Hill sphere

Last updated
In sectional/side view, a two-dimensional representation of the three-dimensional concept of the Hill sphere, here showing the Earth's "gravity well" (gravitational potential of Earth, blue line), the same for the Moon (red line) and their combined potential (black thick line). Point P is the force free spot, where gravitational forces of Earth and Moon cancel. The sizes of Earth and Moon are in the proportion, but distances and energies are not to scale. Earth-moon-gravitational-potential.svg
In sectional/side view, a two-dimensional representation of the three-dimensional concept of the Hill sphere, here showing the Earth's "gravity well" (gravitational potential of Earth, blue line), the same for the Moon (red line) and their combined potential (black thick line). Point P is the force free spot, where gravitational forces of Earth and Moon cancel. The sizes of Earth and Moon are in the proportion, but distances and energies are not to scale.

The Hill sphere is a common model for the calculation of a gravitational sphere of influence. It is the most commonly used model to calculate the spatial extent of gravitational influence of an astronomical body (m) in which it dominates over the gravitational influence of other bodies, particularly a primary (M). [1] It is sometimes confused with other models of gravitational influence, such as the Laplace sphere [1] or being called the Roche sphere, the latter causing confusion with the Roche limit. [2] [3] It was defined by the American astronomer George William Hill, based on the work of the French astronomer Édouard Roche.[ not verified in body ]

Contents

To be retained by a more gravitationally attracting astrophysical object—a planet by a more massive star, a moon by a more massive planet—the less massive body must have an orbit that lies within the gravitational potential represented by the more massive body's Hill sphere.[ not verified in body ] That moon would, in turn, have a Hill sphere of its own, and any object within that distance would tend to become a satellite of the moon, rather than of the planet itself.[ not verified in body ]

A contour plot of the effective gravitational potential of a two-body system, here, the Sun and Earth, indicating the five Lagrange points. Lagrange points2.svg
A contour plot of the effective gravitational potential of a two-body system, here, the Sun and Earth, indicating the five Lagrange points.

One simple view of the extent of the Solar System is that it is bounded by the Hill sphere of the Sun (engendered by the Sun's interaction with the galactic nucleus or other more massive stars). [4] [ verification needed ] A more complex example is the one at right, the Earth's Hill sphere, which extends between the Lagrange points L1 and L2,[ clarification needed ] which lie along the line of centers of the Earth and the more massive Sun.[ not verified in body ] The gravitational influence of the less massive body is least in that direction, and so it acts as the limiting factor for the size of the Hill sphere;[ clarification needed ] beyond that distance, a third object in orbit around the Earth would spend at least part of its orbit outside the Hill sphere, and would be progressively perturbed by the tidal forces of the more massive body, the Sun, eventually ending up orbiting the latter.[ not verified in body ]

For two massive bodies with gravitational potentials and any given energy of a third object of negligible mass interacting with them, one can define a zero-velocity surface in space which cannot be passed, the contour of the Jacobi integral.[ not verified in body ] When the object's energy is low, the zero-velocity surface completely surrounds the less massive body (of this restricted three-body system), which means the third object cannot escape; at higher energy, there will be one or more gaps or bottlenecks by which the third object may escape the less massive body and go into orbit around the more massive one.[ not verified in body ] If the energy is at the border between these two cases, then the third object cannot escape, but the zero-velocity surface confining it touches a larger zero-velocity surface around the less massive body[ verification needed ] at one of the nearby Lagrange points, forming a cone-like point there.[ clarification needed ][ not verified in body ] At the opposite side of the less massive body, the zero-velocity surface gets close to the other Lagrange point.[ not verified in body ] This limiting zero-velocity surface around the less massive body is its Hill "sphere".[ according to whom? ][ original research? ]

Definition

The Hill radius or sphere (the latter defined by the former radius[ citation needed ]) has been described as "the region around a planetary body where its own gravity (compared to that of the Sun or other nearby bodies) is the dominant force in attracting satellites," both natural and artificial. [5] [ better source needed ]

As described by de Pater and Lissauer, all bodies within a system such as the Sun's Solar System "feel the gravitational force of one another", and while the motions of just two gravitationally interacting bodies—constituting a "two-body problem"—are "completely integrable ([meaning]...there exists one independent integral or constraint per degree of freedom)" and thus an exact, analytic solution, the interactions of three (or more) such bodies "cannot be deduced analytically", requiring instead solutions by numerical integration, when possible. [6] :p.26 This is the case, unless the negligible mass of one of the three bodies allows approximation of the system as a two-body problem, known formally as a "restricted three-body problem". [6] :p.26

