Fact-checked by Grok 2 weeks ago

Specific angular momentum

Specific angular momentum, denoted as \vec{h}, is a fundamental quantity in and astrodynamics, representing the of a per and defined as the of its \vec{r} and \vec{v}, yielding \vec{h} = \vec{r} \times \vec{v}. With SI s of square meters per second (m²/s), it quantifies the rotational motion intrinsic to the body's relative to a reference point, such as a central gravitational . This quantity is particularly crucial in orbital contexts, where its magnitude h = r v \sin \phi (with \phi the angle between \vec{r} and \vec{v}) determines key orbital characteristics without dependence on the body's total . In the , specific angular momentum is conserved due to the central nature of the gravitational force, which exerts no on the orbiting body, ensuring \frac{d\vec{h}}{dt} = 0. The vector \vec{h} is always perpendicular to the of motion defined by \vec{r} and \vec{v}, thereby fixing the and orientation in space. This conservation property underpins Kepler's second of planetary motion, where the \frac{dA}{dt} = \frac{h}{2} remains constant, meaning equal areas are swept by the position vector in equal times. Perturbations, such as atmospheric or non-spherical gravitational fields, can alter \vec{h}, but in ideal central-force scenarios, it remains invariant throughout the orbit. Specific angular momentum plays a pivotal role in characterizing orbital shapes and parameters within conic section trajectories, including ellipses, parabolas, and hyperbolas. For elliptical s, h relates directly to the semi-major axis a and e via the and considerations, enabling predictions of perigee and apogee distances. Its magnitude influences the flight path angle and components, which adjust dynamically to maintain constancy as the progresses. In broader astrophysical contexts, such as planetary formation or systems, variations in specific angular momentum help model disk evolution and transfer processes. Applications of specific angular momentum extend to mission design, where it is used to compute orbits, such as Hohmann transfers, by matching \vec{h} at points like perigee. In operations, changes in h detects perturbations, aiding in and collision avoidance. Furthermore, in cometary and dynamics, low specific angular momentum values indicate capture origins, while higher values suggest hyperbolic escapes from solar systems.

Basic Concepts

Definition

Specific angular momentum, denoted as \mathbf{h}, is defined as the angular momentum vector \mathbf{L} of a body divided by its mass m, such that \mathbf{h} = \frac{\mathbf{L}}{m}. This quantity represents the angular momentum on a per-unit-mass basis, commonly used in classical mechanics to describe the rotational dynamics of particles or extended bodies relative to a reference point. Physically, specific angular momentum quantifies the degree of rotational motion per unit mass about the chosen origin, capturing how "far" and "fast" the motion is from pure radial . In isolated systems subject to no external torques, \mathbf{h} is , mirroring the of total but independent of the system's total mass. As a , \mathbf{h} points to the plane containing the and vectors, thereby defining the plane of motion for orbital or rotational trajectories. In the (SI), specific angular momentum has dimensions of length squared per time, or [L² T⁻¹], corresponding to square meters per second (m²/s). Unlike total , which scales with mass and has units of square meters per second (kg·m²/s), the specific form is mass-invariant, making it particularly advantageous for analyzing systems with variable mass—such as rockets—or for comparing motions on a per-particle basis without scaling effects from differing masses.

Mathematical Formulation

The specific angular momentum \mathbf{h} is a vector quantity defined as the cross product of the position vector \mathbf{r} from a chosen reference point (typically the center of mass or the primary body) and the velocity vector \mathbf{v} of the body, given by \mathbf{h} = \mathbf{r} \times \mathbf{v}. This formulation arises in for point masses and is particularly relevant in orbital contexts where the reference point is fixed.\] The vector $\mathbf{h}$ is [perpendicular](/page/Perpendicular) to both $\mathbf{r}$ and $\mathbf{v}$, thus lying normal to the plane of motion.\[ The magnitude of the specific angular momentum is h = |\mathbf{h}| = r v \sin \phi, where r = |\mathbf{r}|, v = |\mathbf{v}|, and \phi is the angle between \mathbf{r} and \mathbf{v}; this follows directly from the geometric definition of the cross product magnitude.\] In polar coordinates for planar motion, where the position is described by radial distance $ r $ and angular position $ \theta $, the magnitude simplifies to $ h = r^2 \dot{\theta} $, with $ \dot{\theta} = d\theta/dt $ representing the angular speed.\[ The Cartesian components of \mathbf{h} are expressed as h_x = y v_z - z v_y, \quad h_y = z v_x - x v_z, \quad h_z = x v_y - y v_x, where (x, y, z) and (v_x, v_y, v_z) are the components of \mathbf{r} and \mathbf{v}, respectively.$$] In the context of Keplerian orbits, the magnitude h relates to the orbital parameters through h = \sqrt{\mu p}, where \mu = G(M + m) is the standard gravitational parameter (with G as the gravitational constant, M the mass of the primary body, and m the mass of the orbiting body) and p is the semi-latus rectum of the conic section orbit.[$$

