Fact-checked by Grok 2 weeks ago

Circular orbit

A circular orbit is a trajectory in which a smaller celestial body or artificial satellite revolves around a larger central body, such as a planet or star, at a constant radial distance from its center, resulting in uniform circular motion. This idealized path occurs when the gravitational force between the bodies exactly balances the centripetal force required for the orbiting object's velocity, ensuring the radius remains fixed throughout the orbit. In , circular orbits represent a special case of the elliptical paths described by Kepler's of planetary motion, characterized by an of exactly zero, which eliminates any variation in distance from the central body. The orbital velocity v for such an orbit is derived from the equilibrium condition \frac{GM m}{r^2} = \frac{m v^2}{r}, yielding v = \sqrt{\frac{GM}{r}}, where G is the , M is the mass of the central body, m is the mass of the orbiting body, and r is the orbital radius; this velocity decreases with increasing altitude, as seen in low Earth orbits () at approximately 7.8 km/s versus geostationary orbits () at about 3.0 km/s. The total mechanical energy of an object in a circular orbit is negative, given by E = -\frac{GM m}{2r}, indicating a bound, configuration under Newtonian . Circular orbits are fundamental to spaceflight applications, enabling consistent communication, Earth observation, and navigation; for instance, GEO satellites at an altitude of approximately 36,000 km above Earth's equator maintain a period matching Earth's sidereal rotation (23 hours 56 minutes 4 seconds), appearing stationary relative to a ground point. The first artificial near-circular orbit was achieved by Sputnik 1 in 1957. Achieving and sustaining these orbits requires precise velocity adjustments during launch and periodic station-keeping maneuvers to counteract perturbations from atmospheric drag, gravitational anomalies, and solar radiation pressure. While natural planetary orbits are rarely perfectly circular due to eccentricities introduced by formation processes and perturbations, circular orbits are preferred for most human-made satellites to simplify mission design and operations.

Fundamentals

Definition

A circular orbit is a trajectory in which a smaller body, such as a satellite or planet, moves around a larger central body, like a star or planet, along a path that forms a perfect circle, maintaining a constant radial distance from the center of the orbit. In this configuration, the orbiting body travels at a uniform speed, with the gravitational attraction between the two bodies providing the force required to sustain the motion without deviation from the circular path. This idealized motion arises in the framework of classical mechanics, where the orbit is stable and periodic. The concept of circular orbits has historical roots in early astronomical models, but it gained precise formulation through in the early , which established that planetary paths are ellipses with at one focus; circular orbits represent the special limiting case of these ellipses where the eccentricity is exactly zero. Prior to Kepler, prevailing astronomical theories, including those of and Copernicus, assumed all celestial orbits were circular due to the philosophical ideal of . Kepler's work, derived from meticulous observations by , shifted this view by accommodating elliptical paths while preserving the circular case as a theoretical possibility. The theoretical foundation for circular orbits rests on the in Newtonian gravity, which considers the interaction between two point masses under the of universal gravitation, with no external forces or third-body perturbations. In this setup, the motion is analyzed relative to the center of mass, allowing the reduced-mass system to describe the relative orbit as if one body is fixed. This approximation is fundamental to and enables the prediction of orbital behavior for systems like Earth-Moon or planetary-satellite pairs.

Assumptions and Idealizations

The modeling of circular orbits relies on several core assumptions rooted in Newtonian mechanics to simplify the complex dynamics of gravitational interactions. The central body is treated as a point , assuming spherical and uniform , which allows the gravitational to be represented by a central acting along the line connecting the centers of . The of the orbiting object is considered negligible compared to that of the central body, enabling the use of the approximation where the orbiting body's motion is analyzed relative to a fixed central . Additionally, the gravitational follows Newton's , with no other forces such as atmospheric drag, tidal effects, or third-body perturbations influencing the orbit. These assumptions facilitate key idealizations that reduce the problem to a tractable form. The system is idealized as a , where the mutual orbit around the common is equivalent to the motion of a single reduced-mass body in a central potential, often further simplified by fixing the more massive body. Motion is confined to a single plane due to the conservation of , assuming no out-of-plane components. The gravitational parameter μ, defined as the product of the gravitational constant and the central body's mass M (μ = M), is held constant, reflecting unchanging masses and a static gravitational environment. While these simplifications enable analytical solutions within the Newtonian framework, they impose limitations on the model's applicability. The assumptions break down for extended or non-spherical bodies, where oblateness introduces perturbations like , requiring more advanced models. Similarly, at relativistic speeds or in strong gravitational fields, such as near black holes, general relativity must account for deviations from Newtonian predictions, though classical approximations suffice for most and solar system orbits.

