per orbital revolution: \Delta\phi = \frac{6\pi GM}{c^2 a (1 - e^2)}, where a is the semi-major axis and e is the eccentricity. For Mercury's orbit around the Sun, with M as the solar mass, this formula predicts an anomalous precession of 43 arcseconds per century, precisely accounting for the long-observed discrepancy after subtracting classical perturbations. This result, derived from the full field equations of general relativity, confirmed the theory's validity in 1915 and remains a cornerstone for precision tests in solar system dynamics. In binary systems, relativistic orbits are further altered by the emission of , which carry away and , causing the orbit to shrink and circularize over time. For quasi-circular inspirals, the averaged power radiated is \left\langle \frac{dE}{dt} \right\rangle \propto \frac{(G \mathcal{M}_\text{chirp})^{5/3} (\pi f)^{10/3}}{c^5}, where \mathcal{M}_\text{chirp} is the and f is the (twice the orbital frequency). This drives the binary through a prolonged inspiral , culminating in merger when the separation approaches the (ISCO). For non-spinning black holes described by the , the ISCO occurs at radial coordinate r = 6GM/c^2, determined by the of the where circular orbits transition from stable to unstable. Closer in, at the r = 3GM/c^2, null geodesics can form unstable circular orbits, marking the limit for light rays grazing the . These features profoundly influence accretion dynamics and gravitational-wave signals from extreme environments.

Orbital Parameters

Elements and Planes

The classical orbital elements provide a standard set of six parameters to uniquely specify the size, shape, and orientation of an orbit in relative to a chosen reference frame. These elements, originally derived from Kepler's laws and refined in , consist of the semi-major axis a, which determines the orbit's size as half the length of the major axis of the elliptical path; the eccentricity e, which describes the orbit's shape with $0 \leq e < 1 for ellipses (circular for e=0, more elongated as e approaches 1); the inclination i, the angle between the orbital plane and the reference plane (typically ranging from 0° to 180°); the longitude of the ascending node \Omega, the angle from the reference direction to the ascending node measured in the reference plane; the argument of periapsis \omega, the angle from the ascending node to the periapsis point within the orbital plane; and the true anomaly \nu, the angle from the periapsis to the current position of the orbiting body measured from the focus. The orbital plane is defined relative to a reference plane, such as the ecliptic for solar system orbits or the equatorial plane for Earth-centered orbits, serving as the fundamental plane against which the orbit's tilt is measured. The inclination i quantifies this tilt as the angle between the two planes, with i=0^\circ indicating a coplanar orbit and i=90^\circ a polar orbit perpendicular to the reference. The ascending node is the point where the orbiting body crosses the reference plane moving in the positive direction (e.g., from south to north for equatorial reference), and \Omega locates this intersection angularly from a fixed reference direction, such as the vernal equinox. As an alternative to the classical elements, equinoctial elements offer improved numerical stability, particularly for orbits with low eccentricity (e \approx 0) or low inclination (i \approx 0^\circ), where classical formulations can introduce singularities or computational issues. These elements include the semi-major axis a (or sometimes the p), two parameters for eccentricity h = e \sin(\omega + \Omega) and k = e \cos(\omega + \Omega), two for inclination p = \tan(i/2) \sin \Omega and q = \tan(i/2) \cos \Omega, and a longitude-like angle L = \Omega + \omega + \nu (or variants thereof). They are particularly useful in trajectory optimization and perturbation analyses for all conic sections, including circular and equatorial orbits, by avoiding trigonometric singularities inherent in the classical set. To obtain classical orbital elements from a body's position and velocity in Cartesian coordinates (state vector), the process involves computing the specific angular momentum and eccentricity vectors to determine the orbital plane and focus, followed by applying successive rotation matrices to align the reference frame with the orbital frame and solve for each element. This conversion typically proceeds by first finding the node vector and inclination from the angular momentum, then deriving \Omega and \omega via projections, and finally obtaining a and e from energy considerations, with \nu from the position vector's direction.

Periods and Speeds

The sidereal orbital period T represents the time required for a celestial body to complete one full revolution around its primary in an elliptical orbit, measured relative to the fixed stars. This duration is derived from Kepler's third law, as generalized by Newton for the two-body problem, yielding T = 2\pi \sqrt{\frac{a^3}{\mu}}, where a is the semi-major axis and \mu = G(M + m) is the standard gravitational parameter (approximating \mu \approx GM when the secondary mass m is negligible compared to the primary M). This relation establishes that periods scale with the cube root of semi-major axes, enabling predictions of orbital timings from geometric parameters like a. The mean motion n, an angular rate characterizing the orbit's average progress, is defined as n = \frac{2\pi}{T} = \sqrt{\frac{\mu}{a^3}}. It serves as a fundamental input for propagating the body's position over time. The evolution of the true anomaly \nu, which measures the angle from periapsis to the current position, is governed by : M = E - e \sin E, where M = n(t - \tau) is the mean anomaly (t is time and \tau the time of periapsis passage), E is the eccentric anomaly, and e is the eccentricity. Solving this transcendental equation iteratively yields E, from which \nu follows via \tan\frac{\nu}{2} = \sqrt{\frac{1+e}{1-e}} \tan\frac{E}{2}, providing the angular position as a function of time. Orbital speeds vary along the trajectory due to conservation of energy and angular momentum. In polar coordinates centered at the primary, the velocity components are the radial speed v_r = \frac{\mu e}{h} \sin \nu and the tangential speed v_\theta = \frac{\mu}{h} (1 + e \cos \nu), where h = \sqrt{\mu p} is the specific and p = a(1 - e^2) the semi-latus rectum. These expressions highlight maximum radial velocity at \nu = 90^\circ and tangential velocity peaking at periapsis (\nu = 0), with the total speed given by the vis-viva equation v = \sqrt{\mu \left( \frac{2}{r} - \frac{1}{a} \right)}. For circular orbits (e = 0), speeds simplify to a constant v = \sqrt{\frac{\mu}{a}}. For relative motions, such as between planets, the synodic period S quantifies the recurrence of alignment as viewed from one body, calculated as \frac{1}{S} = \left| \frac{1}{P} - \frac{1}{Q} \right|, where P and Q are the sidereal periods of the two bodies (assuming P < Q). This period governs observable phenomena like planetary conjunctions; for Earth (P \approx 365.25 days) and Mars (Q \approx 687 days), S \approx 780 days.

