Orbital resonance

Last updated

The three-body Laplace resonance exhibited by three of Jupiter's Galilean moons. Conjunctions are highlighted by brief color changes. There are two Io-Europa conjunctions (green) and three Io-Ganymede conjunctions (grey) for each Europa-Ganymede conjunction (magenta). This diagram is not to scale. Galilean moon Laplace resonance animation 2.gif
The three-body Laplace resonance exhibited by three of Jupiter's Galilean moons. Conjunctions are highlighted by brief color changes. There are two Io-Europa conjunctions (green) and three Io-Ganymede conjunctions (grey) for each Europa-Ganymede conjunction (magenta). This diagram is not to scale.

In celestial mechanics, orbital resonance occurs when orbiting bodies exert regular, periodic gravitational influence on each other, usually because their orbital periods are related by a ratio of small integers. Most commonly, this relationship is found between a pair of objects (binary resonance). The physical principle behind orbital resonance is similar in concept to pushing a child on a swing, whereby the orbit and the swing both have a natural frequency, and the body doing the "pushing" will act in periodic repetition to have a cumulative effect on the motion. Orbital resonances greatly enhance the mutual gravitational influence of the bodies (i.e., their ability to alter or constrain each other's orbits). In most cases, this results in an unstable interaction, in which the bodies exchange momentum and shift orbits until the resonance no longer exists. Under some circumstances, a resonant system can be self-correcting and thus stable. Examples are the 1:2:4 resonance of Jupiter's moons Ganymede, Europa and Io, and the 2:3 resonance between Neptune and Pluto. Unstable resonances with Saturn's inner moons give rise to gaps in the rings of Saturn. The special case of 1:1 resonance between bodies with similar orbital radii causes large planetary system bodies to eject most other bodies sharing their orbits; this is part of the much more extensive process of clearing the neighbourhood, an effect that is used in the current definition of a planet. [1]

Contents

A binary resonance ratio in this article should be interpreted as the ratio of number of orbits completed in the same time interval, rather than as the ratio of orbital periods, which would be the inverse ratio. Thus, the 2:3 ratio above means that Pluto completes two orbits in the time it takes Neptune to complete three. In the case of resonance relationships among three or more bodies, either type of ratio may be used (whereby the smallest whole-integer ratio sequences are not necessarily reversals of each other), and the type of ratio will be specified.

History

Since the discovery of Newton's law of universal gravitation in the 17th century, the stability of the Solar System has preoccupied many mathematicians, starting with Pierre-Simon Laplace. The stable orbits that arise in a two-body approximation ignore the influence of other bodies. The effect of these added interactions on the stability of the Solar System is very small, but at first it was not known whether they might add up over longer periods to significantly change the orbital parameters and lead to a completely different configuration, or whether some other stabilising effects might maintain the configuration of the orbits of the planets.

It was Laplace who found the first answers explaining the linked orbits of the Galilean moons (see below). Before Newton, there was also consideration of ratios and proportions in orbital motions, in what was called "the music of the spheres", or musica universalis .

The article on resonant interactions describes resonance in the general modern setting. A primary result from the study of dynamical systems is the discovery and description of a highly simplified model of mode-locking; this is an oscillator that receives periodic kicks via a weak coupling to some driving motor. The analog here would be that a more massive body provides a periodic gravitational kick to a smaller body, as it passes by. The mode-locking regions are named Arnold tongues.

Types of resonance

The semimajor axes of resonant trans-Neptunian objects (red) are clumped at locations of low-integer resonances with Neptune (vertical red bars near top), in contrast to those of cubewanos (blue) and nonresonant (or not known to be resonant) scattered objects (grey). TheKuiperBelt 75AU All.svg
The semimajor axes of resonant trans-Neptunian objects (red) are clumped at locations of low-integer resonances with Neptune (vertical red bars near top), in contrast to those of cubewanos (blue) and nonresonant (or not known to be resonant) scattered objects (grey).
A chart of the distribution of asteroid semimajor axes, showing the Kirkwood gaps where orbits are destabilized by resonances with Jupiter Kirkwood Gaps.svg
A chart of the distribution of asteroid semimajor axes, showing the Kirkwood gaps where orbits are destabilized by resonances with Jupiter
Spiral density waves in Saturn's A Ring excited by resonances with inner moons. Such waves propagate away from the planet (towards upper left). The large set of waves just below center is due to the 6:5 resonance with Janus. PIA10452 - Saturn A ring spiral density waves.jpg
Spiral density waves in Saturn's A Ring excited by resonances with inner moons. Such waves propagate away from the planet (towards upper left). The large set of waves just below center is due to the 6:5 resonance with Janus.
The eccentric Titan Ringlet in the Columbo Gap of Saturn's C Ring (center) and the inclined orbits of resonant particles in the bending wave just inside it have apsidal and nodal precessions, respectively, commensurate with Titan's mean motion. PIA17173 Titan resonances in Saturn's C ring.jpg
The eccentric Titan Ringlet in the Columbo Gap of Saturn's C Ring (center) and the inclined orbits of resonant particles in the bending wave just inside it have apsidal and nodal precessions, respectively, commensurate with Titan's mean motion.

In general, an orbital resonance may

A mean-motion orbital resonance occurs when two bodies have periods of revolution that are a simple integer ratio of each other. It does not depend only on the existence of such a ratio, and more precisely the ratio of periods is not exactly an rational number, even averaged over a long period. For example, in the case of Pluto and Neptune (see below), the true equation says that the average rate of change of is exactly zero, where is the longitude of Pluto, is the longitude of Neptune, and is the longitude of Pluto's perihelion. Since the rate of motion of the latter is about 0.97×10−4 degrees per year, the ratio of periods is actually 1.503 in the long term. [5]

Depending on the details, mean-motion orbital resonance can either stabilize or destabilize the orbit. Stabilization may occur when the two bodies move in such a synchronised fashion that they never closely approach. For instance:

Orbital resonances can also destabilize one of the orbits. This process can be exploited to find energy-efficient ways of deorbiting spacecraft. [6] [7] For small bodies, destabilization is actually far more likely. For instance:

Most bodies that are in resonance orbit in the same direction; however, the retrograde asteroid 514107 Kaʻepaokaʻawela appears to be in a stable (for a period of at least a million years) 1:−1 resonance with Jupiter. [8] In addition, a few retrograde damocloids have been found that are temporarily captured in mean-motion resonance with Jupiter or Saturn. [9] Such orbital interactions are weaker than the corresponding interactions between bodies orbiting in the same direction. [9] [10] The trans-Neptunian object 2011 KT19 has an orbital inclination of 110° with respect to the planets' orbital plane and is currently in a 7:9 polar resonance with Neptune. [11]

A Laplace resonance is a three-body resonance with a 1:2:4 orbital period ratio (equivalent to a 4:2:1 ratio of orbits). The term arose because Pierre-Simon Laplace discovered that such a resonance governed the motions of Jupiter's moons Io, Europa, and Ganymede. It is now also often applied to other 3-body resonances with the same ratios, [12] such as that between the extrasolar planets Gliese 876 c, b, and e. [13] [14] [15] Three-body resonances involving other simple integer ratios have been termed "Laplace-like" [16] or "Laplace-type". [17]

A Lindblad resonance drives spiral density waves both in galaxies (where stars are subject to forcing by the spiral arms themselves) and in Saturn's rings (where ring particles are subject to forcing by Saturn's moons).