Kinematics

Orbital Velocity

In a circular orbit, the orbital represents the constant tangential speed at which an object must travel to maintain a around a central massive body, such as a or , under the influence of alone. This ensures that the object's motion perpetually counters the inward gravitational , resulting in a uniform circular without any component. The derivation of orbital begins with the principle of balancing the gravitational force pulling the orbiting body toward the center against the required to sustain . Qualitatively, the gravitational pull provides the necessary inward , which must equal the centripetal for the orbit to remain circular; a full quantitative force balance is detailed in subsequent analyses. From this , the orbital v is given by v = \sqrt{\frac{GM}{r}} where G is the , M is the mass of the central body, and r is the orbital radius. This formula yields typical velocities on the order of kilometers per second for common orbits. For instance, satellites in , at an altitude of approximately 200–2,000 km above Earth's surface, require a speed of about 7.8 km/s to remain in circular orbit. Similarly, Earth's own circular orbit around the Sun, with a radius of about 1 , occurs at an average velocity of 29.8 km/s. These values highlight how orbital velocity decreases with increasing radius, as closer orbits demand higher speeds to balance the stronger .

Centripetal Acceleration

In a circular orbit, the centripetal acceleration is the radial component of the object's acceleration that continuously redirects its toward the center of the orbit, maintaining the constant radius despite the tangential motion. This acceleration arises solely from the change in direction of the velocity, as the speed remains constant in uniform . Geometrically, the centripetal acceleration is always perpendicular to the instantaneous , pointing inward along the , and its is for a given orbital and speed. For an object of mass m in with orbital v at r, the of this acceleration is given by a_c = \frac{v^2}{r}. In the context of a gravitational circular orbit around a central of M, this centripetal acceleration equals the at distance r, yielding a_c = \frac{GM}{r^2}, where G is the , thereby linking the kinematic requirement to the strength. This equivalence, \frac{v^2}{r} = \frac{GM}{r^2}, highlights how the orbital velocity v is determined by the balance of these terms. The g(r) = \frac{GM}{r^2} thus serves as the effective providing the necessary centripetal acceleration for the , decreasing inversely with the square of the radial distance.

Speed and

In a circular , the angular speed \omega represents the rate of change of the angular position of the orbiting body with respect to the central body. For a of negligible orbiting a central M at radius r, the angular speed is given by \omega = \sqrt{GM / r^3}, where G is the . This formula arises from the relation \omega = v / r, where v is the orbital , combined with the expression for v derived from gravitational balance. The T, the time required to complete one full revolution around the central body, is the inverse of the speed scaled by $2\pi, yielding T = 2\pi / \omega = 2\pi \sqrt{r^3 / GM}. This period formula can be derived by dividing the orbital $2\pi r by the tangential v = \sqrt{GM / r}, providing a direct measure of the rotational timescale for the . For circular orbits, the satisfies Kepler's third law in its Newtonian form, where [T^2](/page/T+2) \propto r^3, or more precisely [T^2](/page/T+2) = (4\pi^2 / [GM](/page/GM)) r^3. This proportionality holds specifically for the case where the orbiting body's mass is much smaller than the central mass, linking the period directly to the orbital radius without dependence on the orbiting mass. A practical application of these relations is the around , where the period T matches Earth's sidereal rotation period of approximately 23 hours 56 minutes (often rounded to 24 hours for simplicity), resulting in an orbital radius of about 42,000 km from Earth's center. This configuration allows satellites to remain fixed over a point on the , enabling continuous communication coverage.