Perturbations

Perturbation Types

In orbital mechanics, perturbations refer to small deviations from the unperturbed two-body trajectory, where the motion is governed solely by the central gravitational force. These deviations arise from additional forces and are typically analyzed by resolving the perturbing acceleration into three orthogonal components in the orbital reference frame: radial (along the position vector from the primary body, denoted as \Delta v_r or R), transverse (in the orbital plane, perpendicular to the radial direction and aligned with the velocity for circular orbits, denoted as \Delta v_\theta or T), and normal (perpendicular to the orbital plane, denoted as \Delta v_n or N). The radial component primarily affects the semi-major axis and eccentricity, the transverse component influences the angular momentum and thus the semi-major axis and argument of perigee, while the normal component alters the inclination and longitude of the ascending node. The evolution of classical orbital elements under these perturbations is described by Gauss's variational equations, which provide the time rates of change of the elements in response to the perturbing accelerations. For impulsive perturbations, these equations quantify instantaneous changes, but they apply more generally to continuous forces. A key example is the rate of change of the semi-major axis a: \frac{da}{dt} = \frac{2 a^2}{h} \left[ e \sin \nu \, \Delta v_r + \frac{p}{r} \Delta v_\theta \right], where h is the specific angular momentum, e is the eccentricity, \nu is the true anomaly, p = a(1 - e^2) is the semi-latus rectum, and r is the radial distance. Similar expressions exist for other elements, such as eccentricity e, inclination i, and the argument of perigee \omega, showing how radial and transverse components drive in-plane changes, while normal components affect out-of-plane elements. These equations enable the prediction of element variations without full numerical integration of the trajectory. Perturbations are further classified by their temporal behavior into secular and periodic types. Secular perturbations produce long-term, unidirectional drifts in orbital elements, such as gradual precession of the apsides () or nodes, accumulating over many orbits without reversal. In contrast, periodic perturbations cause oscillatory variations in the elements that repeat over short timescales, often tied to the orbital period or harmonics, and average to zero over sufficiently long intervals. This distinction is crucial for distinguishing effects that require corrective maneuvers from those that self-correct. For computing perturbed orbits, Cowell's method offers a direct numerical approach to special perturbations by integrating the full equations of motion. It treats the trajectory as the vector sum of the two-body acceleration and all perturbing accelerations in Cartesian coordinates: \ddot{\mathbf{r}} = -\frac{\mu}{r^3} \mathbf{r} + \mathbf{a}_d, where \mathbf{a}_d encompasses all disturbing forces, integrated step-by-step using numerical propagators like . This method is computationally intensive but highly accurate for arbitrary perturbations, avoiding the need for element-specific formulations.

Stability and Decay

Orbital decay in low Earth orbits primarily results from atmospheric drag, which dissipates the satellite's kinetic energy through collisions with residual atmospheric particles. The drag force acting on a satellite is given by F_{\text{drag}} = \frac{1}{2} \rho v^2 C_d A, where \rho is the atmospheric density, v is the orbital velocity, C_d is the drag coefficient (typically around 2.2 for satellites), and A is the cross-sectional area perpendicular to the velocity vector. This force is tangential and opposite to the direction of motion, gradually reducing the orbital energy and causing the semi-major axis a to decrease over time. For circular low Earth orbits assuming constant density, the rate of change of the semi-major axis is approximately \frac{da}{dt} \approx -\frac{C_d A \rho}{m} \sqrt{G M a}, with m the satellite mass, G the gravitational constant, and M Earth's mass. Resonance effects from planetary perturbations can lead to long-term orbital instability by amplifying small deviations, resulting in depletion zones known as in the asteroid belt. These gaps arise from mean-motion resonances with , where the orbital periods of asteroids align commensurately with Jupiter's orbit, causing repeated gravitational tugs that excite eccentricities and inclinations over time. A prominent example is the 3:1 resonance at approximately 2.5 AU, where asteroids complete three orbits for every one of Jupiter's, leading to chaotic diffusion and eventual ejection from the resonance zone through secular perturbations. This mechanism has cleared the gap over billions of years, with numerical mappings showing that test particles in this resonance experience rapid eccentricity growth, preventing stable populations. In multi-body systems, such as the involving two massive bodies and a test particle, perturbations induce chaotic dynamics characterized by sensitivity to initial conditions, often quantified by Lyapunov exponents. The maximum Lyapunov exponent \lambda measures the exponential divergence rate of nearby trajectories, with positive values (\lambda > 0) indicating ; in the circular restricted , \lambda \approx 0.15 marks a to , while higher values (e.g., \lambda > [1](/page/1)) predict rapid unpredictability within years. This can culminate in orbital ejection, where the test particle's becomes positive, transitioning to a escape trajectory relative to the central system, as seen in simulations where unstable orbits intersect zero-velocity contours or Lagrange points, leading to ejections in a significant fraction of cases. Estimating the lifetime of satellites in low orbits under drag dominance involves simplified models that balance and parameters against atmospheric conditions. A rough for time is \tau \approx \frac{m}{C_d A \rho v}; this yields lifetimes from months at 200 km altitude to decades at 800 km, depending on activity influencing \rho. Such estimates guide mission planning, highlighting how lighter satellites with larger areas faster, often requiring for extension.

Special Phenomena

Oblateness and Tidal Effects

The oblateness of a central body, arising from its rotational flattening, introduces non-spherical components to its gravitational potential, perturbing nearby orbits. The dominant term in this perturbation is the zonal harmonic J₂, which quantifies the body's equatorial bulge. The perturbing potential due to J₂ is given by \delta V = -\frac{J_2 G M R_e^2}{2 r^3} (3 \sin^2 \phi - 1), where G is the gravitational constant, M is the mass of the central body, R_e is its equatorial radius, r is the distance from the center, and \phi is the geocentric latitude. This term causes secular changes in orbital elements, particularly a precession of the ascending node. The rate of nodal precession is \frac{d\Omega}{dt} = -\frac{3}{2} n \left( \frac{R_e}{a} \right)^2 J_2 \cos i, where n is the mean motion, a is the semi-major axis, and i is the inclination; for Earth, J₂ ≈ 1.083 × 10⁻³, leading to significant westward precession for low-inclination orbits. Tidal effects arise from the differential gravitational pull of an orbiting body on the extended central body, deforming it into two bulges aligned with the line connecting the centers of mass. If the central body's rotation is asynchronous with the orbit, friction within its interior or oceans lags these bulges, creating a misalignment that generates a torque. This torque transfers angular momentum from the central body's spin to the orbit, causing orbital expansion and rotational slowdown, while internal friction dissipates energy as heat. Over long timescales, this tidal torque drives the system toward , where the central body's rotation period matches the , achieving synchronous rotation and eliminating the torque. The exemplifies this, locked in a 27.3-day spin-orbit resonance with , always presenting the same face. Darwin's theory of tidal friction, developed in the late , predicted that such interactions evolve binary systems toward 1:1 spin-orbit resonance, with the Moon's recession at about 3.8 cm/year as ongoing evidence. If an orbiting body approaches too closely, forces can exceed its self-gravity, leading to disruption within the . For a , this distance is approximately d \approx 2.44 R \left( \frac{\rho_p}{\rho_m} \right)^{1/3}, where R is the primary's and \rho_p, \rho_m are the densities of the primary and , respectively; beyond this , stable orbits persist, but closer proximity results in shredding, as seen in Saturn's ring systems.