A secular resonance occurs when the precession of two orbits is synchronised (usually a precession of the perihelion or ascending node). A small body in secular resonance with a much larger one (e.g. a planet) will precess at the same rate as the large body. Over long times (a million years, or so) a secular resonance will change the eccentricity and inclination of the small body.

Several prominent examples of secular resonance involve Saturn. There is a near-resonance between the precession of Saturn's rotational axis and that of Neptune's orbital axis (both of which have periods of about 1.87 million years), which has been identified as the likely source of Saturn's large axial tilt (26.7°). [18] [19] [20] Initially, Saturn probably had a tilt closer to that of Jupiter (3.1°). The gradual depletion of the Kuiper belt would have decreased the precession rate of Neptune's orbit; eventually, the frequencies matched, and Saturn's axial precession was captured into a spin-orbit resonance, leading to an increase in Saturn's obliquity. (The angular momentum of Neptune's orbit is 104 times that of Saturn's rotation rate, and thus dominates the interaction.) However, it seems that the resonance no longer exists. Detailed analysis of data from the Cassini spacecraft gives a value of the moment of inertia of Saturn that is just outside the range for the resonance to exist, meaning that the spin axis does not stay in phase with Neptune's orbital inclination in the long term, as it apparently did in the past. One theory for why the resonance came to an end is that there was another moon around Saturn whose orbit destabilized about 100 million years ago, perturbing Saturn. [21] [22]

The perihelion secular resonance between asteroids and Saturn (ν6 = gg6) helps shape the asteroid belt (the subscript "6" identifies Saturn as the sixth planet from the Sun). Asteroids which approach it have their eccentricity slowly increased until they become Mars-crossers, at which point they are usually ejected from the asteroid belt by a close pass to Mars. This resonance forms the inner and "side" boundaries of the asteroid belt around 2 AU, and at inclinations of about 20°.

Numerical simulations have suggested that the eventual formation of a perihelion secular resonance between Mercury and Jupiter (g1 = g5) has the potential to greatly increase Mercury's eccentricity and possibly destabilize the inner Solar System several billion years from now. [23] [24]

The Titan Ringlet within Saturn's C Ring represents another type of resonance in which the rate of apsidal precession of one orbit exactly matches the speed of revolution of another. The outer end of this eccentric ringlet always points towards Saturn's major moon Titan. [2]

A Kozai resonance occurs when the inclination and eccentricity of a perturbed orbit oscillate synchronously (increasing eccentricity while decreasing inclination and vice versa). This resonance applies only to bodies on highly inclined orbits; as a consequence, such orbits tend to be unstable, since the growing eccentricity would result in small pericenters, typically leading to a collision or (for large moons) destruction by tidal forces.

In an example of another type of resonance involving orbital eccentricity, the eccentricities of Ganymede and Callisto vary with a common period of 181 years, although with opposite phases. [25]

Mean-motion resonances in the Solar System

Depiction of Haumea's presumed 7:12 resonance with Neptune in a rotating frame, with Neptune (blue dot at lower right) held stationary. Haumea's shifting orbital alignment relative to Neptune periodically reverses (librates), preserving the resonance. Haumea.GIF
Depiction of Haumea's presumed 7:12 resonance with Neptune in a rotating frame, with Neptune (blue dot at lower right) held stationary. Haumea's shifting orbital alignment relative to Neptune periodically reverses (librates), preserving the resonance.

There are only a few known mean-motion resonances (MMR) in the Solar System involving planets, dwarf planets or larger satellites (a much greater number involve asteroids, planetary rings, moonlets and smaller Kuiper belt objects, including many possible dwarf planets).

Additionally, Haumea is thought to be in a 7:12 resonance with Neptune, [26] [27] and Gonggong is thought to be in a 3:10 resonance with Neptune. [28]

The simple integer ratios between periods hide more complex relations:

As illustration of the latter, consider the well-known 2:1 resonance of Io-Europa. If the orbiting periods were in this relation, the mean motions (inverse of periods, often expressed in degrees per day) would satisfy the following

Substituting the data (from Wikipedia) one will get −0.7395° day−1, a value substantially different from zero.

Actually, the resonance is perfect, but it involves also the precession of perijove (the point closest to Jupiter), . The correct equation (part of the Laplace equations) is:

In other words, the mean motion of Io is indeed double of that of Europa taking into account the precession of the perijove. An observer sitting on the (drifting) perijove will see the moons coming into conjunction in the same place (elongation). The other pairs listed above satisfy the same type of equation with the exception of Mimas-Tethys resonance. In this case, the resonance satisfies the equation

The point of conjunctions librates around the midpoint between the nodes of the two moons.

Laplace resonance

Illustration of Io-Europa-Ganymede resonance. From the centre outwards: Io (yellow), Europa (gray), and Ganymede (dark) TheLaplaceResonance2.png
Illustration of Io–Europa–Ganymede resonance. From the centre outwards: Io (yellow), Europa (gray), and Ganymede (dark)

The Laplace resonance involving Io–Europa–Ganymede includes the following relation locking the orbital phase of the moons:

where are mean longitudes of the moons (the second equals sign ignores libration).

This relation makes a triple conjunction impossible. (A Laplace resonance in the Gliese 876 system, in contrast, is associated with one triple conjunction per orbit of the outermost planet, ignoring libration.) The graph illustrates the positions of the moons after 1, 2, and 3 Io periods. librates about 180° with an amplitude of 0.03°. [29]

Another "Laplace-like" resonance involves the moons Styx, Nix, and Hydra of Pluto: [16]

This reflects orbital periods for Styx, Nix, and Hydra, respectively, that are close to a ratio of 18:22:33 (or, in terms of the near resonances with Charon's period, 3+3/11:4:6; see below); the respective ratio of orbits is 11:9:6. Based on the ratios of synodic periods, there are 5 conjunctions of Styx and Hydra and 3 conjunctions of Nix and Hydra for every 2 conjunctions of Styx and Nix. [16] [30] As with the Galilean satellite resonance, triple conjunctions are forbidden. librates about 180° with an amplitude of at least 10°. [16]

Hydra, Nix, Styx conjunctions cycle.png
Hydra, Nix, Styx conjunctions cycle.png
Sequence of conjunctions of Hydra (blue), Nix (red), and Styx (black) over one third of their resonance cycle. Movements are counterclockwise and orbits completed are tallied at upper right of diagrams (click on image to see the whole cycle).

Plutino resonances

The dwarf planet Pluto is following an orbit trapped in a web of resonances with Neptune. The resonances include:

One consequence of these resonances is that a separation of at least 30 AU is maintained when Pluto crosses Neptune's orbit. The minimum separation between the two bodies overall is 17 AU, while the minimum separation between Pluto and Uranus is just 11 AU [31] (see Pluto's orbit for detailed explanation and graphs).