Dynamics

Equation of Motion

In the two-body problem under Newtonian gravity, the relative motion of two point masses m_1 and m_2 can be described using polar coordinates (r, \theta) in the orbital plane, where r is the radial separation and \theta is the angular coordinate. The system reduces to an equivalent one-body problem with reduced mass \mu = \frac{m_1 m_2}{m_1 + m_2} subject to a central force F = -\frac{G m_1 m_2}{r^2} = -\frac{G M \mu}{r^2}, where M = m_1 + m_2 is the total mass and G is the gravitational constant. The radial component of the equation of motion, derived from Newton's second law in polar coordinates, is given by \ddot{r} - r \dot{\theta}^2 = -\frac{G M}{r^2}, where the term r \dot{\theta}^2 represents the centrifugal acceleration outward, balancing the inward gravitational acceleration. This equation governs the radial dynamics, with angular momentum conservation ensuring \dot{\theta} = \frac{L}{\mu r^2}, where L is the constant angular momentum. For a circular orbit, the radius r remains constant, implying \dot{r} = 0 and \ddot{r} = 0. Substituting these into the radial equation yields the balance condition r \dot{\theta}^2 = \frac{G M}{r^2}, or equivalently, \dot{\theta} = \omega = \sqrt{\frac{G M}{r^3}}, where \omega is the constant angular speed. The angular motion is then uniform, described by \theta(t) = \omega t + \phi_0, with \phi_0 an initial phase. This confirms that r(t) = constant satisfies the equation under the assumptions of a central inverse-square force and conserved angular momentum, resulting in stable circular motion.

Force Balance

In a circular orbit, assuming the orbiting body has much smaller mass than the central body (m \ll M), the gravitational force acting on the orbiting body provides the exact required to maintain the constant radius of the path; this approximates the general two-body relative motion where the gravitational parameter is GM with M \approx m_1 + m_2. The gravitational force F_g between the central body of mass M and the orbiting body of mass m at a r is given by : F_g = \frac{G M m}{r^2}, directed radially inward toward the center of the central body, where G is the . For uniform circular motion, this gravitational force must equal the F_c necessary to keep the orbiting body moving in a : F_c = \frac{m v^2}{r}, where v is the . This requirement arises from the radial component of Newton's second law, ensuring the toward the center matches the kinematic demands of . Setting F_g = F_c yields the force balance equation: \frac{G M m}{r^2} = \frac{m v^2}{r}. The masses m cancel out, simplifying to \frac{G M}{r^2} = \frac{v^2}{r}, which highlights that the depends on the interplay between the central M, the orbital radius r, and the speed v, without reliance on the orbiting body's . This equality ensures no net radial force beyond what sustains the circular path, meaning the gravitational pull precisely counters the tendency for the body to move in a straight line, resulting in stable orbital motion at fixed r. The magnitude of these forces scales inversely with r^2 for and as $1/r for the centripetal requirement at fixed speed, but in , the orbital radius r directly determines the force magnitude needed for balance, as larger orbits weaken the gravitational pull while requiring less for the same speed.

Energy Considerations

Total Mechanical Energy

In a circular orbit, the kinetic energy K of an orbiting body of mass m at radius r from a central body of mass M is K = \frac{1}{2} m v^2, where v is the constant . Substituting the expression for orbital speed yields K = \frac{G M m}{2 r}, with G denoting the ./Book%3A_University_Physics_I_-Mechanics_Sound_Oscillations_and_Waves(OpenStax)/13%3A_Gravitation/13.05%3A_Satellite_Orbits_and_Energy) The gravitational potential energy U for this two-body system is U = -\frac{G M m}{r}, adopting the convention where potential is zero at infinite separation./Book%3A_University_Physics_I_-Mechanics_Sound_Oscillations_and_Waves(OpenStax)/13%3A_Gravitation/13.05%3A_Satellite_Orbits_and_Energy) The total E is thus the sum E = K + U = -\frac{G M m}{2 r}. This value remains constant throughout the due to the conservative nature of the gravitational force and is negative, signifying a where the orbiting body cannot escape to without additional energy input./Book%3A_University_Physics_I_-Mechanics_Sound_Oscillations_and_Waves(OpenStax)/13%3A_Gravitation/13.05%3A_Satellite_Orbits_and_Energy) The provides insight into this energy partition for stable, self-gravitating systems under an , stating that the time-averaged satisfies $2 \langle K \rangle = -\langle \mathbf{r} \cdot \mathbf{F} \rangle. For a circular orbit, this simplifies to |U| = 2 K, confirming the equipartition where equals half the magnitude of ./03%3A_Systems_of_Particles/3.13%3A_The_Virial_Theorem) In , the v_{\rm esc} corresponds to zero total energy (E = 0) and is given by v_{\rm esc} = \sqrt{\frac{2 G M}{r}} = \sqrt{2} \, v, exceeding the circular orbital speed by a factor of \sqrt{2} and thus requiring twice the to reach unbound conditions./Book%3A_University_Physics_I_-Mechanics_Sound_Oscillations_and_Waves(OpenStax)/13%3A_Gravitation/13.05%3A_Satellite_Orbits_and_Energy)