Exotic Orbits

Exotic orbits encompass configurations that deviate from the standard Keplerian conics due to multi-body gravitational influences or extreme conditions, often requiring advanced dynamical models to describe their and . These orbits highlight the limitations of two-body approximations and reveal complex behaviors in systems with resonant or perturbed interactions. Horseshoe and orbits represent co-orbital librations in the circular restricted , where two bodies of comparable mass share an orbit with nearly identical periods around a dominant central mass, leading to resonant exchanges rather than collisions. In horseshoe orbits, the secondary bodies librate around the L3 in a wide, U-shaped path in the co-rotating frame, while orbits involve narrower librations encircling the L4 or L5 points. Saturn's moons and exemplify a stable horseshoe configuration, discovered during the flyby, where the pair swaps orbital paths every approximately four Earth years due to their 1:1 , with semi-major axes of about 151,000 km. This setup is enabled by subtle many-body perturbations that maintain long-term over billions of years. Retrograde orbits, characterized by inclinations exceeding 90 degrees relative to the primary's equatorial or , result in motion opposite to the prograde direction and are typically associated with dynamical capture rather than in-situ formation. Such orbits often exhibit high and inclination, making them unstable over long timescales without mechanisms. In the Solar System, Saturn's outer moon follows a retrograde path with an inclination of about 151 degrees and a period of 550 days, indicative of capture from the . Theoretical models for exoplanetary systems predict retrograde exomoons around gas giants, particularly irregular ones captured during planetary scattering events, which could comprise a significant fraction of stable satellite populations in multi-planet architectures. In the restricted , zero-velocity curves delineate the boundaries of Hill's regions, which constrain the allowable positions for a test particle's motion by setting the Jacobi constant such that vanishes at these loci. These curves form closed contours around the primary bodies at higher energies, opening to permit transfers between regions at lower energies near the Lagrange points, thus defining forbidden zones inaccessible to bounded orbits. Hill's regions provide a phase-space of orbital accessibility, essential for analyzing capture, , and resonant behaviors in hierarchical systems like exoplanets. Theoretical exotic orbits include relativistic zoom-whirl trajectories around black holes, where inspiral dynamics produce alternating zoom (radial plunge) and whirl (near-circular ) phases due to strong-field . These orbits, studied in the post-Newtonian approximation for Kerr metrics, exhibit unstable spirals that emit characteristic before merger, contrasting classical bounded paths.

Applications

Astrodynamics

Astrodynamics is the branch of orbital mechanics focused on the practical application of gravitational dynamics to spacecraft navigation, trajectory optimization, and mission design. It encompasses the calculation of velocity changes required for transfers, insertions, and corrections to achieve desired orbital paths around celestial bodies. Central to astrodynamics are methods for solving boundary value problems and budgeting propellant usage through delta-v (Δv) analysis, enabling efficient spaceflight from launch to interplanetary exploration. A key problem in astrodynamics is , which determines the initial velocity vector required for a to travel between two specified position vectors in a finite time under two-body motion. This is essential for preliminary trajectory design in , reentry, and interplanetary transfers. The universal variable formulation, introduced by and Blanchard, unifies the solution across elliptic, parabolic, and hyperbolic regimes by using the variable χ, defined as χ = √|α| Δt where α is related to the semi-major axis and Δt is the transfer time. The solution involves solving a for χ using Stumpff functions c2(z) and c3(z), ensuring and convergence for all conic sections. Delta-v budgets quantify the total velocity change a must achieve for a , serving as a fundamental metric for system sizing via the . For launches to (), the budget typically totals around 9.4 km/s, accounting for the ideal insertion velocity of approximately 7.8 km/s plus losses from atmospheric drag (∼0.2 km/s) and (∼1.5 km/s during ascent). An approximation for the ideal Δv required to reach a at altitude h, neglecting losses, is given by Δv_launch ≈ √(2GM/R) [1 - √(R/(R + h))], where G is the , M is the central body's , and R is its radius; drag losses are added empirically based on vehicle and profile. This budget informs design, ensuring sufficient for achieving orbital velocity while countering Earth's rotational and gravitational effects. Orbit insertion maneuvers establish a trajectory after launch or transfer, often requiring precise to circularize or adjust parameters. For changes, which align the with a target inclination, the cost is minimized when performed at apogee where is lowest, following the vector addition = 2 v sin(Δi/2), with v as the local and Δi the inclination shift. This formula arises from the geometry of velocity vectors before and after the impulsive burn, assuming instantaneous thrust. For example, a 30° change in (v ≈ 7.8 km/s) demands ≈ 4.0 km/s, highlighting the expense of such maneuvers and the preference for launch site matching to reduce initial Δi. Interplanetary transfers rely on the patched conics approximation, which divides the trajectory into segments dominated by successive gravitational influences: heliocentric between planets and planetocentric near departure/arrival. This method simplifies n-body dynamics by matching velocity and position at spherical "spheres of influence" boundaries, enabling solutions for each leg. Gravity assists, or slingshots, further optimize by leveraging planetary motion to alter hyperbolic excess velocity, providing effective boosts without propellant. The Voyager missions exemplified this, using and Saturn flybys to gain ∼10-15 km/s in heliocentric speed, reducing required launch by over 50% compared to direct trajectories and enabling visits to all four outer planets.