Conservation Laws

In Central Force Fields

In central force fields, the force acting on a particle is directed along the position vector \mathbf{r} from the force center and depends only on the magnitude r = |\mathbf{r}|, expressed as \mathbf{F} = f(r) \hat{\mathbf{r}}, where \hat{\mathbf{r}} = \mathbf{r}/r and f(r) is a scalar function. The torque \boldsymbol{\tau} = \mathbf{r} \times \mathbf{F} vanishes because \mathbf{F} is parallel to \mathbf{r}, yielding \boldsymbol{\tau} = 0. This implies that the time derivative of the angular momentum \mathbf{L} = m \mathbf{r} \times \mathbf{v} is zero, \frac{d\mathbf{L}}{dt} = \boldsymbol{\tau} = 0, so \mathbf{L} is conserved; consequently, the specific angular momentum \mathbf{h} = \mathbf{r} \times \mathbf{v} (per unit mass) is also conserved, with \frac{d\mathbf{h}}{dt} = 0. A general proof follows from the definition: \frac{d\mathbf{h}}{dt} = \frac{d}{dt} (\mathbf{r} \times \mathbf{v}) = \mathbf{v} \times \mathbf{v} + \mathbf{r} \times \mathbf{a} = \mathbf{r} \times \mathbf{a}, where \mathbf{a} = \mathbf{v}/dt = \mathbf{F}/m is parallel to \mathbf{r} for a central force, making \mathbf{r} \times \mathbf{a} = 0. Thus, \mathbf{h} remains constant in magnitude and direction. Conservation of \mathbf{h} confines the particle's motion to a perpendicular to \mathbf{h}, as any out-of-plane component would vary. Additionally, the is constant: the rate at which area is swept by the position vector is dA/dt = |\mathbf{h}|/2, leading to equal areas in equal times for any central force. Examples include the gravitational force, \mathbf{F} = -GMm/r^2 \hat{\mathbf{r}} (an ), and more generally, inverse-power-law forces \mathbf{F} \propto r^{-n} \hat{\mathbf{r}} (for n > 0), where the central ensures holds. The isotropic harmonic oscillator, \mathbf{F} = -k \mathbf{r}, also exemplifies this for linear restoring forces.