The next largest body in a similar 2:3 resonance with Neptune, called a plutino , is the probable dwarf planet Orcus. Orcus has an orbit similar in inclination and eccentricity to Pluto's. However, the two are constrained by their mutual resonance with Neptune to always be in opposite phases of their orbits; Orcus is thus sometimes described as the "anti-Pluto". [32]

Depiction of the resonance between Neptune's moons Naiad (whose orbital motion is shown in red) and Thalassa, in a view that co-rotates with the latter Naiad-Thalassa 73-69 orbital resonance.jpg
Depiction of the resonance between Neptune's moons Naiad (whose orbital motion is shown in red) and Thalassa, in a view that co-rotates with the latter

Naiad:Thalassa 73:69 resonance

Neptune's innermost moon, Naiad, is in a 73:69 fourth-order resonance with the next outward moon, Thalassa. As it orbits Neptune, the more inclined Naiad successively passes Thalassa twice from above and then twice from below, in a cycle that repeats every ~21.5 Earth days. The two moons are about 3540 km apart when they pass each other. Although their orbital radii differ by only 1850 km, Naiad swings ~2800 km above or below Thalassa's orbital plane at closest approach. As is common, this resonance stabilizes the orbits by maximizing separation at conjunction, but it is unusual for the role played by orbital inclination in facilitating this avoidance in a case where eccentricities are minimal. [33] [34] [note 1]

Mean-motion resonances among extrasolar planets

Resonant planetary system of two planets with a 1:2 orbit ratio Resonant planetary system.gif
Resonant planetary system of two planets with a 1:2 orbit ratio

While most extrasolar planetary systems discovered have not been found to have planets in mean-motion resonances, chains of up to five resonant planets [36] and up to seven at least near resonant planets [37] have been uncovered. Simulations have shown that during planetary system formation, the appearance of resonant chains of planetary embryos is favored by the presence of the primordial gas disc. Once that gas dissipates, 90–95% of those chains must then become unstable to match the low frequency of resonant chains observed. [38]

Cases of extrasolar planets close to a 1:2 mean-motion resonance are fairly common. Sixteen percent of systems found by the transit method are reported to have an example of this (with period ratios in the range 1.83–2.18), [43] as well as one sixth of planetary systems characterized by Doppler spectroscopy (with in this case a narrower period ratio range). [69] Due to incomplete knowledge of the systems, the actual proportions are likely to be higher. [43] Overall, about a third of radial velocity characterized systems appear to have a pair of planets close to a commensurability. [43] [69] It is much more common for pairs of planets to have orbital period ratios a few percent larger than a mean-motion resonance ratio than a few percent smaller (particularly in the case of first order resonances, in which the integers in the ratio differ by one). [43] This was predicted to be true in cases where tidal interactions with the star are significant. [70]

Coincidental 'near' ratios of mean motion

Depiction of asteroid Pallas' 18:7 near resonance with Jupiter in a rotating frame (click for animation). Jupiter (pink loop at upper left) is held nearly stationary. The shift in Pallas' orbital alignment relative to Jupiter increases steadily over time; it never reverses course (i.e., there is no libration). PallasJupiter.GIF
Depiction of asteroid Pallas' 18:7 near resonance with Jupiter in a rotating frame (click for animation). Jupiter (pink loop at upper left) is held nearly stationary. The shift in Pallas' orbital alignment relative to Jupiter increases steadily over time; it never reverses course (i.e., there is no libration).
Depiction of the Earth:Venus 8:13 near resonance. With Earth held stationary at the center of a nonrotating frame, the successive inferior conjunctions of Venus over eight Earth years trace a pentagrammic pattern (reflecting the difference between the numbers in the ratio). Venus pentagram.png
Depiction of the Earth:Venus 8:13 near resonance. With Earth held stationary at the center of a nonrotating frame, the successive inferior conjunctions of Venus over eight Earth years trace a pentagrammic pattern (reflecting the difference between the numbers in the ratio).
Diagram of the orbits of Pluto's small outer four moons, which follow a 3:4:5:6 sequence of near resonances relative to the period of its large inner satellite Charon. The moons Styx, Nix and Hydra are also involved in a true 3-body resonance. Moons of Pluto.png
Diagram of the orbits of Pluto's small outer four moons, which follow a 3:4:5:6 sequence of near resonances relative to the period of its large inner satellite Charon. The moons Styx, Nix and Hydra are also involved in a true 3-body resonance.

A number of near-integer-ratio relationships between the orbital frequencies of the planets or major moons are sometimes pointed out (see list below). However, these have no dynamical significance because there is no appropriate precession of perihelion or other libration to make the resonance perfect (see the detailed discussion in the section above). Such near resonances are dynamically insignificant even if the mismatch is quite small because (unlike a true resonance), after each cycle the relative position of the bodies shifts. When averaged over astronomically short timescales, their relative position is random, just like bodies that are nowhere near resonance. For example, consider the orbits of Earth and Venus, which arrive at almost the same configuration after 8 Earth orbits and 13 Venus orbits. The actual ratio is 0.61518624, which is only 0.032% away from exactly 8:13. The mismatch after 8 years is only 1.5° of Venus' orbital movement. Still, this is enough that Venus and Earth find themselves in the opposite relative orientation to the original every 120 such cycles, which is 960 years. Therefore, on timescales of thousands of years or more (still tiny by astronomical standards), their relative position is effectively random.

The presence of a near resonance may reflect that a perfect resonance existed in the past, or that the system is evolving towards one in the future.

Some orbital frequency coincidences include:

Table of some orbital frequency coincidences in the Solar system
RatioBodiesMismatch
after one
cycle [lower-alpha 1]
Randmztn.
time [lower-alpha 2]
Probability [lower-alpha 3] [lower-alpha 4]
Trans-planetary resonances
9:23 VenusMercury 4.0°200 y 19%
1:4Earth-Mercury54.8°3 y0.3%
8:13 EarthVenus [71] [72] [lower-alpha 5] 1.5°1000 y 6.5%
243:395 EarthVenus [71] [73] 0.8°50,000 y 68%
1:3 MarsVenus 20.6°20 y11%
1:2 MarsEarth 42.9°8 y24%
193:363Mars-Earth0.9°70,000 y0.6%
1:12 JupiterEarth [lower-alpha 6] 49.1°40 y28%
3:19Jupiter-Mars28.7°200 y0.4%
2:5 SaturnJupiter [lower-alpha 7] 12.8°800 y13%
1:7 UranusJupiter 31.1°500 y18%
7:20 UranusSaturn 5.7°20,000 y20%
5:28 NeptuneSaturn 1.9°80,000 y5.2%
1:2 NeptuneUranus 14.0°2000 y7.8%
Mars' satellite system
1:4 DeimosPhobos [lower-alpha 8] 14.9°0.04 y8.3%
Major asteroids' resonances
1:1 PallasCeres [75] [76] 0.7°1000 y0.39% [lower-alpha 9]
7:18 JupiterPallas [77] 0.10°100,000 y0.4% [lower-alpha 10]
87 Sylvia's satellite system [lower-alpha 11]
17:45 RomulusRemus 0.7°40 y6.7%
Jupiter's satellite system
1:6 IoMetis 0.6°2 y0.31%
3:5 AmaltheaAdrastea 3.9°0.2 y6.4%
3:7 CallistoGanymede [78] 0.7°30 y1.2%
Saturn's satellite system
2:3 EnceladusMimas 33.2°0.04 y33%
2:3 DioneTethys [lower-alpha 12] 36.2°0.07 y0.36
3:5 RheaDione 17.1°0.4 y26%
2:7 TitanRhea 21.0°0.7 y22%
1:5 IapetusTitan 9.2°4 y5.1%
Major centaurs' resonances [lower-alpha 13]
3:4 UranusChariklo 4.5°10,000 y7.3%
Uranus' satellite system
3:5 RosalindCordelia [80] 0.22°4 y0.37%
1:3 UmbrielMiranda [lower-alpha 14] 24.5°0.08 y14%
3:5 UmbrielAriel [lower-alpha 15] 24.2°0.3 y35%
1:2 TitaniaUmbriel 36.3°0.1 y20%
2:3 OberonTitania 33.4°0.4 y34%
Neptune's satellite system
1:20 TritonNaiad 13.5°0.2 y7.5%
1:2 ProteusLarissa [83] [84] 8.4°0.07 y4.7%
5:6 ProteusHippocamp 2.1°1 y5.7%
Pluto's satellite system
1:3 StyxCharon [85] 58.5°0.2 y33%
1:4 NixCharon [85] [86] 39.1°0.3 y22%
1:5 KerberosCharon [85] 9.2°2 y5%
1:6 HydraCharon [85] [86] 6.6°3 y3.7%
Haumea's satellite system
3:8 HiʻiakaNamaka [lower-alpha 16] 42.5°2 y55%
  1. Mismatch in orbital longitude of the inner body, as compared to its position at the beginning of the cycle (with the cycle defined as n orbits of the outer body – see below). Circular orbits are assumed (i.e., precession is ignored).
  2. The randomization time is the amount of time needed for the mismatch from the initial relative longitudinal orbital positions of the bodies to grow to 180°. The listed number is rounded to the nearest first significant digit.
  3. Estimated probability of obtaining by chance an orbital coincidence of equal or smaller mismatch, at least once in n attempts, where n is the integer number of orbits of the outer body per cycle, and the mismatch is assumed to randomly vary between 0° and 180°. The value is calculated as 1 − ( 1 −  mismatch / 180° )n . This is a crude calculation that only attempts to give a rough idea of relative probabilities.
  4. Smaller is better: The smaller the probability of an apparently resonant relationship arising as a mere chance alignment of random numbers, the more credible the proposal that gravitational interaction causes persistence of the relationship, or prolongs it / delays its ultimate dissolution by other, disruptive perturbations.
  5. The two near commensurabilities listed for Earth and Venus are reflected in the timing of transits of Venus, which occur in pairs 8 years apart, in a cycle that repeats every 243 years. [71] [73]
  6. The near 1:12 resonance between Jupiter and Earth has the coincidental side-effect of making the Alinda asteroids, which occupy (or are close to) the 3:1 resonance with Jupiter, to be close to a 1:4 resonance with Earth.
  7. The long-known near resonance between Jupiter and Saturn has traditionally been called the Great Inequality . It was first described by Laplace in a series of papers published 1784–1789.
  8. Resonances with a now-vanished inner moon are likely to have been involved in the formation of Phobos and Deimos. [74]
  9. Based on the proper orbital periods, 1684.869 and 1681.601 days, for Pallas and Ceres, respectively.
  10. Based on the "proper" orbital period of Pallas, 1684.869 days, and 4332.59 days for Jupiter.
  11. 87 Sylvia is the first asteroid discovered to have more than one moon.
  12. This resonance may have been occupied in the past. [79]
  13. Some definitions of centaurs require that they not be resonant.
  14. This resonance may have been occupied in the past. [81]
  15. This resonance may have been occupied in the past. [82]
  16. The results for the Haumea system aren't very meaningful because, contrary to the assumptions implicit in the calculations, Namaka has an eccentric, non-Keplerian orbit that precesses rapidly (see below). Hiʻiaka and Namaka are much closer to a 3:8 resonance than indicated, and may actually be in it. [87]

The least probable orbital correlation in the list – meaning the relationship that seems most likely to have not just be by random chance – is that between Io and Metis, followed by those between Rosalind and Cordelia, Pallas and Ceres, Jupiter and Pallas, Callisto and Ganymede, and Hydra and Charon, respectively.

Possible past mean-motion resonances

A past resonance between Jupiter and Saturn may have played a dramatic role in early Solar System history. A 2004 computer model by Alessandro Morbidelli of the Observatoire de la Côte d'Azur in Nice suggested the formation of a 1:2 resonance between Jupiter and Saturn due to interactions with planetesimals that caused them to migrate inward and outward, respectively. In the model, this created a gravitational push that propelled both Uranus and Neptune into higher orbits, and in some scenarios caused them to switch places, which would have doubled Neptune's distance from the Sun. The resultant expulsion of objects from the proto-Kuiper belt as Neptune moved outwards could explain the Late Heavy Bombardment 600 million years after the Solar System's formation and the origin of Jupiter's Trojan asteroids. [88] An outward migration of Neptune could also explain the current occupancy of some of its resonances (particularly the 2:5 resonance) within the Kuiper belt.

While Saturn's mid-sized moons Dione and Tethys are not close to an exact resonance now, they may have been in a 2:3 resonance early in the Solar System's history. This would have led to orbital eccentricity and tidal heating that may have warmed Tethys' interior enough to form a subsurface ocean. Subsequent freezing of the ocean after the moons escaped from the resonance may have generated the extensional stresses that created the enormous graben system of Ithaca Chasma on Tethys. [79]

The satellite system of Uranus is notably different from those of Jupiter and Saturn in that it lacks precise resonances among the larger moons, while the majority of the larger moons of Jupiter (3 of the 4 largest) and of Saturn (6 of the 8 largest) are in mean-motion resonances. In all three satellite systems, moons were likely captured into mean-motion resonances in the past as their orbits shifted due to tidal dissipation, a process by which satellites gain orbital energy at the expense of the primary's rotational energy, affecting inner moons disproportionately. In the Uranian system, however, due to the planet's lesser degree of oblateness, and the larger relative size of its satellites, escape from a mean-motion resonance is much easier. Lower oblateness of the primary alters its gravitational field in such a way that different possible resonances are spaced more closely together. A larger relative satellite size increases the strength of their interactions. Both factors lead to more chaotic orbital behavior at or near mean-motion resonances. Escape from a resonance may be associated with capture into a secondary resonance, and/or tidal evolution-driven increases in orbital eccentricity or inclination.