Earth-Centric Orbits

Earth-centric orbits refer to the trajectories of artificial satellites and around , classified primarily by altitude and purpose, which determine their operational characteristics and environmental challenges. These orbits enable a wide range of applications, from to global communications, with design principles drawn from astrodynamics to achieve desired coverage and stability. Low Earth orbit (LEO), typically ranging from 160 km to 2,000 km in altitude, features short orbital periods of approximately 90 minutes, allowing satellites to complete multiple passes over specific regions daily. At these low altitudes, atmospheric drag from residual upper atmosphere molecules significantly perturbs orbits, causing gradual decay that requires periodic boosts to maintain altitude. The (ISS), operating at around 400 km, exemplifies LEO use for microgravity research and serves as a hub for crewed missions, where drag-induced orbital lowering necessitates regular reboost maneuvers to counteract decay rates that can reach several kilometers per month during high solar activity. Medium Earth orbit (MEO) spans altitudes from 2,000 km to 35,786 km, encompassing regions above the densest atmospheric effects but within zones of intense radiation. Satellites in MEO, such as the Global Positioning System (GPS) constellation at approximately 20,200 km, maintain 12-hour orbital periods that enable consistent global coverage with a minimum of 24 satellites. This altitude balances visibility for ground receivers with reduced drag compared to LEO, though MEO satellites must incorporate robust shielding against radiation. Geostationary orbit (GEO), at precisely 35,786 km above the , provides a 24-hour matching , resulting in a zero inclination for geostationary configurations where appear fixed in the sky relative to ground observers. This fixed positioning is ideal for continuous and , as the remains over the same without apparent motion. Achieving and maintaining GEO requires precise launches to the equatorial plane, with station-keeping maneuvers to counter drifts from lunar and solar perturbations. Specific perturbations in Earth-centric orbits are amplified by proximity to the planet. The Van Allen radiation belts, consisting of trapped high-energy protons and electrons between about 1,000 km and 60,000 km altitude, pose significant risks to electronics and materials in MEO and higher satellites through total ionizing dose and single-event effects, necessitating specialized radiation-hardened components. In , exposure is lower but notable in the , where belts dip closer to . Additionally, Earth's oblateness, modeled by the J2 gravitational harmonic, induces stronger perturbations in low-altitude orbits like , causing rapid of the (nodal ) and argument of perigee at rates up to several degrees per day, which must be accounted for in mission planning to avoid unintended shifts in ground tracks.

Planetary and Exoplanet Orbits

In the Solar System, planetary orbits exhibit a clear progression from the inner terrestrial planets to the outer gas giants, governed by Kepler's laws and gravitational influences. The inner rocky planets—Mercury, , , and Mars—occupy compact orbits with relatively short periods ranging from 88 days for Mercury to 687 days for Mars, and generally low eccentricities that keep their paths nearly circular, except for Mercury's notably higher value of 0.2056. These close-in orbits result in higher orbital speeds and stronger heating, contributing to the planets' diverse surface conditions. In contrast, the outer gas giants—Jupiter, Saturn, , and —trace wider, more elliptical paths with much longer periods, from 11.86 years for Jupiter to 164.8 years for ; their eccentricities are modest but allow for greater dynamical interactions, such as the 5:2 mean-motion between and Saturn, where completes five orbits for every two of Saturn's, stabilizing their relative positions over billions of years. Exoplanetary systems, with over 7,800 confirmed exoplanets as of November 2025, reveal a broader diversity of orbital architectures beyond the Solar System's orderly arrangement. Hot Jupiters, massive gas giants orbiting perilously close to their host stars (often within 0.05 ), typically exhibit short periods of days and initially high eccentricities that tidal forces circularize over time, leading to inflated atmospheres and extreme temperatures exceeding 1,000 K. These planets, first identified in the , challenge formation theories as they likely migrate inward from cooler outer regions. Circumbinary orbits, where planets encircle pairs, add further exoticism; , a Saturn-mass world with a 229-day period at about 0.7 from its binary hosts, exemplifies this configuration, detected via the transit method that precisely measures semi-major axis a and inclination i. The transit technique has been instrumental in cataloging such systems, enabling statistical insights into their prevalence around roughly 1-2% of binary stars. Mean-motion resonances play a crucial role in maintaining across these systems, linking orbital periods in integer ratios to mitigate chaotic perturbations. In the Solar System, the Jupiter-Saturn 5:2 resonance exemplifies this, influencing long-term orbital evolution and contributing to the "Great Inequality" in their positions. examples abound, such as the system, a compact of seven Earth-sized planets around an star, locked in a series of first-order resonances (e.g., 8:5, 5:3) that form a Laplace , ensuring dynamical despite their close spacing at 0.01-0.06 . This resonant architecture, spanning periods from 1.5 to 12 days, highlights how such configurations can persist post-formation, even in multi-planet setups prone to instability. Orbital scaling in these systems follows Kepler's third law, which relates a planet's orbital radius r to its period T as r \propto T^{2/3} for bodies orbiting the same central mass, providing a framework to estimate . For Sun-like stars, this scaling delineates the —where liquid water could exist on rocky planets—at approximately 0.95 to 1.67 , encompassing and allowing for Earth-like worlds with periods of 300-500 days. In exoplanet surveys, applying this law to data helps identify candidates in these zones, such as those around Kepler and TESS targets, underscoring the law's enduring utility in probing potential biospheres despite variations from stellar type and planetary atmosphere.