For such two- or restricted three-body problems as its simplest examples—e.g., one more massive primary astrophysical body, mass of m1, and a less massive secondary body, mass of m2—the concept of a Hill radius or sphere is of the approximate limit to the secondary mass's "gravitational dominance", [6] a limit defined by "the extent" of its Hill sphere, which is represented mathematically as follows: [6] :p.29 [7]

,

where, in this representation, major axis "a" can be understood as the "instantaneous heliocentric distance" between the two masses (elsewhere abbreviated rp). [6] :p.29 [7]

More generally, if the less massive body, , orbits a more massive body (m1, e.g., as a planet orbiting around the Sun) and has a semi-major axis , and an eccentricity of , then the Hill radius or sphere, of the less massive body, calculated at the pericenter, is approximately: [8] [ non-primary source needed ][ better source needed ]

When eccentricity is negligible (the most favourable case for orbital stability), this expression reduces to the one presented above.[ citation needed ]

Example and derivation

A schematic, not-to-scale representation of Hill spheres (as 2D radii) and Roche limits of each body of the Sun-Earth-Moon system. The actual Hill radius for the Earth-Moon pair is on the order of 60,000 km (i.e., extending less than one-sixth the distance of the 378,000 km between the Moon and the Earth). Comparison of Hill sphere and Roche limit.svg
A schematic, not-to-scale representation of Hill spheres (as 2D radii) and Roche limits of each body of the Sun-Earth-Moon system. The actual Hill radius for the Earth-Moon pair is on the order of 60,000 km (i.e., extending less than one-sixth the distance of the 378,000 km between the Moon and the Earth).

In the Earth-Sun example, the Earth (5.97×1024 kg) orbits the Sun (1.99×1030 kg) at a distance of 149.6 million km, or one astronomical unit (AU). The Hill sphere for Earth thus extends out to about 1.5 million km (0.01 AU). The Moon's orbit, at a distance of 0.384 million km from Earth, is comfortably within the gravitational sphere of influence of Earth and it is therefore not at risk of being pulled into an independent orbit around the Sun.

The earlier eccentricity-ignoring formula can be re-stated as follows:

, or ,

where M is the sum of the interacting masses.

Derivation

The expression for the Hill radius can be found by equating gravitational and centrifugal forces acting on a test particle (of mass much smaller than ) orbiting the secondary body. Assume that the distance between masses and is , and that the test particle is orbiting at a distance from the secondary. When the test particle is on the line connecting the primary and the secondary body, the force balance requires that

where is the gravitational constant and is the (Keplerian) angular velocity of the secondary about the primary (assuming that ). The above equation can also be written as

which, through a binomial expansion to leading order in , can be written as

Hence, the relation stated above

If the orbit of the secondary about the primary is elliptical, the Hill radius is maximum at the apocenter, where is largest, and minimum at the pericenter of the orbit. Therefore, for purposes of stability of test particles (for example, of small satellites), the Hill radius at the pericenter distance needs to be considered.

To leading order in , the Hill radius above also represents the distance of the Lagrangian point L1 from the secondary.

Regions of stability

The Hill sphere is only an approximation, and other forces (such as radiation pressure or the Yarkovsky effect) can eventually perturb an object out of the sphere.[ citation needed ] As stated, the satellite (third mass) should be small enough that its gravity contributes negligibly. [6] :p.26ff

Detailed numerical calculations show that orbits at or just within the Hill sphere are not stable in the long term; it appears that stable satellite orbits exist only inside 1/2 to 1/3 of the Hill radius.[ citation needed ]

The region of stability for retrograde orbits at a large distance from the primary is larger than the region for prograde orbits at a large distance from the primary. This was thought to explain the preponderance of retrograde moons around Jupiter; however, Saturn has a more even mix of retrograde/prograde moons so the reasons are more complicated. [10]

Further examples

It is possible for a Hill sphere to be so small that it is impossible to maintain an orbit around a body. For example, an astronaut could not have orbited the 104 ton Space Shuttle at an orbit 300 km above the Earth, because a 104-ton object at that altitude has a Hill sphere of only 120 cm in radius, much smaller than a Space Shuttle. A sphere of this size and mass would be denser than lead, and indeed, in low Earth orbit, a spherical body must be more dense than lead in order to fit inside its own Hill sphere, or else it will be incapable of supporting an orbit. Satellites further out in geostationary orbit, however, would only need to be more than 6% of the density of water to fit inside their own Hill sphere.[ citation needed ]