Proof in the Two-Body Problem

In the two-body problem, two point masses m_1 and m_2 interact solely through their mutual gravitational attraction, with no external forces acting on the system. The dynamics can be reduced to an equivalent one-body problem by introducing the relative position vector \mathbf{r} = \mathbf{r}_1 - \mathbf{r}_2, where \mathbf{r}_1 and \mathbf{r}_2 are the position vectors of the two masses relative to an inertial frame. The reduced mass is defined as \mu = \frac{m_1 m_2}{m_1 + m_2}, which governs the motion of this effective single body in the relative coordinate system. The gravitational force between the masses is central and directed along the line joining them, given by \mathbf{F} = -\frac{[G](/page/G) m_1 m_2}{r^2} \hat{\mathbf{r}} = -\frac{[G](/page/G) m_1 m_2 \mathbf{r}}{r^3}, where [G](/page/G) is the , r = |\mathbf{r}|, and \hat{\mathbf{r}} = \mathbf{r}/r. This force law ensures that the acceleration of the relative vector is \ddot{\mathbf{r}} = -\frac{[G](/page/G) (m_1 + m_2)}{r^3} \mathbf{r}, as derived from Newton's second law applied to the system: \mu \ddot{\mathbf{r}} = \mathbf{F}. The specific angular momentum \mathbf{h} for the relative motion is defined as \mathbf{h} = \mathbf{r} \times \mathbf{v}, where \mathbf{v} = \dot{\mathbf{r}} is the relative velocity; this quantity has units of area per unit time and is independent of the reduced mass. To prove its constancy, compute the time derivative: \frac{d\mathbf{h}}{dt} = \frac{d}{dt} (\mathbf{r} \times \mathbf{v}) = \dot{\mathbf{r}} \times \mathbf{v} + \mathbf{r} \times \dot{\mathbf{v}} = \mathbf{v} \times \mathbf{v} + \mathbf{r} \times \mathbf{a} = \mathbf{r} \times \mathbf{a}, where \mathbf{a} = \ddot{\mathbf{r}} is the relative acceleration. Substituting the gravitational acceleration yields \mathbf{r} \times \mathbf{a} = \mathbf{r} \times \left( -\frac{G (m_1 + m_2)}{r^3} \mathbf{r} \right) = -\frac{G (m_1 + m_2)}{r^3} (\mathbf{r} \times \mathbf{r}) = \mathbf{0}, since the cross product of any vector with itself is zero. Thus, \frac{d\mathbf{h}}{dt} = \mathbf{0}, confirming that \mathbf{h} is constant in both magnitude and direction. This proof holds under the assumptions of an inertial reference frame, absence of external forces or torques on the system, and treatment of the masses as point particles (or spherically symmetric bodies, ensuring the gravitational force remains central). For an orbiting body where one mass dominates (e.g., m_2 \ll m_1), the specific angular momentum approximates \mathbf{h} \approx \mathbf{r}_2 \times \mathbf{v}_2 relative to the , retaining its constant magnitude h = r v_\perp (with v_\perp the perpendicular component) and perpendicular to the .

Implications for Keplerian Orbits

Second Law: Equal Areas

Kepler's second law, also known as the law of equal areas, states that a joining a to sweeps out equal areas during equal intervals of time. This empirical observation, derived by from Brahe's precise astronomical data in the early 17th century, was later explained by in his (1687) as a direct consequence of under a central gravitational force. In , the specific angular momentum \mathbf{h} = \mathbf{r} \times \mathbf{v} remains constant for motion in a , as proven for the . The magnitude h = |\mathbf{h}| determines the , the rate at which area is swept by the radius vector \mathbf{r}. The infinitesimal area dA swept in time dt is given by dA = \frac{1}{2} r^2 d[\theta](/page/Theta), where \theta is the polar angle, leading to the \frac{dA}{dt} = \frac{1}{2} r^2 \dot{[\theta](/page/Theta)}. Since the tangential component relates to h = r^2 \dot{\theta} (or more generally \frac{dA}{dt} = \frac{1}{2} |\mathbf{r} \times \mathbf{v}| = \frac{h}{2}), the areal velocity is constant at \frac{h}{2}. To derive this explicitly in polar coordinates, consider the transverse equation of motion for a , which yields \frac{d}{dt}(r^2 \dot{\theta}) = 0, implying r^2 \dot{\theta} = h (constant). Thus, \dot{\theta} = \frac{h}{r^2}, and substituting into the area element gives \frac{dA}{dt} = \frac{1}{2} h. Integrating over a finite \Delta t results in the total area swept \Delta A = \frac{h}{2} \Delta t, demonstrating that equal time intervals correspond to equal areas regardless of the planet's varying speed. Geometrically, this implies that the radius vector accelerates near the central body (where r is small, so \dot{\theta} increases to maintain h) but slows farther away, yet the net area swept remains uniform over time, providing a kinematic signature of central force without dependence on the force law's specifics.