Mean-motion resonances that probably once existed in the Uranus System include (3:5) Ariel-Miranda, (1:3) Umbriel-Miranda, (3:5) Umbriel-Ariel, and (1:4) Titania-Ariel. [82] [81] Evidence for such past resonances includes the relatively high eccentricities of the orbits of Uranus' inner satellites, and the anomalously high orbital inclination of Miranda. High past orbital eccentricities associated with the (1:3) Umbriel-Miranda and (1:4) Titania-Ariel resonances may have led to tidal heating of the interiors of Miranda and Ariel, [89] respectively. Miranda probably escaped from its resonance with Umbriel via a secondary resonance, and the mechanism of this escape is believed to explain why its orbital inclination is more than 10 times those of the other regular Uranian moons (see Uranus' natural satellites). [90] [91]

Similar to the case of Miranda, the present inclinations of Jupiter's moonlets Amalthea and Thebe are thought to be indications of past passage through the 3:1 and 4:2 resonances with Io, respectively. [92]

Neptune's regular moons Proteus and Larissa are thought to have passed through a 1:2 resonance a few hundred million years ago; the moons have drifted away from each other since then because Proteus is outside a synchronous orbit and Larissa is within one. Passage through the resonance is thought to have excited both moons' eccentricities to a degree that has not since been entirely damped out. [83] [84]

In the case of Pluto's satellites, it has been proposed that the present near resonances are relics of a previous precise resonance that was disrupted by tidal damping of the eccentricity of Charon's orbit (see Pluto's natural satellites for details). The near resonances may be maintained by a 15% local fluctuation in the Pluto-Charon gravitational field. Thus, these near resonances may not be coincidental.

The smaller inner moon of the dwarf planet Haumea, Namaka, is one tenth the mass of the larger outer moon, Hiʻiaka. Namaka revolves around Haumea in 18 days in an eccentric, non-Keplerian orbit, and as of 2008 is inclined 13° from Hiʻiaka. [87] Over the timescale of the system, it should have been tidally damped into a more circular orbit. It appears that it has been disturbed by resonances with the more massive Hiʻiaka, due to converging orbits as it moved outward from Haumea because of tidal dissipation. The moons may have been caught in and then escaped from orbital resonance several times. They probably passed through the 3:1 resonance relatively recently, and currently are in or at least close to an 8:3 resonance. Namaka's orbit is strongly perturbed, with a current precession of about −6.5° per year. [87]

See also

Notes

  1. The nature of this resonance (ignoring subtleties like libration and precession) can be crudely obtained from the orbital periods as follows. From Showalter et al., 2019, [35] the periods of Naiad (Pn) and Thalassa (Pt) are 0.294396 and 0.311484 days, respectively. From these, the period between conjunctions can be calculated as 5.366 days (1/[1/Pn – 1/Pt]), which is 18.23 (≈ 18.25) orbits of Naiad and 17.23 (≈ 17.25) orbits of Thalassa. Thus, after four conjunction periods, 73 orbits of Naiad and 69 orbits of Thalassa have elapsed, and the original configuration will be restored.

Related Research Articles

<span class="mw-page-title-main">Kuiper belt</span> Area of the Solar System beyond the planets, comprising small bodies

The Kuiper belt is a circumstellar disc in the outer Solar System, extending from the orbit of Neptune at 30 astronomical units (AU) to approximately 50 AU from the Sun. It is similar to the asteroid belt, but is far larger—20 times as wide and 20–200 times as massive. Like the asteroid belt, it consists mainly of small bodies or remnants from when the Solar System formed. While many asteroids are composed primarily of rock and metal, most Kuiper belt objects are composed largely of frozen volatiles, such as methane, ammonia, and water. The Kuiper belt is home to most of the objects that astronomers generally accept as dwarf planets: Orcus, Pluto, Haumea, Quaoar, and Makemake. Some of the Solar System's moons, such as Neptune's Triton and Saturn's Phoebe, may have originated in the region.

<span class="mw-page-title-main">Planet</span> Large, round non-stellar astronomical object

A planet is a large, rounded astronomical body that is neither a star nor its remnant. The best available theory of planet formation is the nebular hypothesis, which posits that an interstellar cloud collapses out of a nebula to create a young protostar orbited by a protoplanetary disk. Planets grow in this disk by the gradual accumulation of material driven by gravity, a process called accretion. The Solar System has at least eight planets: the terrestrial planets Mercury, Venus, Earth, and Mars, and the giant planets Jupiter, Saturn, Uranus, and Neptune.

<span class="mw-page-title-main">Planets beyond Neptune</span> Hypothetical planets further than Neptune

Following the discovery of the planet Neptune in 1846, there was considerable speculation that another planet might exist beyond its orbit. The search began in the mid-19th century and continued at the start of the 20th with Percival Lowell's quest for Planet X. Lowell proposed the Planet X hypothesis to explain apparent discrepancies in the orbits of the giant planets, particularly Uranus and Neptune, speculating that the gravity of a large unseen ninth planet could have perturbed Uranus enough to account for the irregularities.

In astronomy, the plutinos are a dynamical group of trans-Neptunian objects that orbit in 2:3 mean-motion resonance with Neptune. This means that for every two orbits a plutino makes, Neptune orbits three times. The dwarf planet Pluto is the largest member as well as the namesake of this group. The next largest members are Orcus, (208996) 2003 AZ84, and Ixion. Plutinos are named after mythological creatures associated with the underworld.

<span class="mw-page-title-main">Quasi-satellite</span> Type of satellite in sync with another orbit

A quasi-satellite is an object in a specific type of co-orbital configuration with a planet where the object stays close to that planet over many orbital periods.

<span class="mw-page-title-main">Planetary migration</span> Astronomical phenomenon

Planetary migration occurs when a planet or other body in orbit around a star interacts with a disk of gas or planetesimals, resulting in the alteration of its orbital parameters, especially its semi-major axis. Planetary migration is the most likely explanation for hot Jupiters. The generally accepted theory of planet formation from a protoplanetary disk predicts that such planets cannot form so close to their stars, as there is insufficient mass at such small radii and the temperature is too high to allow the formation of rocky or icy planetesimals.

In astronomy, a resonant trans-Neptunian object is a trans-Neptunian object (TNO) in mean-motion orbital resonance with Neptune. The orbital periods of the resonant objects are in a simple integer relations with the period of Neptune, e.g. 1:2, 2:3, etc. Resonant TNOs can be either part of the main Kuiper belt population, or the more distant scattered disc population.

<span class="mw-page-title-main">Scattered disc</span> Collection of bodies in the extreme Solar System

The scattered disc (or scattered disk) is a distant circumstellar disc in the Solar System that is sparsely populated by icy small Solar System bodies, which are a subset of the broader family of trans-Neptunian objects. The scattered-disc objects (SDOs) have orbital eccentricities ranging as high as 0.8, inclinations as high as 40°, and perihelia greater than 30 astronomical units (4.5×109 km; 2.8×109 mi). These extreme orbits are thought to be the result of gravitational "scattering" by the gas giants, and the objects continue to be subject to perturbation by the planet Neptune.

<span class="mw-page-title-main">Moons of Pluto</span> Natural satellites orbiting Pluto

The dwarf planet Pluto has five natural satellites. In order of distance from Pluto, they are Charon, Styx, Nix, Kerberos, and Hydra. Charon, the largest, is mutually tidally locked with Pluto, and is massive enough that Pluto and Charon are sometimes considered a binary dwarf planet.

<span class="mw-page-title-main">Formation and evolution of the Solar System</span> Modelling its structure and composition

There is evidence that the formation of the Solar System began about 4.6 billion years ago with the gravitational collapse of a small part of a giant molecular cloud. Most of the collapsing mass collected in the center, forming the Sun, while the rest flattened into a protoplanetary disk out of which the planets, moons, asteroids, and other small Solar System bodies formed.