References

  1. [1]
    What Is an Orbit? | NASA Space Place – NASA Science for Kids
    The Short Answer: An orbit is a regular, repeating path that one object in space takes around another one. An object in an orbit is called a satellite.
  2. [2]
    Chapter 3: Gravity & Mechanics - NASA Science
    Jan 16, 2025 · Sir Isaac Newton realized that the force that makes apples fall to the ground is the same force that makes the planets "fall" around the Sun.
  3. [3]
    Orbits and Kepler's Laws - NASA Science
    May 21, 2024 · Kepler's three laws describe how planetary bodies orbit the Sun. They describe how (1) planets move in elliptical orbits with the Sun as a focus.
  4. [4]
    Universe Glossary A-G - NASA Science
    It is defined as the body's mass multiplied by its velocity and its distance from the center of motion. In a closed system, it is a conserved quantity in ...<|control11|><|separator|>
  5. [5]
    Catalog of Earth Satellite Orbits
    Sep 4, 2009 · There are essentially three types of Earth orbits: high Earth orbit, medium Earth orbit, and low Earth orbit.
  6. [6]
    Chapter 5: Planetary Orbits - NASA Science
    As an orbiting body crosses the reference plane going north, the node is referred to as the ascending node; going south, it is the descending node. The ...
  7. [7]
    Description of Orbits and Ephemerides - JPL Solar System Dynamics
    If one object is much more massive than the other (like the Sun), it is the small object that does most of the moving and is said to be “orbiting” the massive ...
  8. [8]
    ESA - Types of orbits - European Space Agency
    An orbit is the curved path that an object in space (like a star, planet, moon, asteroid or spacecraft) follows around another object due to gravity.Missing: mechanics | Show results with:mechanics
  9. [9]
    [PDF] 1 CHAPTER 2 CONIC SECTIONS 2.1 Introduction A particle moving ...
    Mar 2, 2019 · A particle moving under the influence of an inverse square force moves in an orbit that is a conic section; that is to say an ellipse, ...
  10. [10]
    [PDF] Introduction to Orbits - NASA Technical Reports Server (NTRS)
    Orbits are the paths a spacecraft takes in space;. They are used to go from one location in space to another, for example, from the ...
  11. [11]
    What Is an Orbit? (Grades 5-8) - NASA
    Jul 7, 2010 · An orbit is a regular, repeating path that one object in space takes around another one. An object in an orbit is called a satellite.
  12. [12]
    [PDF] Fundamentals of Orbital Mechanics - NASA
    Jan 1, 2000 · 7.5 Types of Orbits. Trajectories of space objects are not only classified with regard to the theory by which their ephemerides are generated ...
  13. [13]
    Lecture 19: Orbits
    Oct 10, 2010 · The shape of an orbit is a conic section with the center of mass at one focus. There are two parts to Newton's formulation of Kepler's First Law ...
  14. [14]
    Circular Orbit - an overview | ScienceDirect Topics
    A circular orbit is defined as a special case of an elliptical orbit where the eccentricity is zero (e = 0), resulting in a constant radial distance from ...
  15. [15]
    Chapter 4: Trajectories - NASA Science
    Nov 4, 2024 · Hohmann Transfer Orbits​​ To launch a spacecraft from Earth to an outer planet such as Mars using the least propellant possible, first consider ...
  16. [16]
    Satellite Orbits
    A zero inclination orbit would mean the satellite is orbiting directly over the equator, an inclination of 90 degrees is a perfectly polar orbit.
  17. [17]
    Babylonian Astronomy – Robert Hatch
    The motions of the planets are rather complicated when viewed from earth. We believe we know why on the basis of Newtonian physics, as amended by Einstein ...
  18. [18]
    Kidinnu and ancient Babylonian astronomy
    A similar system was used for the movements of the sun and the five planets (which the Babylonians called Nabu, Istar, Nergal, Marduk and Ninurta). This is ...
  19. [19]
    Astronomy the Babylonian Way - NASA ADS
    ... positions of the Sun and Moon, and occasionally of the planets and stars. Babylonian motivation for astronomical recording seems to have been primarily ...
  20. [20]
    Introduction to Aristotle's Celestial and Terrestrial Physics
    Nov 7, 2012 · By nature celestial bodies move continuously; their motion is circular, everlasting, without beginning or end. The heavenly bodies, including ...
  21. [21]
    Aristotle - Department of Physics and Astronomy
    ARISTOTLE. (384-322 B.C.) Stagira. geocentrist ... natural movement. earthly elements: earth & water ... celestial bodies on spheres w/ circular motions
  22. [22]
    Ptolemy - Ptolemaic model
    He refined the geometric model of the Solar system using epicycles, deferents, and equants to explain the motion of the planets. In the Ptolemaic model ...
  23. [23]
    ptolemy_epicycle_equant.html - UNLV Physics
    The geometric center of the deferent is the center of the Solar System which was the whole universe to Ptolemy. The Earth is displaced from the geometric center ...
  24. [24]
    [PDF] The Ptolemaic System: A Detailed Synopsis
    Apr 3, 2015 · Ptolemy also used an equant point but, unlike the others, the other side of the earth from the center of the eccentric deferent! tolemaic ...<|separator|>
  25. [25]
    Early Astronomy - Northern Arizona University
    Like most ancient people, Babylonians believed that studying planetary movements could help them predict the future. In fact, according to a biblical story ...
  26. [26]
    Astronomers A
    Al-Battani, Abu Allah Mohammed ibn Jabir (850-928) - also known as Albategniuis, showed that the distance from the Earth to the Sun changes during the year. ...
  27. [27]
    [PDF] To Save the Phenomena - MIT OpenCourseWare
    Copernicus adopted the Tusi couple in his De Revolutionibus to avoid. Ptolemy's equant, without attribution. Copernicus, De Revolutionibus,. 1543. Arabic ...<|separator|>
  28. [28]
    Kepler's Laws of Planetary Motion - NASA Earth Observatory
    Jul 7, 2009 · Thinking on Kepler's laws, Newton realized that all motion, whether it was the orbit of the Moon around the Earth or an apple falling from a ...
  29. [29]
    The Watershed :Tycho Brahe & Johannes Kepler
    Published the results of his work in 1609 in the Astronomia Nova (New Astronomy), which contained the first and second of his three laws of planetary motion.
  30. [30]
    Kepler and the First Law of Planetary Motion - UC Berkeley math
    After eight years of calculations, Kepler's first law of planetary motion was incorporated into the Astronomia nova, along with his second law: a planet sweeps ...
  31. [31]
    Orbits and Kepler's Laws - NASA Science
    May 2, 2024 · What Newton realized was that the orbits of objects in space depend on their masses, which led him to discover gravity. Newton's generalized ...<|control11|><|separator|>
  32. [32]
    Classics of Astronomy by Johannes Kepler - Williams College
    Nov 23, 2010 · Kepler's five years of research into Mars' orbit therefore began with using the observations of Mars Tycho supplied to try to investigate the ...
  33. [33]
    Johannes Kepler (1571–1630) | High Altitude Observatory
    Kepler's first two Laws of Planetary Motion were first outlined in his 1609 "Astronomia Nova," but laid out in detail together with his Third Law in book IV of ...Johannes Kepler (1571--1630) · Kepler's Observations · The Rudolphine Tables<|control11|><|separator|>