: Elliptical Orbits

In the context of Keplerian orbits under an inverse-square central force, the shape of the trajectory is determined by combining the of specific angular momentum h with the of total \mathcal{E}. The specific angular momentum h = r^2 \dot{\theta} remains constant, allowing the radial equation of motion to be expressed in terms of the polar angle \theta. By introducing the u = 1/r (Binet's ), the radial reduce to a second-order : \frac{d^2 u}{d\theta^2} + u = \frac{\mu}{h^2}, where \mu is the (product of the and central mass). The general solution is u = \frac{\mu}{h^2} + A \cos(\theta - \theta_0), which inverts to the polar orbit equation r(\theta) = \frac{h^2 / \mu}{1 + e \cos(\theta - \theta_0)}, where e is the and \theta_0 sets the orientation of the periapsis. This equation describes conic sections with the central body at one , classified by the e: for e < 1, the orbit is an ellipse (bound trajectory); e = 0 yields a circle (special elliptical case); e = 1 a parabola (marginally unbound); and e > 1 a (unbound ). The parameter p = h^2 / \mu is the semi-latus rectum, representing the radial distance at \theta = 90^\circ from the , which scales the overall size of the conic. The e emerges from the , specifically e = \sqrt{1 + \frac{2 \mathcal{E} h^2}{\mu^2}}, linking the orbit's shape directly to the specific values of h and \mathcal{E}. The role of specific angular momentum h is pivotal in shaping elliptical orbits: for a fixed total energy \mathcal{E} < 0 (bound motion), larger h reduces e, resulting in more nearly circular orbits by increasing the semi-latus rectum p and widening the radial excursion. This can be understood through the effective potential framework, where the radial motion is governed by an effective potential V_{\text{eff}}(r) = -\frac{\mu}{r} + \frac{h^2}{2 r^2}. The centrifugal term \frac{h^2}{2 r^2} creates an angular momentum barrier that prevents collapse to the center, confining bound orbits between turning points r_{\min} and r_{\max} for \mathcal{E} < 0, with the barrier's strength scaling as h^2. Thus, h not only determines the angular scale of the orbit but also stabilizes elliptical paths against radial perturbations.

Third Law: Harmonic Periods

Kepler's third law relates the orbital period T to the semi-major axis a of an elliptical orbit around a central mass, stating that T^2 \propto a^3. In the Newtonian two-body problem, this takes the precise form T = 2\pi \sqrt{\frac{a^3}{\mu}}, where \mu = G(M + m) is the standard gravitational parameter, with G the gravitational constant, M the central mass (typically much larger than the orbiting mass m), and a the semi-major axis representing the average scale of the orbit. This formula arises from the balance of gravitational attraction and the orbital dynamics under inverse-square law forces. The specific angular momentum h, conserved in central force fields, links to Kepler's third law through the geometry of the elliptical orbit. For an ellipse, h = \sqrt{\mu p}, where p is the semi-latus rectum, and p = a(1 - e^2) with e the eccentricity; thus, h^2 = \mu a (1 - e^2). The period T depends solely on a and \mu, independent of e, so h influences T indirectly by determining the orbit's shape via e. As detailed in the first law, elliptical parameters like a and e define the conic section under gravitational forces. A derivation of the period formula integrates the orbital motion using the vis-viva equation and specific angular momentum. The vis-viva equation, from energy conservation, gives the speed as v^2 = \mu \left( \frac{2}{r} - \frac{1}{a} \right), relating instantaneous velocity to radial distance r. Combined with h = r^2 \dot{\theta} (angular momentum conservation), the time element is dt = \frac{r^2}{h} d\theta. Substituting the conic orbit equation r = \frac{h^2 / \mu}{1 + e \cos \theta} and integrating over one full revolution (\theta from 0 to $2\pi) yields T = 2\pi \sqrt{\frac{a^3}{\mu}}, confirming the third law without reliance on area rates. This relation generalizes to any conic-section orbit in a central gravitational field, but the orbital period T is defined only for bound elliptical orbits (e < 1); for parabolic (e = 1) or hyperbolic (e > 1) trajectories, no periodic closure occurs.