In astronomy, a co-orbital configuration is a configuration of two or more astronomical objects orbiting at the same, or very similar, distance from their primary; i.e., they are in a 1:1 mean-motion resonance..

The stability of the Solar System is a subject of much inquiry in astronomy. Though the planets have historically been stable as observed, and will be in the "short" term, their weak gravitational effects on one another can add up in ways that are not predictable by any simple means.

<span class="mw-page-title-main">Nice model</span> Scenario for the dynamical evolution of the Solar System

The Nicemodel is a scenario for the dynamical evolution of the Solar System. It is named for the location of the Côte d'Azur Observatory—where it was initially developed in 2005—in Nice, France. It proposes the migration of the giant planets from an initial compact configuration into their present positions, long after the dissipation of the initial protoplanetary disk. In this way, it differs from earlier models of the Solar System's formation. This planetary migration is used in dynamical simulations of the Solar System to explain historical events including the Late Heavy Bombardment of the inner Solar System, the formation of the Oort cloud, and the existence of populations of small Solar System bodies such as the Kuiper belt, the Neptune and Jupiter trojans, and the numerous resonant trans-Neptunian objects dominated by Neptune.

<span class="mw-page-title-main">Regular moon</span> Satellites which formed around their parent planet

In astronomy, a regular moon or a regular satellite is a natural satellite following a relatively close, stable, and circular orbit which is generally aligned to its primary's equator. They form within discs of debris and gas that once surrounded their primary, usually the aftermath of a large collision or leftover material accumulated from the protoplanetary disc. Young regular moons then begin to accumulate material within the circumplanetary disc in a process similar to planetary accretion, as opposed to irregular moons, which formed independently before being captured into orbit around the primary.

The five-planet Nice model is a numerical model of the early Solar System that is a revised variation of the Nice model. It begins with five giant planets, the four that exist today plus an additional ice giant between Saturn and Uranus in a chain of mean-motion resonances.

The jumping-Jupiter scenario specifies an evolution of giant-planet migration described by the Nice model, in which an ice giant is scattered inward by Saturn and outward by Jupiter, causing their semi-major axes to jump, and thereby quickly separating their orbits. The jumping-Jupiter scenario was proposed by Ramon Brasser, Alessandro Morbidelli, Rodney Gomes, Kleomenis Tsiganis, and Harold Levison after their studies revealed that the smooth divergent migration of Jupiter and Saturn resulted in an inner Solar System significantly different from the current Solar System. During this migration secular resonances swept through the inner Solar System exciting the orbits of the terrestrial planets and the asteroids, leaving the planets' orbits too eccentric, and the asteroid belt with too many high-inclination objects. The jumps in the semi-major axes of Jupiter and Saturn described in the jumping-Jupiter scenario can allow these resonances to quickly cross the inner Solar System without altering orbits excessively, although the terrestrial planets remain sensitive to its passage.

The Nice 2 model is a model of the early evolution of the Solar System. The Nice 2 model resembles the original Nice model in that a late instability of the outer Solar System results in gravitational encounters between planets, the disruption of an outer planetesimal disk, and the migrations of the outer planets to new orbits. However, the Nice 2 model differs in its initial conditions and in the mechanism for triggering the late instability. These changes reflect the analysis of the orbital evolution of the outer Solar System during the gas disk phase and the inclusion of gravitational interactions between planetesimals in the outer disk into the model.

<span class="mw-page-title-main">Grand tack hypothesis</span> Theory of early changes in Jupiters orbit

In planetary astronomy, the grand tack hypothesis proposes that Jupiter formed at a distance of 3.5 AU from the Sun, then migrated inward to 1.5 AU, before reversing course due to capturing Saturn in an orbital resonance, eventually halting near its current orbit at 5.2 AU. The reversal of Jupiter's planetary migration is likened to the path of a sailboat changing directions (tacking) as it travels against the wind.

<span class="mw-page-title-main">Satellite system (astronomy)</span> Set of gravitationally bound objects in orbit

A satellite system is a set of gravitationally bound objects in orbit around a planetary mass object or minor planet, or its barycenter. Generally speaking, it is a set of natural satellites (moons), although such systems may also consist of bodies such as circumplanetary disks, ring systems, moonlets, minor-planet moons and artificial satellites any of which may themselves have satellite systems of their own. Some bodies also possess quasi-satellites that have orbits gravitationally influenced by their primary, but are generally not considered to be part of a satellite system. Satellite systems can have complex interactions including magnetic, tidal, atmospheric and orbital interactions such as orbital resonances and libration. Individually major satellite objects are designated in Roman numerals. Satellite systems are referred to either by the possessive adjectives of their primary, or less commonly by the name of their primary. Where only one satellite is known, or it is a binary with a common centre of gravity, it may be referred to using the hyphenated names of the primary and major satellite.

The hypothetical Planet Nine would modify the orbits of extreme trans-Neptunian objects via a combination of effects. On very long timescales exchanges of angular momentum with Planet Nine cause the perihelia of anti-aligned objects to rise until their precession reverses direction, maintaining their anti-alignment, and later fall, returning them to their original orbits. On shorter timescales mean-motion resonances with Planet Nine provides phase protection, which stabilizes their orbits by slightly altering the objects' semi-major axes, keeping their orbits synchronized with Planet Nine's and preventing close approaches. The inclination of Planet Nine's orbit weakens this protection, resulting in a chaotic variation of semi-major axes as objects hop between resonances. The orbital poles of the objects circle that of the Solar System's Laplace plane, which at large semi-major axes is warped toward the plane of Planet Nine's orbit, causing their poles to be clustered toward one side.