Delta-v for Achieving Circular Orbit

Delta-v, denoted as \Delta v, represents the impulse-induced change in a spacecraft's velocity, typically achieved through , and is a fundamental metric in astrodynamics for quantifying the propellant requirements of orbital maneuvers. Achieving a circular orbit from a surface launch involves imparting sufficient \Delta v to reach the orbital while overcoming gravitational and atmospheric losses. In an idealized scenario without losses, the minimum \Delta v approximates the circular orbital \sqrt{GM/r}, where G is the gravitational constant, M is the central body's mass, and r is the orbital radius; however, real launches require additional \Delta v for gravity losses during ascent and drag. For transferring between two circular orbits, the Hohmann transfer provides a minimum-energy elliptical path requiring two impulsive burns. The first \Delta v at the initial radius r_1 is \Delta v_1 = \sqrt{\frac{GM}{r_1}} \left( \sqrt{\frac{2 r_2}{r_1 + r_2}} - 1 \right), which boosts the to enter the transfer ellipse. The second \Delta v at the target radius r_2 is \Delta v_2 = \sqrt{\frac{GM}{r_2}} \left( 1 - \sqrt{\frac{2 r_1}{r_1 + r_2}} \right), circularizing the by adjusting to the local circular . These burns occur tangentially at perigee and apogee of the transfer to minimize total \Delta v. In more general scenarios, such as transitioning from an elliptical orbit to a circular one, \Delta v is the vector difference between the instantaneous velocity and the required circular velocity at the burn point, with minimum-energy paths favoring burns at perigee (to raise apogee) or apogee (to raise perigee). For instance, circularization from an elliptical orbit at its apogee involves \Delta v = v_c - v_a, where v_c is the circular velocity and v_a the apogee velocity, ensuring efficient energy addition for bound orbits. Radial velocity components, if present, require additional \Delta v to align the trajectory tangentially before circularization. A practical example is inserting a into () at approximately 200-400 km altitude, where the ideal orbital velocity is about 7.8 km/s, but the total from ground launch reaches around 9.5 km/s to account for ascent inefficiencies. This budget includes roughly 1.5-2 km/s for and losses, highlighting the gap between theoretical and operational requirements in real missions.

Relativistic Corrections

Orbital Velocity in General Relativity

In the Newtonian framework, the orbital velocity for a circular orbit around a central mass M at radius r is given by v = \sqrt{GM/r}, derived from balancing gravitational attraction with . modifies this expression through the curvature of , particularly in strong fields. In the post-Newtonian approximation for weak fields, the incorporates relativistic terms from the expanded , yielding a corrected orbital velocity v \approx \sqrt{GM/r} \left(1 - \frac{3}{2} \frac{GM}{c^2 r}\right), where c is the . This first-order correction arises from the 1PN terms in the , reducing the velocity slightly compared to the Newtonian value for a given r. Within the , which describes the around a non-rotating, spherically symmetric , circular geodesics exist but stability is limited by the (ISCO). Stable circular orbits are possible only for r > 3 r_s, where r_s = 2GM/c^2 is the ; orbits between $1.5 r_s and $3 r_s are unstable, and none exist closer than the at $1.5 r_s. This restriction stems from the shape of the , where the at the ISCO marks the boundary between stable and unstable motion. These relativistic modifications have practical implications across scales. For GPS satellites orbiting at altitudes of approximately 20,000 km, where GM/(c^2 r) \sim 10^{-10}, the corrections to orbital velocity and related clock rates are small but essential for sub-nanosecond timing accuracy, incorporated via post-Newtonian adjustments in satellite ephemerides. In contrast, black hole accretion disks, where orbital radii can approach a few times the ISCO (e.g., r \sim 6GM/c^2), experience substantial relativistic effects; velocities deviate significantly from Keplerian values, influencing disk structure, radiation spectra, and angular momentum transport.