Modern Applications

Spacecraft Trajectories

In spacecraft trajectory design, specific angular momentum plays a crucial role in determining orbital parameters and enabling efficient maneuvers. For instance, the , an elliptical path used to shift a between two coplanar circular orbits, relies on impulsive delta-v burns at perigee and apogee to alter the specific angular momentum magnitude while conserving its direction. This method minimizes propellant consumption by achieving the transfer with the least energy, as the semi-major axis of the transfer ellipse is the average of the initial and final orbital radii, directly influencing the required change in specific angular momentum. The vector components of specific angular momentum are integral to defining key in mission planning. Its direction, perpendicular to the , specifies the inclination—the angle between the and the reference equatorial plane—and the of the ascending node, which locates the point where the crosses the from south to north. Meanwhile, the magnitude of specific angular momentum, for a given semi-major axis, governs the 's , quantifying its deviation from circularity; higher magnitudes correspond to more circular orbits, while lower ones yield more elliptical paths. These relationships allow engineers to predict and adjust orientation relative to or other central bodies during insertion or correction phases. Real-world orbits experience perturbations that gradually modify specific angular momentum, necessitating active . In , atmospheric drag exerts a tangential force that primarily reduces the semi-major axis and eccentricity but affects specific angular momentum magnitude slowly over multiple orbits, as the drag acceleration is on the order of micro-g. Similarly, Earth's J2 oblateness , arising from its , induces in the and node regression, altering the direction of specific angular momentum at rates up to several degrees per day for inclined low-altitude orbits. To counteract these effects and maintain desired trajectories, employ station-keeping maneuvers, such as periodic firings, which restore specific angular momentum to nominal values and ensure long-term stability. A practical example is geostationary satellites, which Earth at an altitude of approximately 35,786 km to match the planet's sidereal rotation period of 23 hours 56 minutes. These satellites require a specific angular momentum magnitude of about 130,000 km²/s to achieve the necessary circular equatorial with zero inclination, ensuring they remain fixed over a single on Earth's surface for continuous communication coverage. The vector's alignment with Earth's equatorial plane is critical, as any deviation would cause apparent motion relative to ground stations, requiring corrective delta-v to preserve the geostationary configuration. For trajectory optimization, numerical tools like NASA's General Mission Analysis Tool (GMAT) and AGI's (STK) incorporate specific angular momentum computations to simulate and refine spacecraft paths. GMAT, an open-source platform, models two-body dynamics and perturbations to propagate orbits and optimize delta-v sequences, outputting specific angular momentum vectors for evaluating maneuver efficiency in interplanetary transfers or Earth-orbit adjustments. STK complements this by providing visualization and multi-body propagation, allowing analysts to iterate on specific angular momentum profiles for fuel-optimal trajectories in complex mission scenarios.

Astrophysical Systems

In protoplanetary disks, the specific angular momentum of infalling gas parcels is conserved during the collapse of cores, resulting in a radial distribution that establishes Keplerian profiles where the orbital velocity scales as v \propto r^{-1/2}. This conservation drives the flattening of the collapsing material into a disk, with the disk's size determined by the initial specific angular momentum of the core, typically on the order of $10^{20} to $10^{21} cm² s⁻¹ for solar-mass systems. Torques from gravitational instabilities, , or interactions with the central star can alter this specific angular momentum, facilitating and the formation of gaps in the disk that influence growth. In systems, the specific angular momentum is defined for each star relative to the system's , contributing to the total orbital J = \mu \sqrt{G (M_1 + M_2) a (1 - e^2)}, where \mu is the , a the semi-major , and e the . During Roche lobe overflow, from the donor star carries away specific angular momentum, often comparable to that of the donor's , which can widen or shrink the binary separation depending on whether the transferred material is isotropic or carries excess angular momentum from the donor. This process is critical in close binaries, such as those evolving into cataclysmic variables, where angular momentum loss via or magnetic braking further tightens the and influences the stability of . Accretion disks around compact objects, such as black holes or neutron stars, rely on the transport of specific outward through viscous processes to enable inward radial infall of material. In the Shakura-Sunyaev α-disk model, the viscosity parameter \alpha parameterizes turbulent stresses that redistribute , creating a gradient where inner regions lose specific (typically Keplerian, h \approx \sqrt{G M r}) to outer parts, driving accretion rates up to \dot{M} \sim 10^{-8} M_\odot yr⁻¹ in active galactic nuclei. This mechanism ensures steady-state disk structure, with the inner edge truncated at the , beyond which prevents further infall without additional dissipation. In galactic dynamics, the specific angular momentum of stars orbiting the is conserved in axisymmetric potentials, analogous to central force fields, allowing stars to maintain nearly circular orbits in the disk while precessing slowly due to non-spherical perturbations. This conservation, quantified as h_z = R v_\phi where R is the cylindrical radius and v_\phi the azimuthal velocity, underpins the stability of galactic disks, with typical values around $10^{3} kpc km s⁻¹ for stars, enabling the formation of spiral arms through density waves that temporarily alter radial positions without changing h_z. Deviations from axisymmetry, such as bars, can redistribute this angular momentum, driving secular evolution and fueling central bulges. Specific angular momentum in astrophysical systems is often measured through , particularly observations of host stars, which yield orbital periods, semi-major axes, and eccentricities to compute h = \sqrt{G M_* a (1 - e^2)} per , constraining formation scenarios like disk versus in-situ growth. For instance, hot Jupiters with relatively low specific angular momentum (~10^{15} m² s⁻¹) suggest inward from outer disk regions, while higher values in temperate giants align with models incorporating torques. These measurements, combined with for inclination, refine angular momentum distributions across exoplanetary systems, revealing correlations with host star and multiplicity.