References

  1. "IAU 2006 General Assembly: Resolutions 5 and 6" (PDF). IAU. 24 August 2006. Retrieved 23 June 2009.
  2. 1 2 Porco, C.; Nicholson, P. D.; Borderies, N.; Danielson, G. E.; Goldreich, P.; Holdberg, J. B.; Lane, A. L. (1984). "The eccentric Saturnian ringlets at 1.29Rs and 1.45Rs". Icarus . 60 (1): 1–16. Bibcode:1984Icar...60....1P. doi:10.1016/0019-1035(84)90134-9.
  3. Rosen, P. A.; Lissauer, J. J. (1988). "The Titan −1:0 Nodal Bending Wave in Saturn's Ring C". Science . 241 (4866): 690–694. Bibcode:1988Sci...241..690R. doi:10.1126/science.241.4866.690. PMID   17839081. S2CID   32938282.
  4. Chakrabarti, S. K.; Bhattacharyya, A. (2001). "Constraints on the C ring parameters of Saturn at the Titan -1:0 resonance". Monthly Notices of the Royal Astronomical Society . 326 (2): L23. Bibcode:2001MNRAS.326L..23C. doi: 10.1046/j.1365-8711.2001.04813.x .
  5. Williams, James G.; Benson, G. S. (1971). "Resonances in the Neptune-Pluto System". Astronomical Journal. 76: 167. Bibcode:1971AJ.....76..167W. doi: 10.1086/111100 . S2CID   120122522.
  6. Witze, A. (5 September 2018). "The quest to conquer Earth's space junk problem". Nature. 561 (7721): 24–26. Bibcode:2018Natur.561...24W. doi: 10.1038/d41586-018-06170-1 . PMID   30185967.
  7. Daquin, J.; Rosengren, A. J.; Alessi, E. M.; Deleflie, F.; Valsecchi, G. B.; Rossi, A. (2016). "The dynamical structure of the MEO region: long-term stability, chaos, and transport". Celestial Mechanics and Dynamical Astronomy. 124 (4): 335–366. arXiv: 1507.06170 . Bibcode:2016CeMDA.124..335D. doi:10.1007/s10569-015-9665-9. S2CID   119183742.
  8. Wiegert, P.; Connors, M.; Veillet, C. (30 March 2017). "A retrograde co-orbital asteroid of Jupiter". Nature. 543 (7647): 687–689. Bibcode:2017Natur.543..687W. doi:10.1038/nature22029. PMID   28358083. S2CID   205255113.
  9. 1 2 Morais, M. H. M.; Namouni, F. (21 September 2013). "Asteroids in retrograde resonance with Jupiter and Saturn". Monthly Notices of the Royal Astronomical Society Letters . 436: L30–L34. arXiv: 1308.0216 . Bibcode:2013MNRAS.436L..30M. doi:10.1093/mnrasl/slt106. S2CID   119263066.
  10. Morais, Maria Helena Moreira; Namouni, Fathi (12 October 2013). "Retrograde resonance in the planar three-body problem". Celestial Mechanics and Dynamical Astronomy. 117 (4): 405–421. arXiv: 1305.0016 . Bibcode:2013CeMDA.117..405M. doi:10.1007/s10569-013-9519-2. ISSN   1572-9478. S2CID   254379849.
  11. Morais, M. H. M.; Nomouni, F. (2017). "First transneptunian object in polar resonance with Neptune". Letters. Monthly Notices of the Royal Astronomical Society. 472 (1): L1–L4. arXiv: 1708.00346 . Bibcode:2017MNRAS.472L...1M. doi:10.1093/mnrasl/slx125.
  12. Barnes, R. (2011). "Laplace Resonance". In Gargaud, M. (ed.). Encyclopedia of Astrobiology. Springer Science+Business Media. pp. 905–906. doi:10.1007/978-3-642-11274-4_864. ISBN   978-3-642-11271-3.
  13. 1 2 3 Rivera, E. J.; Laughlin, G.; Butler, R. P.; Vogt, S. S.; Haghighipour, N.; Meschiari, S. (2010). "The Lick-Carnegie Exoplanet Survey: A Uranus-mass Fourth Planet for GJ 876 in an Extrasolar Laplace Configuration". The Astrophysical Journal . 719 (1): 890–899. arXiv: 1006.4244 . Bibcode:2010ApJ...719..890R. doi:10.1088/0004-637X/719/1/890. S2CID   118707953.
  14. Nelson, B. E.; Robertson, P. M.; Payne, M. J.; Pritchard, S. M.; Deck, K. M.; Ford, E. B.; Wright, J. T.; Isaacson, H. T. (2015). "An empirically derived three-dimensional Laplace resonance in the Gliese 876 planetary system". Monthly Notices of the Royal Astronomical Society. 455 (3): 2484–2499. arXiv: 1504.07995 . doi: 10.1093/mnras/stv2367 .
  15. Marti, J. G.; Giuppone, C. A.; Beauge, C. (2013). "Dynamical analysis of the Gliese-876 Laplace resonance". Monthly Notices of the Royal Astronomical Society . 433 (2): 928–934. arXiv: 1305.6768 . Bibcode:2013MNRAS.433..928M. doi:10.1093/mnras/stt765. S2CID   118643833.
  16. 1 2 3 4 Showalter, M. R.; Hamilton, D. P. (2015). "Resonant interactions and chaotic rotation of Pluto's small moons". Nature . 522 (7554): 45–49. Bibcode:2015Natur.522...45S. doi:10.1038/nature14469. PMID   26040889. S2CID   205243819.
  17. Murray, C. D.; Dermott, S. F. (1999). Solar System Dynamics. Cambridge University Press. p. 17. ISBN   978-0-521-57597-3.
  18. Beatty, J. K. (23 July 2003). "Why Is Saturn Tipsy?". Sky & Telescope . Archived from the original on 3 September 2009. Retrieved 25 February 2009.
  19. Ward, W. R.; Hamilton, D. P. (2004). "Tilting Saturn. I. Analytic Model". The Astronomical Journal . 128 (5): 2501–2509. Bibcode:2004AJ....128.2501W. doi: 10.1086/424533 .
  20. Hamilton, D. P.; Ward, W. R. (2004). "Tilting Saturn. II. Numerical Model". The Astronomical Journal . 128 (5): 2510–2517. Bibcode:2004AJ....128.2510H. doi:10.1086/424534. S2CID   33083447.
  21. Maryame El Moutamid (15 September 2022). "How Saturn got its tilt and its rings". Science. 377 (6612): 1264–1265. Bibcode:2022Sci...377.1264E. doi:10.1126/science.abq3184. PMID   36108002. S2CID   252309068.
  22. Jack Wisdom; et al. (15 September 2022). "Loss of a satellite could explain Saturn's obliquity and young rings". Science. 377 (6612): 1285–1289. Bibcode:2022Sci...377.1285W. doi:10.1126/science.abn1234. PMID   36107998. S2CID   252310492.
  23. Laskar, J. (2008). "Chaotic diffusion in the Solar System". Icarus . 196 (1): 1–15. arXiv: 0802.3371 . Bibcode:2008Icar..196....1L. doi:10.1016/j.icarus.2008.02.017. S2CID   11586168.
  24. Laskar, J.; Gastineau, M. (2009). "Existence of collisional trajectories of Mercury, Mars and Venus with the Earth". Nature . 459 (7248): 817–819. Bibcode:2009Natur.459..817L. doi:10.1038/nature08096. PMID   19516336. S2CID   4416436.
  25. Musotto, S.; Varad, F.; Moore, W.; Schubert, G. (2002). "Numerical Simulations of the Orbits of the Galilean Satellites". Icarus . 159 (2): 500–504. Bibcode:2002Icar..159..500M. doi:10.1006/icar.2002.6939.