Derivation of Relativistic Effects

In , the motion of particles in the of a spheroidal, non-rotating is described by geodesics in the , which is the exact solution to Einstein's field equations for such a source. The in standard coordinates (t, r, \theta, \phi) is given by ds^2 = -\left(1 - \frac{2GM}{c^2 r}\right) c^2 dt^2 + \left(1 - \frac{2GM}{c^2 r}\right)^{-1} dr^2 + r^2 d\theta^2 + r^2 \sin^2\theta d\phi^2, where G is the , M is the , and c is the . For timelike geodesics corresponding to massive particles in equatorial motion (\theta = \pi/2, d\theta = 0), the \tau satisfies ds^2 = -c^2 d\tau^2. The metric's symmetries yield two conserved quantities: the \tilde{E} = (1 - 2GM/(c^2 r)) c (dt/d\tau) (energy per unit rest mass at infinity) and the L = r^2 (d\phi/d\tau) (angular momentum per unit rest mass). Substituting these into the condition g_{\mu\nu} (dx^\mu/d\tau) (dx^\nu/d\tau) = -c^2 leads to the radial equation \left(\frac{dr}{d\tau}\right)^2 = \tilde{E}^2 - V_\text{eff}(r), where the effective potential is V_\text{eff}(r) = \left(1 - \frac{2GM}{c^2 r}\right) \left(c^2 + \frac{L^2}{r^2}\right). This form arises directly from rearranging the geodesic equation using the conserved quantities. To connect to Newtonian mechanics and derive corrections, the effective potential is expanded in the weak-field limit ($2GM/(c^2 r) \ll 1), yielding a form analogous to the Newtonian orbital energy: V_\text{eff}(r) \approx - \frac{GM}{r} + \frac{L^2}{2 r^2} - \frac{GM L^2}{c^2 r^3}, where the first two terms recover the classical gravitational and centrifugal potentials, and the third term is the leading general relativistic correction. This expansion is obtained by Taylor series in the small parameter GM/(c^2 r). For circular orbits, the radius r is constant, so dr/d\tau = 0 and dV_\text{eff}/dr = 0. Differentiating the approximate effective potential gives \frac{dV_\text{eff}}{dr} = \frac{GM}{r^2} - \frac{L^2}{r^3} + \frac{3 GM L^2}{c^2 r^4} = 0. Solving for L^2 yields L^2 = \frac{GM r}{1 - 3 GM/(c^2 r)}, which reduces to the Newtonian L^2 = GM r in the limit GM/(c^2 r) \to 0. Substituting back, the coordinate angular speed \omega = d\phi/dt = (d\phi/d\tau)/(dt/d\tau) = (L/r^2) / (\tilde{E}/(c (1 - 2GM/(c^2 r)))) for circular orbits simplifies to the Keplerian form with a relativistic correction: \omega^2 = \frac{GM}{r^3} \left(1 - \frac{3 GM}{c^2 r}\right). This follows from expressing \tilde{E} at equilibrium and the force balance in the post-Newtonian approximation. The coordinate orbital velocity is then v_\text{coord} = \omega r \approx \sqrt{GM/r} \left[1 - \frac{3}{2} \frac{GM}{c^2 r}\right], obtained by taking the square root and expanding to first order in GM/(c^2 r). Stability of circular orbits requires d^2 V_\text{eff}/dr^2 > 0. Setting the second derivative to zero determines the boundary between stable and unstable orbits, yielding the innermost stable circular orbit (ISCO) at r = 6 GM / c^2, where the correction term causes the effective potential minimum to disappear for smaller radii. Below this radius, no stable circular orbits exist, as perturbations lead to infall.