References

  1. [1]
    [PDF] Introduction to Orbital Mechanics and Spacecraft Attitudes ... - NASA
    Mar 20, 2020 · And from our earlier study of the two-body problem, we recall that the specific angular momentum, ℎ is also constant: ℎ = constant. Transfer ...
  2. [2]
    3.5 Orbital Mechanics – A Guide to CubeSat Mission and Bus Design
    The specific angular momentum (h) of a satellite remains constant. As r and v change along the orbit, the flight path angle () must change so as to keep h ...
  3. [3]
    [PDF] Orbital Mechanics
    Let h be the specific angular momentum (i.e. the angular momentum per unit mass) of m, h = r ×. ˙ r. (3). The × sign indicates the cross product. Taking the ...
  4. [4]
    [PDF] 2. Orbital Mechanics MAE 342 2016 - Robert F. Stengel
    Feb 12, 2020 · Specific Angular Momentum. Vector of a Satellite. hS = m m r × v = r × v = r × qr … is the angular momentum per unit of the satellite's mass ...
  5. [5]
    [PDF] Lecture D30 - Orbit Transfers - DSpace@MIT
    Since the velocity at the perigee is orthogonal to the position vector, the specific angular momentum of the transfer orbit is, h = r1vπ = 6.787 × 1010 m2/s ,.
  6. [6]
    Comets and Conserved Physical Quantities
    Note below the plot showing Specific Angular Momentum plotted against perihelion distance. Along the upper curve the aperiodic comets lie, while the periodic ...
  7. [7]
    The Law of Conservation of Angular Momentum - Physics
    The Law of Conservation of Angular Momentum states that, when no external torques act on a system, the angular momentum of the system is conserved. Always ...
  8. [8]
    [PDF] Central Forces - Oregon State University
    Since the angular momentum is a constant in central force problems, it's magnitude l is also constant. Therefore (43) can be used to rewrite differential.
  9. [9]
    None
    ### Summary of Central Force Problems (Chapter 4)
  10. [10]
    [PDF] Kepler's Laws - Central Force Motion - MIT OpenCourseWare
    We note that the constant of integration, h, that will be determined by the initial conditions, is precisely the magnitude of the specific angular momentum ...
  11. [11]
    [PDF] The Two-Body Problem - UCSB Physics
    The combination of linear momentum, angular momentum, and energy con- servation in our system has led to a dramatic simplification - a system of two particles ...
  12. [12]
    [PDF] 4. Central Forces - DAMTP
    We already saw in Section 2.2. 2 that angular momentum is conserved in a central potential. The proof is straightforward:<|control11|><|separator|>
  13. [13]
    [PDF] 1 CHAPTER 9 THE TWO BODY PROBLEM IN TWO DIMENSIONS ...
    In this chapter we show how Kepler's laws can be derived from Newton's laws of motion and gravitation, and conservation of angular momentum, and we derive ...
  14. [14]
    [PDF] Newton's derivation of Kepler's laws (outline) - UTK Math
    Thus Kepler's second law is equiv- alent to conservation of angular momentum, and is true for any radial force. (not necessarily following an inverse-square law) ...
  15. [15]
    [PDF] Lecture D28 - Central Force Motion: Kepler's Laws - DSpace@MIT
    In a time dt, the area, dA, swept by r will be dA = r rdθ/2. Therefore, dA dt = 1 2 r2 ˙θ = h 2 , which proves Kepler's second law.
  16. [16]
    Deriving Kepler's Laws - Elementary Differential Equations
    ... ′+rθ″=0. Conservation of Angular Momentum. From the second equation we can derive Kepler's second law. The key trick is to observe ddt(r2θ′)=2rr′θ ...
  17. [17]
    Kepler Orbits - Galileo and Einstein
    Kepler's Second Law is just conservation of angular momentum! Second, conservation of energy: this time, we act on the equation of motion with ˙→r⋅ : m˙→r ...