Within the Solar System, the planet with the largest Hill radius is Neptune, with 116 million km, or 0.775 au; its great distance from the Sun amply compensates for its small mass relative to Jupiter (whose own Hill radius measures 53 million km). An asteroid from the asteroid belt will have a Hill sphere that can reach 220,000 km (for 1 Ceres), diminishing rapidly with decreasing mass. The Hill sphere of 66391 Moshup, a Mercury-crossing asteroid that has a moon (named Squannit), measures 22 km in radius. [11]

A typical extrasolar "hot Jupiter", HD 209458 b, [12] has a Hill sphere radius of 593,000 km, about eight times its physical radius of approx 71,000 km. Even the smallest close-in extrasolar planet, CoRoT-7b, [13] still has a Hill sphere radius (61,000 km), six times its physical radius (approx 10,000 km). Therefore, these planets could have small moons close in, although not within their respective Roche limits.[ citation needed ]

Hill spheres for the solar system

The following table and logarithmic plot show the radius of the Hill spheres of some bodies of the Solar System calculated with the first formula stated above (including orbital eccentricity), using values obtained from the JPL DE405 ephemeris and from the NASA Solar System Exploration website. [14]

Radius of the Hill spheres of some bodies of the Solar System
BodyMillion km au Body radii Arcminutes [note 1] Farthest moon (au)
Mercury 0.17530.001271.910.7
Venus 1.00420.0067165.931.8
Earth 1.47140.0098230.733.70.00257
Mars 0.98270.0066289.314.90.00016
Jupiter 50.57360.3381707.4223.20.1662
Saturn 61.63400.41201022.7147.80.1785
Uranus 66.78310.44642613.180.00.1366
Neptune 115.03070.76894644.687.90.3360
Ceres 0.20480.0014433.01.7
Pluto 5.99210.04015048.13.50.00043
Eris 8.11760.05436979.92.70.00025
Logarithmic plot of the Hill radii for the bodies of the solar system Hill sphere of the planets.png
Logarithmic plot of the Hill radii for the bodies of the solar system

See also

Explanatory notes

  1. At average distance, as seen from the Sun. The angular size as seen from Earth varies depending on Earth's proximity to the object.

Related Research Articles

<span class="mw-page-title-main">Double planet</span> A binary system where two planetary-mass objects share an orbital axis external to both

In astronomy, a double planet is a binary satellite system where both objects are planets, or planetary-mass objects, that share an orbital axis external to both planetary bodies.

<span class="mw-page-title-main">Lagrange point</span> Equilibrium points near two orbiting bodies

In celestial mechanics, the Lagrange points are points of equilibrium for small-mass objects under the gravitational influence of two massive orbiting bodies. Mathematically, this involves the solution of the restricted three-body problem.

<span class="mw-page-title-main">Orbit</span> Curved path of an object around a point

In celestial mechanics, an orbit is the curved trajectory of an object such as the trajectory of a planet around a star, or of a natural satellite around a planet, or of an artificial satellite around an object or position in space such as a planet, moon, asteroid, or Lagrange point. Normally, orbit refers to a regularly repeating trajectory, although it may also refer to a non-repeating trajectory. To a close approximation, planets and satellites follow elliptic orbits, with the center of mass being orbited at a focal point of the ellipse, as described by Kepler's laws of planetary motion.

<span class="mw-page-title-main">Tidal force</span> A gravitational effect also known as the differential force and the perturbing force

The tidal force or tide-generating force is a gravitational effect that stretches a body along the line towards and away from the center of mass of another body due to spatial variations in strength in gravitational field from the other body. It is responsible for the tides and related phenomena, including solid-earth tides, tidal locking, breaking apart of celestial bodies and formation of ring systems within the Roche limit, and in extreme cases, spaghettification of objects. It arises because the gravitational field exerted on one body by another is not constant across its parts: the nearer side is attracted more strongly than the farther side. The difference is positive in the near side and negative in the far side, which causes a body to get stretched. Thus, the tidal force is also known as the differential force, residual force, or secondary effect of the gravitational field.

<span class="mw-page-title-main">Escape velocity</span> Concept in celestial mechanics

In celestial mechanics, escape velocity or escape speed is the minimum speed needed for an object to escape from contact with or orbit of a primary body, assuming:

<span class="mw-page-title-main">Roche limit</span> Orbital radius that will break up a satellite

In celestial mechanics, the Roche limit, also called Roche radius, is the distance from a celestial body within which a second celestial body, held together only by its own force of gravity, will disintegrate because the first body's tidal forces exceed the second body's self-gravitation. Inside the Roche limit, orbiting material disperses and forms rings, whereas outside the limit, material tends to coalesce. The Roche radius depends on the radius of the first body and on the ratio of the bodies' densities.