References

  1. [1]
    Chapter 5: Planetary Orbits - NASA Science
    ... orbit's ellipse (described in Chapter 3. Recall an eccentricity of zero indicates a circular orbit). Inclination is the angular distance of the orbital ...<|control11|><|separator|>
  2. [2]
    Astronautics, Space & Astrodynamics – Introduction to Aerospace ...
    Circular orbits occur when the gravitational force acting on a satellite balances the centrifugal force required to maintain uniform circular motion, as ...
  3. [3]
    [PDF] The Two-Body Problem - UCSB Physics
    the radial coordinate will be constant for all time. This of course corresponds to a circular orbit. Thus, we see that so long as the effective potential admits.
  4. [4]
    Kepler's Laws - MacTutor History of Mathematics
    Observational background​​ Kepler originally investigated the orbit of Mars because that was the task allocated to him by Tycho Brahe (1546-1601), when Kepler ...
  5. [5]
    [PDF] 2 - The Two Body Problem
    The two-body problem (two point masses interacting via gravity, with no other forces present) is the fundamental building block of celestial mechanics.<|control11|><|separator|>
  6. [6]
    3.5 Orbital Mechanics – A Guide to CubeSat Mission and Bus Design
    To treat the spacecraft motion as a two-body problem, we have to make the following four important assumptions: The mass of the satellite (m2) is negligible ...
  7. [7]
    [PDF] Second–Order Relative Motion Equations - VTechWorks
    2.1 Orbital Mechanics. 2.1.1 The Two-Body Problem. The basic assumptions of the Two-Body problem are [25]:. 1. The equations of motion are expressed in a non ...
  8. [8]
    [PDF] Chapter 3 Two Body Central Forces - Rutgers Physics
    Consider two particles of masses m1 and m2, with the only forces those of their mutual interaction, which we assume is given by a potential which is a.
  9. [9]
    One Universe: motion knowledge concept 15
    In equations: a = GM/r2 = v2/r Cancelling like terms in these ratios gives GM/r = v2 Taking the square root gives the equation v = sqrt(GM/r) This is the ...
  10. [10]
    AST 101: Forces, Orbits, and Energy
    V = (GM/d)1/2 . The circumference of a circular orbit with a radius d is 2πd. The velocity (or more correctly, the speed) of an object is the distance ...<|control11|><|separator|>
  11. [11]
    ESA - Low Earth orbit - European Space Agency
    Mar 2, 2020 · Satellites in this orbit travel at a speed of around 7.8 km per second; at this speed, a satellite takes approximately 90 minutes to circle ...
  12. [12]
    Solar System Data - HyperPhysics
    Earth Data ; Minimum distance from Sun: 0.983 AU=1.471x108 km ; Mean orbital velocity: 29.8 km/s ; Siderial period: 365.256 days ; Rotation period: 23.9345 hours.
  13. [13]
    Uniform Circular Motion - Imagine the Universe! - NASA
    Sep 23, 2020 · Centripetal forces are always directed toward the center of the circular path. By definition, acceleration is the rate of change in velocity of ...Missing: formula | Show results with:formula<|control11|><|separator|>
  14. [14]
    6.2 Centripetal Acceleration – College Physics chapters 1-17
    We can express the magnitude of centripetal acceleration using either of two equations: a c = v 2 r ; a c = r ω 2 .
  15. [15]
    Uniform circular motion
    The distance the moon travels in one orbital period T is d = 2πr. Its speed is v = distance/time = d/T. The centripetal acceleration of the moon is v2/r. ...
  16. [16]
    [PDF] Lecture 16: Centripetal Acceleration, ac = v 2/r
    Circular motion at constant speed, v leads to ac = v2/r. It may surprise you to learn that when a mass moves around a circle of radius r at constant speed, ...
  17. [17]
    Central Forces (Orbits, Scattering, etc) - UMD Physics
    ... G M m / r . If we consider the centripetal acceleration, held into a circular orbit by a central gravitational force, we can write that. mv2r=GMmr ...<|control11|><|separator|>