  34. [34]
    Johannes Kepler: Everything you need to know - Space
    Dec 22, 2023 · Johannes Kepler was an astronomer best known for his three laws of planetary motion, which describe how the planets move in ellipses around the ...
  35. [35]
    Newton's Philosophiae Naturalis Principia Mathematica
    Dec 20, 2007 · By reason of this deviation of the Sun from the center of gravity the centripetal force does not always tend to that immobile center, and hence ...
  36. [36]
    Newton's theory of "Universal Gravitation" - PWG Home - NASA
    Newton went further and proposed that gravity was a "universal" force, and that the Sun's gravity was what held planets in their orbits.
  37. [37]
    The Key to Newton's Dynamics - UC Press E-Books Collection
    Proposition 1— Theorem 1. The areas that bodies driven in orbits describe with radii having been constructed to a stationary center of forces, ...
  38. [38]
    The Apple, the Moon, and the Inverse Square Law
    The inverse square law proposed by Newton suggests that the force of gravity acting between any two objects is inversely proportional to the square of the ...Missing: conic | Show results with:conic
  39. [39]
    Newton and the comets - Comètes
    The orbits of the planets are ellipses in which the Sun occupies one of the foci, or sometimes - in the case of comets - parabolas, in which the Sun occupies ...
  40. [40]
    Newton's Philosophy
    Oct 13, 2006 · The criticisms of Newton's theory of gravity by Leibniz and Huygens, outlined briefly above, would prove essential to the Continental reception ...
  41. [41]
    (PDF) The Reception of Newton's Principia - ResearchGate
    Jul 9, 2015 · Newton's Principia, when it appeared in 1687, was received with the greatest admiration, not only by the foremost mathematicians and astronomers in Europe.
  42. [42]
    [PDF] The Two-Body Problem - UCSB Physics
    is the reduced mass of the system. Thus, our problem has effectively been reduced to a one-particle system - mathematically, it is no different than a single.
  43. [43]
    11.2: Equivalent one-body Representation for two-body motion
    Jul 5, 2021 · This reduction to an equivalent one-body problem greatly simplifies the solution of the motion, but it misrepresents the actual trajectories and ...
  44. [44]
    [PDF] The Two-Body problem - DAMTP
    Thus, is we effectively reduced the 2-body problem to a 1-body problem. Second symmetry, isotropy of space, implies that the vector of the angular momentum ...
  45. [45]
    [PDF] CHAPTER 3 - The Two-Body Central Force Problem
    2 dt. The conservation of angular momentum is thus equivalent to saying the areal velocity is constant. Here we have the proof of the well-known Kepler's ...
  46. [46]
    13.5 Kepler's Laws of Planetary Motion - University Physics Volume 1
    Sep 19, 2016 · Since the angular momentum is constant, the areal velocity must also be constant. This is exactly Kepler's second law. As with Kepler's first ...
  47. [47]
    [PDF] Chapter 3 Two Body Central Forces - Rutgers Physics
    First we reduce it to a one-body problem, and then we reduce the dimensionality. 3.1.1 Reduction to a one-body problem. As there are no external forces, we ...
  48. [48]
    [PDF] Orbital Mechanics
    These notes provide an alternative and elegant derivation of Kepler's three laws for the motion of two bodies resulting from their gravitational force on each ...Missing: d²u/ dθ² +
  49. [49]
    [PDF] Expression of conic section in polar coordinates - Purdue Engineering
    Referring back to equation 1, this is the equation of a conic section in polar coordinates. e is the eccentricity and p is the semi-latus rectum. θ is the true ...
  50. [50]
    [PDF] Lecture D29 - Central Force Motion: Orbits - DSpace@MIT
    Recall that the gravitational potential per unit mass is given by −µ/r. That is, F /m = −V(−µ/r) = −(µ/r2)er.
  51. [51]
    13.3 Gravitational Potential Energy and Total Energy
    The total energy of a system is the sum of kinetic and gravitational potential energy, and this total energy is conserved in orbital motion. Objects must have ...
  52. [52]
    13.3: Gravitational Potential Energy and Total Energy
    Potential energy is particularly useful for forces that change with position, as the gravitational force does over large distances.
  53. [53]
    [PDF] Lecture 3: Planar Orbital Elements: True Anomaly, Eccentricity, and ...
    The Vis-Viva Equation for Elliptic/Hyperbolc Orbits. E = −. µ. 2a. Note for an elliptic orbit, the total energy is negative, as expected. This yields the ...Missing: Newtonian | Show results with:Newtonian
  54. [54]
    7.2: Kepler's Laws - Physics LibreTexts
    May 21, 2023 · Kepler's First Law: The paths of bodies trapped in orbits form closed ellipses, with the gravitating body at one of the foci.Kepler's First Law · Kepler's Third Law · Reconciling Kepler's Laws...
  55. [55]
    Deriving Kepler's Laws from the Inverse-Square Law - Galileo
    We present here a calculus-based derivation of Kepler's Laws. You should be familiar with the results, but need not worry about the details of the derivation.
  56. [56]
    [PDF] Newton's derivation of Kepler's laws (outline) - UTK Math
    In sections 2-8 we present Newton's derivation of Kepler's laws from the inverse-square law for gravity, which only uses basic calculus. (Nothing is original in ...
  57. [57]
    [PDF] Derivation of Kepler's laws from the inverse square ... - Purdue Math
    We use complex numbers and represent the motion as z(t). We first derive the equal area law for an object moving under any central force:.
  58. [58]
    Kepler's Second Law - Richard Fitzpatrick
    We conclude that Kepler's second law of planetary motion is a direct consequence of angular momentum conservation.
  59. [59]
    5.6: Kepler's Laws - Physics LibreTexts
    Jul 8, 2023 · We can derive Kepler's third law by starting with Newton's laws of motion and the universal law of gravitation. We can therefore demonstrate ...
  60. [60]
    phy105 - celestial mechanics - newton's derivation of kepler's laws
    Hence we have verified Kepler's second law. kepler III, Newton's form of Kepler's third law can be derived by considering two bodies of masses m1 and m2 ...<|control11|><|separator|>
  61. [61]
    The Orbit Equation - Orbital Mechanics & Astrodynamics
    The orbit equation describes conic sections, meaning that all orbits are one of four types, as shown in Fig. 31. The particular type of orbit is determined ...Missing: classification | Show results with:classification
  62. [62]
    [PDF] The Classical Gravitational N-Body Problem - arXiv
    Certainly, no general analytical solution is known. There are five equilibrium solutions, discovered by Euler and Lagrange. (see Fig.1). They lie at critical ...
  63. [63]
    A remarkable periodic solution of the three-body problem in the case ...
    Nov 1, 2000 · A remarkable periodic solution of the three-body problem in the case of equal masses. Authors:Alain Chenciner, Richard Montgomery.
  64. [64]
    Resonance | SpringerLink
    His lunar theory [22] is a pioneer work in many respects. We credit Delaunay ... Lagrange equations for the variation of the orbital elements under a perturbation ...
  65. [65]