The orbital period is the amount of time a given astronomical object takes to complete one orbit around another object. In astronomy, it usually applies to planets or asteroids orbiting the Sun, moons orbiting planets, exoplanets orbiting other stars, or binary stars. It may also refer to the time it takes a satellite orbiting a planet or moon to complete one orbit.

<span class="mw-page-title-main">Orbital mechanics</span> Field of classical mechanics concerned with the motion of spacecraft

Orbital mechanics or astrodynamics is the application of ballistics and celestial mechanics to the practical problems concerning the motion of rockets, satellites, and other spacecraft. The motion of these objects is usually calculated from Newton's laws of motion and the law of universal gravitation. Orbital mechanics is a core discipline within space-mission design and control.

<span class="mw-page-title-main">Barycenter (astronomy)</span> Center of mass of multiple bodies orbiting each other

In astronomy, the barycenter is the center of mass of two or more bodies that orbit one another and is the point about which the bodies orbit. A barycenter is a dynamical point, not a physical object. It is an important concept in fields such as astronomy and astrophysics. The distance from a body's center of mass to the barycenter can be calculated as a two-body problem.

In gravitationally bound systems, the orbital speed of an astronomical body or object is the speed at which it orbits around either the barycenter or, if one body is much more massive than the other bodies of the system combined, its speed relative to the center of mass of the most massive body.

Newton's law of universal gravitation says that every particle attracts every other particle in the universe with a force that is proportional to the product of their masses and inversely proportional to the square of the distance between their centers. Separated objects attract and are attracted as if all their mass were concentrated at their centers. The publication of the law has become known as the "first great unification", as it marked the unification of the previously described phenomena of gravity on Earth with known astronomical behaviors.

<span class="mw-page-title-main">Earth's orbit</span> Trajectory of Earth around the Sun

Earth orbits the Sun at an average distance of 149.60 million km in a counterclockwise direction as viewed from above the Northern Hemisphere. One complete orbit takes 365.256 days, during which time Earth has traveled 940 million km. Ignoring the influence of other Solar System bodies, Earth's orbit, also known as Earth's revolution, is an ellipse with the Earth-Sun barycenter as one focus with a current eccentricity of 0.0167. Since this value is close to zero, the center of the orbit is relatively close to the center of the Sun.

In celestial mechanics, the standard gravitational parameterμ of a celestial body is the product of the gravitational constant G and the total mass M of the bodies. For two bodies, the parameter may be expressed as G(m1 + m2), or as GM when one body is much larger than the other:

A sphere of influence (SOI) in astrodynamics and astronomy is the oblate-spheroid-shaped region around a celestial body where the primary gravitational influence on an orbiting object is that body. This is usually used to describe the areas in the Solar System where planets dominate the orbits of surrounding objects such as moons, despite the presence of the much more massive but distant Sun. In the patched conic approximation, used in estimating the trajectories of bodies moving between the neighbourhoods of different masses using a two body approximation, ellipses and hyperbolae, the SOI is taken as the boundary where the trajectory switches which mass field it is influenced by. It is not to be confused with the sphere of activity which extends well beyond the sphere of influence.

<span class="mw-page-title-main">Orbital eccentricity</span> Amount by which an orbit deviates from a perfect circle

In astrodynamics, the orbital eccentricity of an astronomical object is a dimensionless parameter that determines the amount by which its orbit around another body deviates from a perfect circle. A value of 0 is a circular orbit, values between 0 and 1 form an elliptic orbit, 1 is a parabolic escape orbit, and greater than 1 is a hyperbola. The term derives its name from the parameters of conic sections, as every Kepler orbit is a conic section. It is normally used for the isolated two-body problem, but extensions exist for objects following a rosette orbit through the Galaxy.

For most numbered asteroids, almost nothing is known apart from a few physical parameters and orbital elements. Some physical characteristics can only be estimated. The physical data is determined by making certain standard assumptions.

"Clearing the neighbourhood" around a celestial body's orbit describes the body becoming gravitationally dominant such that there are no other bodies of comparable size other than its natural satellites or those otherwise under its gravitational influence.

<span class="mw-page-title-main">Jupiter mass</span> Unit of mass equal to the total mass of the planet Jupiter

Jupiter mass, also called Jovian mass, is the unit of mass equal to the total mass of the planet Jupiter. This value may refer to the mass of the planet alone, or the mass of the entire Jovian system to include the moons of Jupiter. Jupiter is by far the most massive planet in the Solar System. It is approximately 2.5 times as massive as all of the other planets in the Solar System combined.