  18. [18]
    Earth Orbits - HyperPhysics
    Above the earth's surface at a height of h = m = x 106 m, which corresponds to a radius r = x earth radius, the acceleration of gravity is g = m/s2 = x g on the ...
  19. [19]
    [PDF] CHAPTER 11 SATELLITE ORBITS
    and using u = 1/r this becomes [d2u/dθ2 + u] = GMm2/L2 . The solution has the form of the equation for a conic section 1/r = [GMm2/L2] [1 + ε cosθ] where the ...Missing: formula | Show results with:formula
  20. [20]
    Kepler's Third Law - Physics
    Let's derive it for a circular orbit, assuming a mass m is orbiting a mass M, with r being the radius of the orbit. ΣF = ma. GmM/r2 = mv2/r. Cancelling the m's ...
  21. [21]
    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.
  22. [22]
    ESA - Orbits - European Space Agency
    GEO is a circular orbit 35 786 kilometres above Earth's equator and follows the direction of Earth's rotation. An object in GEO has an orbital period equal to ...
  23. [23]
    [PDF] Orbits 1
    The Two-Body Problem. Consider a pair of particles1with masses M1 and M2 ... equation of motion to polar coordinates. µ. ( r − r ˙θ. 2). = −. GM1M2 r2. (4).<|control11|><|separator|>
  24. [24]
    [PDF] Chapter 9 Circular Motion Dynamics - MIT OpenCourseWare
    According to Newton's Second Law, there must be a centripetal force acting on the. Moon directed towards the center of the Earth that accounts for this inward ...
  25. [25]
    Newtonian Gravitation | ASTRO 801: Planets, Stars, Galaxies, and ...
    Newtonian gravitation states the force of gravity between two masses is F = G (m1 x m2) / d2, where G is a constant and d is distance.
  26. [26]
    [PDF] Lecture D30 - Orbit Transfers - DSpace@MIT
    A Hohmann Transfer is a two-impulse elliptical transfer between two co-planar circular orbits. The transfer itself consists of an elliptical orbit with a ...
  27. [27]
    [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.
  28. [28]
    [PDF] On-Orbit Cryogenic Refueling - OhioLINK ETD Center
    LEO – Low Earth Orbit. LLO – Low Lunar Orbit. LV – Launch Vehicle m ... 9.5 km/s of delta-V at initial launch. Since the mission delta-V from LEO in ...
  29. [29]
    The Post-Newtonian Approximation for Relativistic Compact Binaries
    We show a method to derive post-Newtonian equations of motion for relativistic compact binaries based on a surface integral approach and the strong field point ...
  30. [30]
    Innermost stable circular orbits of spinning test particles in ... - arXiv
    Mar 24, 2015 · We consider the motion of classical spinning test particles in Schwarzschild and Kerr metrics and investigate innermost stable circular orbits (ISCO).Missing: formula | Show results with:formula
  31. [31]
    [PDF] arXiv:gr-qc/0507121v3 13 Dec 2006
    Dec 13, 2006 · In general, the satellites of the GNSS are affected by Relativity in three different ways: in the equations of motion, in the signal propagation ...
  32. [32]
    Foundations of Black Hole Accretion Disk Theory - PMC
    This review covers the main aspects of black hole accretion disk theory. We begin with the view that one of the main goals of the theory is to better ...
  33. [33]
    7. The Schwarzschild Solution and Black Holes
    Circular orbits will be stable if they correspond to a minimum of the potential, and unstable if they correspond to a maximum. Bound orbits which are not ...
  34. [34]
    Orbits in the Schwarzschild Metric
    Orbits in the Schwarzschild Metric. Start with the dot product of the 4-momentum with itself and use the conserved quantities above to eliminate ...
  35. [35]
    [PDF] General Relativity Fall 2019 Lecture 20: Geodesics of Schwarzschild
    Nov 7, 2019 · From classical mechanics, we know that circular orbits where d2Veff/dr2 > 0 are stable, whereas those with. d2Veff/dr2 < 0 are unstable. This ...