  18. [18]
  19. [19]
    13.5 Kepler's Laws of Planetary Motion - University Physics Volume 1
    Sep 19, 2016 · For a circular orbit, the semi-major axis (a) is the same as the radius for the orbit. In fact, Equation 13.8 gives us Kepler's third law if we ...<|control11|><|separator|>
  20. [20]
    [PDF] Lecture 3: Planar Orbital Elements: True Anomaly, Eccentricity, and ...
    An important consequence is that the rate of sweep of area is determined only by angular momentum, h. • This means that the period of the orbit depends only on ...
  21. [21]
    Hohmann Transfer - an overview | ScienceDirect Topics
    Solving for the angular momentum h, we get. (6.2) h = 2 μ r a r p r a + r p. This is a useful formula for analyzing Hohmann transfers, because knowing h we can ...
  22. [22]
    A Study on the Effects of J2 Perturbations on a Drag-Free Control ...
    The predominant perturbative forces acting on a spacecraft in LEO are J2 and higher order gravitational components, the effects of which are fairly easy to ...
  23. [23]
    [PDF] 13 AAS/AIAA Space Flight Mechanics Meeting - Dr. Hanspeter Schaub
    For low Earth orbits (LEO), the J2 gravitational perturbation is the dominant perturba- tion for a formation with spacecraft of equal type and built. While the ...
  24. [24]
    [PDF] Dynamics of a Geostationary Satellite - HAL
    Dec 4, 2017 · The specific angular momentum is perpendicular to the position and the velocities of the spacecraft and is a constant vector. Therefore, the ...
  25. [25]
    [PDF] Describing Orbits
    Instead, we define inclination as the angle between two vectors: one perpen- dicular to the orbital plane, (the specific angular momentum vector), and one ...
  26. [26]
    General Mission Analysis Tool (GMAT), Version 2011A(GSC-16228-1)
    GMAT is a software system for trajectory optimization, mission analysis, trajectory estimation, and prediction. Analysts use GMAT to design spacecraft ...
  27. [27]
    [PDF] AAS 19-824 REVISITING TRAJECTORY DESIGN WITH STK ... - Agi
    The higher momentum- transfer efficiency, or Specific Impulse, of low-thrust engines results in greater effective-payload-to- propellant ratio, or ...Missing: GMAT angular
  28. [28]
    [2405.07334] The Formation of Protoplanetary Disks through Pre ...
    May 12, 2024 · Additionally, we predict how the specific angular momentum of protoplanetary disks scales with stellar mass. Our conclusions are based on a ...
  29. [29]
    orbital periods of subdwarf B binaries produced by the first stable ...
    ... Roche Lobe overflow (RLOF), a s. ... The wind is lost from the binary by taking away the specific angular momentum as pertains to the giant.
  30. [30]
    [PDF] Chapter 9 Accretion Disks for Beginners
    In an accretion disk, angular momentum is transferred by viscous torques from the inner regions of the disk to the outer regions. The importance of accretion ...
  31. [31]
    10. Equilibria of galactic disks
    As shown in Chapter 9.1, orbits in a static, axisymmetric gravitational potential have two classical integrals of motion: the specific energy E and the ...<|separator|>
  32. [32]
    Angle–action estimation in a general axisymmetric potential
    The total angular momentum is only conserved when we are considering spherical potentials. In a spherical potential the vertical action is simply Jz = L − |Lz|.
  33. [33]
    Angular Momentum Distributions for Observed and Modeled ...
    Jan 18, 2022 · In this study, we calculate the spin angular momentum of host stars and the orbital angular momentum of their planets using data from the NASA ...