    [PDF] The Lagrange Points - Wilkinson Microwave Anisotropy Probe
    The term \restricted" refers to the condition that two of the masses are very much heavier than the third. Today we know that the full three-body problem is ...
  66. [66]
    [PDF] Chaotic Capture of a Retrograde Moon by Venus and the Reversal ...
    Dec 28, 2023 · In a hierarchical three-body system, a planet is in the center of an imaginary volume traditionally called the Hill sphere, within which the ...Missing: retention | Show results with:retention
  67. [67]
    [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 ...
  68. [68]
    [PDF] A study of symplectic integrators for planetary system problems - arXiv
    Mar 1, 2017 · Explicit symplectic integrators can be interpreted as simply modifying rapidly oscillating terms in the N-body Hamiltonian (Wisdom & Holman 1991) ...
  69. [69]
    [PDF] On the unreasonable effectiveness of the post-Newtonian ... - arXiv
    Feb 25, 2011 · The post-Newtonian approximation solves Einstein's equations for slow, weak-field systems, but is effective in strong-field, fast-motion ...
  70. [70]
    Gravitational Self-Force Correction to the Innermost Stable Circular ...
    The innermost stable circular orbit (ISCO) of a test particle around a Schwarzschild black hole of mass M has (areal) radius r i s c o = 6 ⁢ M ⁢ G ⁡ / c 2 .Missing: primary | Show results with:primary
  71. [71]
    Understanding photon sphere and black hole shadow in ...
    May 31, 2019 · We have derived the differential equation governing the evolution of the photon sphere for dynamical black hole spacetimes with or without spherical symmetry.
  72. [72]
    Orbital Elements - Ulysses - ESA Cosmos - European Space Agency
    The definitions of the quantities making up the Classical Orbital Elements are as follows: Mean Distance (a) - the semi-major axis of the orbit measured in ...
  73. [73]
    [PDF] Describing Orbits
    We use inclination to define several different kinds of orbits. For example, an Earth orbit with an inclination of 0° or 180° is an equatorial orbit, because ...
  74. [74]
    [PDF] Cartesian State Vectors to Keplerian Orbit Elements
    Cartesian State Vectors −→ Keplerian Orbit Elements (Memorandum № 2). 2 Constants and Conversion Factors. Universal Constants. Symbol. Description. Value.
  75. [75]
    Deriving Kepler's Formula for Binary Stars - Imagine the Universe!
    Sep 23, 2020 · Canceling, rearranging, and gathering the unknowns to the left side of the equation gives Kepler's Third Law. Eqn 4: P^2/a^3 = 4 pi^2/ Kepler's ...
  76. [76]
    [PDF] ORBIT MECHANICS ABOUT SMALL ASTEROIDS
    The additional parameter n = √µ/a3 is the mean motion. For each of these equations, there are situations where we wish to pose them in a frame rotating with the ...
  77. [77]
    How Orbital Motion is Calculated - PWG Home - NASA
    Oct 13, 2016 · M = M(0) + 360°(t/T). We assume the period T is known (this requires the 3rd law and is discussed for circular orbits in sections 20 and 20a).Missing: third | Show results with:third
  78. [78]
    [PDF] ht Ma ansit r T - Space Math @ NASA
    Mar 30, 2010 · The synodic period is the time it takes a planet viewed from Earth to be observed at exactly the same illumination phase as it had previously.
  79. [79]
    [PDF] Predicting the Transits of the Stars Kepler-16 A and B
    happen is called the Synodic Period and is calculated using the formula. 360 360 360 t. T. P. -. = In our problem t = 41 days and T = 229 days so P = 50 days ...
  80. [80]
    Radial, transverse and normal satellite position perturbations due to ...
    The perturbations are given in the radial, transverse (or alongtrack) and normal (or cross-track) components. The solution is obtained by projecting the Kepler ...
  81. [81]
    Gauss planetary equations
    Gauss planetary equations. At any given instance in time, a perturbed planetary orbit is completely determined by six osculating orbital elements.
  82. [82]
    Impulsive orbit correction using second-order Gauss's variational ...
    Feb 10, 2020 · Gauss's variational equations (GVEs) provide the instantaneous rates of change of the orbital elements for a small assigned acceleration vector ...
  83. [83]
    [PDF] Spacecraft Dynamics and Control - Lecture 12: Orbital Perturbations
    Long Periodic - Cycles last longer than one orbital period. • Secular - Does not cycle. Disturbances mount over time. Secular Disturbances must be corrected. M.
  84. [84]
    [PDF] 19700014880.pdf - NASA Technical Reports Server (NTRS)
    This completes the development of the variational equations for third-body perturbations only. DRAG r. For the variations dues to the presence od an atmosphere, ...
  85. [85]
    [PDF] Satellite Drag: Aerodynamic Forces in LEO - NASA CCMC
    An Illustra?on of Orbital Decay. Page 16. Energy Dissipa?on Rate (EDR). ​𝜀 =​1 ... Work done by aerodynamic drag along the orbital path l. ​B/2 𝜌​‖​​V ...Missing: formula | Show results with:formula
  86. [86]
    Effect of atmospheric drag on artificial satellite orbits
    The atmospheric drag force causes the major radius and eccentricity of the satellite's orbit to both decay monotonically in time.Missing: da/ | Show results with:da/
  87. [87]
    The origin of the Kirkwood gaps - A mapping for asteroidal motion ...
    SUMMARY AND CONCLUSIONS The mappings in this paper are models for the motion of asteroids near the 3/1 commensurability. They have the correct secular and ...
  88. [88]
    Motions of asteroids at the Kirkwood gaps: I. On the 3:1 resonance ...
    The behavior of asteroids at the 3:1 resonance is studied in a large parameter region, and the various motions that these asteroids show are summarized.Missing: paper | Show results with:paper
  89. [89]
    The instability transition for the restricted 3-body problem
    It is shown that the maximum Lyapunov exponent can be used as an indicator for chaotic behaviour of planetary orbits, which is consistent with previous ...
  90. [90]
    [PDF] Earth orbital lifetime prediction model and program
    An earth orbital satellite lifetime deck has been developed and pro grammed in Fortran IV language for the IBM 7094. The deck represents the.Missing: retention | Show results with:retention
  91. [91]
    [PDF] Lecture 13: The Effect of a Non-Spherical Earth - Matthew M. Peet
    We will need to change to ECI and ultimately RTN coordinates in order to apply the orbit perturbation equations. • This is one of those cases where RTN is not ...
  92. [92]
    [PDF] Observed Tidal Braking in the Earth/Moon/Sun System
    This phase lag in the tidal response produces a torque which causes a transfer of angular momentum in the earth/moon/sun system.
  93. [93]
    Secular Acceleration of the Moon - NASA Eclipse
    Jan 29, 2009 · Conversely, the Moon's gravitational tug on this mass exerts a torque that decelerates the rotation of Earth.
  94. [94]
    Tidal Locking - NASA Science
    The Moon rotates exactly once each time it orbits Earth so the same side of the Moon always faces our planet. This is synchronous rotation.
  95. [95]
    Tidal Friction - an overview | ScienceDirect Topics
    Besides changing the rotation rate of a planet, the tidal torque also changes the obliquity as the tidal bulge is carried outside the orbital plane by rotation ...
  96. [96]
    [PDF] Tidal Forces - Let 'er Rip! - Space Math @ NASA