The two-body problem in general relativity is the determination of the motion and gravitational field of two bodies as described by the field equations of general relativity. Solving the Kepler problem is essential to calculate the bending of light by gravity and the motion of a planet orbiting its sun. Solutions are also used to describe the motion of binary stars around each other, and estimate their gradual loss of energy through gravitational radiation.

The innermost stable circular orbit is the smallest marginally stable circular orbit in which a test particle can stably orbit a massive object in general relativity. The location of the ISCO, the ISCO-radius, depends on the mass and angular momentum (spin) of the central object. The ISCO plays an important role in black hole accretion disks since it marks the inner edge of the disk.

References

  1. 1 2 Souami, D.; Cresson, J.; Biernacki, C.; Pierret, F. (2020). "On the local and global properties of gravitational spheres of influence". Monthly Notices of the Royal Astronomical Society. 496 (4): 4287–4297. arXiv: 2005.13059 . doi:10.1093/mnras/staa1520.
  2. Williams, Matt (2015-12-30). "How Many Moons Does Mercury Have?". Universe Today. Retrieved 2023-11-08.
  3. Hill, Roderick J. (2022). "Gravitational clearing of natural satellite orbits". Publications of the Astronomical Society of Australia. Cambridge University Press. 39. Bibcode:2022PASA...39....6H. doi:10.1017/pasa.2021.62. ISSN   1323-3580. S2CID   246637375.
  4. Chebotarev, G. A. (March 1965). "On the Dynamical Limits of the Solar System". Soviet Astronomy. 8: 787. Bibcode:1965SvA.....8..787C.
  5. Lauretta, Dante and the Staff of the Osiris-Rex Asteroid Sample Return Mission (2023). "Word of the Week: Hill Sphere". Osiris-Rex Asteroid Sample Return Mission (AsteroidMission.org). Tempe, AZ: University of Arizona. Retrieved July 22, 2023.
  6. 1 2 3 4 5 6 de Pater, Imke & Lissauer, Jack (2015). "Dynamics (The Three-Body Problem, Perturbations and Resonances)". Planetary Sciences (2nd ed.). Cambridge, England: Cambridge University Press. pp. 26, 28–30, 34. ISBN   9781316195697 . Retrieved 22 July 2023.{{cite book}}: CS1 maint: multiple names: authors list (link)
  7. 1 2 Higuchi1, A. & Ida, S. (April 2017). "Temporary Capture of Asteroids by an Eccentric Planet". The Astronomical Journal. Washington, DC: The American Astronomical Society. 153 (4): 155. arXiv: 1702.07352 . Bibcode:2017AJ....153..155H. doi: 10.3847/1538-3881/aa5daa . S2CID   119036212.{{cite journal}}: CS1 maint: multiple names: authors list (link) CS1 maint: numeric names: authors list (link)
  8. Hamilton, D.P. & Burns, J.A. (March 1992). "Orbital Stability Zones About Asteroids: II. The Destabilizing Effects of Eccentric Orbits and of Solar Radiation". Icarus. New York, NY: Academic Press. 96 (1): 43–64. Bibcode:1992Icar...96...43H. doi: 10.1016/0019-1035(92)90005-R .{{cite journal}}: CS1 maint: multiple names: authors list (link) See also Hamilton, D.P. & Burns, J.A. (March 1991). "Orbital Stability Zones About Asteroids" (PDF). Icarus. New York, NY: Academic Press. 92 (1): 118–131. Bibcode:1991Icar...92..118H. doi:10.1016/0019-1035(91)90039-V . Retrieved 22 July 2023.{{cite journal}}: CS1 maint: multiple names: authors list (link) cited therein.
  9. Follows, Mike (4 October 2017). "Ever Decreasing Circles". NewScientist.com . Retrieved 23 July 2023. The moon's Hill sphere has a radius of 60,000 kilometres, about one-sixth of the distance between it and Earth.
  10. Astakhov, Sergey A.; Burbanks, Andrew D.; Wiggins, Stephen & Farrelly, David (2003). "Chaos-assisted capture of irregular moons". Nature . 423 (6937): 264–267. Bibcode:2003Natur.423..264A. doi:10.1038/nature01622. PMID   12748635. S2CID   16382419.
  11. Johnston, Robert (20 October 2019). "(66391) Moshup and Squannit". Johnston's Archive. Retrieved 30 March 2017.
  12. "HD 209458 b". Extrasolar Planets Encyclopaedia . Archived from the original on 2010-01-16. Retrieved 2010-02-16.
  13. "Planet CoRoT-7 b". Extrasolar Planets Encyclopaedia .
  14. "NASA Solar System Exploration". NASA. Retrieved 2020-12-22.

Further reading