    Problem 1 - The location of the tidal radius (also called the Roche Limit) for two bodies is given by the formula d = 2.4x ... Roche Limit for a star like our Sun ...
  97. [97]
    [PDF] Constellations of co-orbital planets: horseshoe dynamics ... - arXiv
    Apr 18, 2023 · Satellites Janus and Epimetheus orbit Saturn in a horseshoe co- orbital configuration (Smith et al. 1980; Dermott & Murray 1981a,b). The two ...
  98. [98]
    [PDF] arXiv:1003.0557v2 [astro-ph.EP] 10 Sep 2010
    Sep 10, 2010 · The orbital motion of Janus and Epimetheus presents a peculiar horseshoe-shaped orbit resulting from a. 1:1 orbital resonance (e.g.. Dermott and ...
  99. [99]
    [PDF] Formation of the Janus-Epimetheus system through collisions - arXiv
    Sep 23, 2015 · Libration around. L3, L4, and L5 constitutes a horseshoe orbit (Brown 1911). Many trojan asteroids are in co-orbital trajectories with planets, ...
  100. [100]
    Phoebe - NASA Science
    as well as most objects in the solar ...
  101. [101]
    Orbital Stability of Exomoons and Submoons with Applications to ...
    May 14, 2020 · 2006). Retrograde orbits that are typically associated with the irregular satellites of Jupiter (Jewitt & Haghighipour 2007) also have high ...
  102. [102]
    [PDF] arXiv:1701.02125v1 [astro-ph.EP] 9 Jan 2017
    Jan 9, 2017 · Detectable exomoons around gas giants may be able to form by co-accretion or capture, but determining the upper limit on likely moon masses at ...
  103. [103]
    [PDF] Classifying orbits in the restricted three-body problem - arXiv
    The white domains correspond to the Hill's region, gray shaded domains indicate the forbidden regions, while the thick black lines depict the Zero Velocity ...
  104. [104]
    [PDF] arXiv:2010.05130v1 [math.DS] 11 Oct 2020
    Oct 11, 2020 · The boundaries of the Hill's regions are called zero velocity curves because they are the locus in the configuration space where the kinetic ...
  105. [105]
    [PDF] Geometry of transit orbits in the periodically-perturbed restricted ...
    Mar 31, 2022 · The boundary of the Hill's region, beyond which lies the forbidden realm, is called the zero-velocity surface in the spatial case and zero- ...
  106. [106]
    Zoom & Whirl: Eccentric equatorial orbits around spinning black ...
    Mar 25, 2002 · We study eccentric equatorial orbits of a test-body around a Kerr black hole under the influence of gravitational radiation reaction.
  107. [107]
    [0907.0671] Zoom-Whirl Orbits in Black Hole Binaries - arXiv
    Jul 3, 2009 · Title:Zoom-Whirl Orbits in Black Hole Binaries. Authors:James Healy, Janna Levin, Deirdre Shoemaker. View a PDF of the paper titled Zoom-Whirl ...
  108. [108]
    [PDF] A unified form of lambert's theorem - NASA Technical Reports Server
    In this paper a unified form of. Lambert's theorem will be presented which is valid for elliptic, hyperbolic, and parabolic orbits. The key idea involves the ...
  109. [109]
    [PDF] Lv =v1n-
    To achieve low earth orbit, approximately 7.5kmls delta-v is required in an ideal situation, however launching from the surface of the earth is far from ideal.
  110. [110]
    [PDF] 16.89J / ESD.352J Space Systems Engineering
    May 17, 1999 · using the simplified plane change formula in Equation 29. ∆Vplane change = 2V sin(θ 2). Equation 29. ∆V is the change in velocity required ...
  111. [111]
    [PDF] aas 07-160 comparison of a simple patched conic trajectory code to ...
    The patched conic approximation subdivides the planetary mission into three distinct trajectories and patches them together to create a single trajectory path.Missing: seminal | Show results with:seminal
  112. [112]
    Planetary Voyage - NASA Science
    The Voyager mission was designed to take advantage of a rare geometric arrangement of the outer planets in the late 1970s and the 1980s which allowed for a four ...
  113. [113]
    [PDF] A Researcher's Guide to Space Environmental Effects - NASA
    In terms of materials degradation in space, the low‑Earth orbit (LEO) environment, defined as 200‑1,000 km above Earth's surface, is a particularly harsh ...
  114. [114]
    ISS - NASA's All Sky Fireball Network
    It orbits Earth at an altitude of about 400~km; at this height, atmospheric drag causes orbital decay and the station requires regular reboosts to maintain its ...
  115. [115]
    GNSS - Global Navigation Satellite System - NASA Earthdata
    The current GPS constellation includes 24 satellites, each traveling in a 12-hour, circular orbit 20,200 kilometers above Earth. The satellites are positioned ...
  116. [116]
    [PDF] HANDBOOK FOR LIMITING ORBITAL DEBRIS - NASA Standards
    Jul 30, 2008 · Geosynchronous Earth Orbit (GEO): An orbit with a period equal to the sidereal day. ... The normal altitude of a circular GEO is 35,786 km and the ...
  117. [117]
    What are the Van Allen Belts and why do they matter? - NASA Science
    Jul 23, 2023 · Van Allen calculated that it was possible to fly through the weaker regions of radiation to reach outer space. In 1968, NASA's Apollo Mission 8 ...
  118. [118]
    NASA's Van Allen Probes Survive Extreme Radiation Five Years On
    Sep 1, 2017 · Most satellites, not designed to withstand high levels of particle radiation, wouldn't last a day in the Van Allen Radiation belts.
  119. [119]
    [PDF] An introduction to orbit dynamics and its application to satellite ...
    The basic orbit dynamics of satellite motion are covered in detail. Of particular interest are orbit plane precession,. Sun-synchronous orbits, and ...Missing: astrodynamics | Show results with:astrodynamics
  120. [120]
    Planetary Physical Parameters - JPL Solar System Dynamics
    The following tables contain selected physical characteristics of the planets and dwarf planets, respectively. Table column headings are described below.Missing: synodic | Show results with:synodic
  121. [121]
    2 Mean-Motion Resonance in the Jupiter–Saturn Planetary System
    The motion of the Jupiter–Saturn planetary system near the 5 : 2 mean-motion resonance is modeled analytically in the frame of the planar general three-body ...
  122. [122]
    NASA Exoplanet Archive
    6,042. Confirmed Planets. 10/30/2025 ; 708. TESS Confirmed Planets. 10/30/2025 ; 7,710. TESS Project Candidates. 10/26/2025 ; View more Planet and Candidate ...Planet Counts · Data · 2025 Exoplanet Archive News · Kepler Mission Information
  123. [123]
    Hot Jupiters: Origins, Structure, Atmospheres - AGU Journals - Wiley
    Feb 8, 2021 · The atmospheres of hot Jupiters lie in a unique thermal, chemical, and dynamical regime characterized by strong incident radiation, large ...Abstract · Discovery of Hot Jupiters · Atmospheres · Conclusions and Prospects for...
  124. [124]
    Kepler-16b: In the Light of Two Suns - NASA
    Sep 16, 2011 · This artist's concept illustrates Kepler-16b, the first planet known to definitively orbit two stars – what's called a circumbinary planet.
  125. [125]
    TRAPPIST-1: Dynamical analysis of the transit-timing variations and ...
    TRAPPIST-1 has seven Earth-sized planets in a resonant chain, possibly the longest known, with two- and three-planet resonances.
  126. [126]
  127. [127]
    The Habitable Zone - NASA Science
    The definition of “habitable zone” is the distance from a star at which liquid water could exist on orbiting planets' surfaces.