  26. Brown, M. E.; Barkume, K. M.; Ragozzine, D.; Schaller, E. L. (2007). "A collisional family of icy objects in the Kuiper belt" (PDF). Nature . 446 (7133): 294–296. Bibcode:2007Natur.446..294B. doi:10.1038/nature05619. PMID   17361177. S2CID   4430027.
  27. Ragozzine, D.; Brown, M. E. (2007). "Candidate members and age estimate of the family of Kuiper Belt object 2003 EL61". The Astronomical Journal . 134 (6): 2160–2167. arXiv: 0709.0328 . Bibcode:2007AJ....134.2160R. doi:10.1086/522334. S2CID   8387493.
  28. Buie, M. W. (24 October 2011). "Orbit Fit and Astrometric record for 225088". SwRI (Space Science Department). Retrieved 14 November 2014.
  29. Sinclair, A. T. (1975). "The Orbital Resonance Amongst the Galilean Satellites of Jupiter". Monthly Notices of the Royal Astronomical Society . 171 (1): 59–72. Bibcode:1975MNRAS.171...59S. doi: 10.1093/mnras/171.1.59 .
  30. Witze, A. (3 June 2015). "Pluto's moons move in synchrony". Nature News . doi:10.1038/nature.2015.17681. S2CID   134519717.
  31. Malhotra, R. (1997). "Pluto's Orbit" . Retrieved 26 March 2007.
  32. Brown, M. E. (23 March 2009). "S/2005 (90482) 1 needs your help". Mike Brown's Planets . Retrieved 25 March 2009.
  33. "NASA Finds Neptune Moons Locked in 'Dance of Avoidance'". Jet Propulsion Laboratory. 14 November 2019. Retrieved 15 November 2019.
  34. Brozović, M.; Showalter, M. R.; Jacobson, R. A.; French, R. S.; Lissauer, J. J.; de Pater, I. (31 October 2019). "Orbits and resonances of the regular moons of Neptune". Icarus. 338 (2): 113462. arXiv: 1910.13612 . doi:10.1016/j.icarus.2019.113462. S2CID   204960799.
  35. Showalter, M. R.; de Pater, I.; Lissauer, J. J.; French, R. S. (2019). "The seventh inner moon of Neptune" (PDF). Nature. 566 (7744): 350–353. Bibcode:2019Natur.566..350S. doi:10.1038/s41586-019-0909-9. PMC   6424524 . PMID   30787452.
  36. 1 2 3 Shale, C. J.; Vanderburg, A. (2017). "Identifying Exoplanets With Deep Learning: A Five Planet Resonant Chain Around Kepler-80 And An Eighth Planet Around Kepler-90" (PDF). The Astrophysical Journal . 155 (2): 94. arXiv: 1712.05044 . Bibcode:2018AJ....155...94S. doi: 10.3847/1538-3881/aa9e09 . S2CID   4535051 . Retrieved 15 December 2017.
  37. 1 2 Luger, R.; Sestovic, M.; Kruse, E.; Grimm, S. L.; Demory, B.-O.; Agol, E.; Bolmont, E.; Fabrycky, D.; Fernandes, C. S.; Van Grootel, V.; Burgasser, A.; Gillon, M.; Ingalls, J. G.; Jehin, E.; Raymond, S. N.; Selsis, F.; Triaud, A. H. M. J.; Barclay, T.; Barentsen, G.; Delrez, L.; de Wit, J.; Foreman-Mackey, D.; Holdsworth, D. L.; Leconte, J.; Lederer, S.; Turbet, M.; Almleaky, Y.; Benkhaldoun, Z.; Magain, P.; Morris, B. (22 May 2017). "A seven-planet resonant chain in TRAPPIST-1". Nature Astronomy. 1 (6): 0129. arXiv: 1703.04166 . Bibcode:2017NatAs...1E.129L. doi:10.1038/s41550-017-0129. S2CID   54770728.
  38. Izidoro, A.; Ogihara, M.; Raymond, S. N.; Morbidelli, A.; Pierens, A.; Bitsch, B.; Cossou, C.; Hersant, F. (2017). "Breaking the chains: hot super-Earth systems from migration and disruption of compact resonant chains". Monthly Notices of the Royal Astronomical Society. 470 (2): 1750–1770. arXiv: 1703.03634 . Bibcode:2017MNRAS.470.1750I. doi:10.1093/mnras/stx1232. S2CID   119493483.
  39. Laughlin, G. (23 June 2010). "A second Laplace resonance". Systemic: Characterizing Planets. Archived from the original on 29 December 2013. Retrieved 30 June 2015.
  40. Marcy, Ge. W.; Butler, R. P.; Fischer, D.; Vogt, S. S.; Lissauer, J. J.; Rivera, E. J. (2001). "A Pair of Resonant Planets Orbiting GJ 876". The Astrophysical Journal . 556 (1): 296–301. Bibcode:2001ApJ...556..296M. doi: 10.1086/321552 .
  41. "Planet Kepler-223 b". Extrasolar Planets Encyclopaedia . Archived from the original on 22 January 2018. Retrieved 21 January 2018.
  42. Beatty, K. (5 March 2011). "Kepler Finds Planets in Tight Dance". Sky and Telescope . Retrieved 16 October 2012.
  43. 1 2 3 4 5 6 Lissauer, J. J.; et al. (2011). "Architecture and dynamics of Kepler's candidate multiple transiting planet systems". The Astrophysical Journal Supplement Series . 197 (1): 1–26. arXiv: 1102.0543 . Bibcode:2011ApJS..197....8L. doi:10.1088/0067-0049/197/1/8. S2CID   43095783.
  44. 1 2 Mills, S. M.; Fabrycky, D. C.; Migaszewski, C.; Ford, E. B.; Petigura, E.; Isaacson, H. (11 May 2016). "A resonant chain of four transiting, sub-Neptune planets". Nature. 533 (7604): 509–512. arXiv: 1612.07376 . Bibcode:2016Natur.533..509M. doi:10.1038/nature17445. PMID   27225123. S2CID   205248546.
  45. Koppes, S. (17 May 2016). "Kepler-223 System: Clues to Planetary Migration". Jet Propulsion Lab . Retrieved 18 May 2016.
  46. MacDonald, M. G.; Ragozzine, D.; Fabrycky, D. C.; Ford, E. B.; Holman, M. J.; Isaacson, H. T.; Lissauer, J. J.; Lopez, E. D.; Mazeh, T. (1 January 2016). "A Dynamical Analysis of the Kepler-80 System of Five Transiting Planets". The Astronomical Journal. 152 (4): 105. arXiv: 1607.07540 . Bibcode:2016AJ....152..105M. doi: 10.3847/0004-6256/152/4/105 . S2CID   119265122.
  47. Leleu, A.; Alibert, Y.; Hara, N. C.; Hooton, M. J.; Wilson, T. G.; Robutel, P.; Delisle, J. -B.; Laskar, J.; Hoyer, S.; Lovis, C.; Bryant, E. M.; Ducrot, E.; Cabrera, J.; Delrez, L.; Acton, J. S.; Adibekyan, V.; Allart, R.; Prieto, Allende; Alonso, R.; Alves, D.; et al. (20 January 2021). "Six transiting planets and a chain of Laplace resonances in TOI-178". Astronomy & Astrophysics. 649: A26. arXiv: 2101.09260 . Bibcode:2021A&A...649A..26L. doi:10.1051/0004-6361/202039767. ISSN   0004-6361. S2CID   231693292.
  48. Gillon, M.; Triaud, A. H. M. J.; Demory, B.-O.; Jehin, E.; Agol, E.; Deck, K. M.; Lederer, S. M.; de Wit, J.; Burdanov, A. (22 February 2017). "Seven temperate terrestrial planets around the nearby ultracool dwarf star TRAPPIST-1". Nature. 542 (7642): 456–460. arXiv: 1703.01424 . Bibcode:2017Natur.542..456G. doi:10.1038/nature21360. PMC   5330437 . PMID   28230125.
  49. Tamayo, D.; Rein, H.; Petrovich, C.; Murray, N. (10 May 2017). "Convergent Migration Renders TRAPPIST-1 Long-lived". The Astrophysical Journal. 840 (2): L19. arXiv: 1704.02957 . Bibcode:2017ApJ...840L..19T. doi: 10.3847/2041-8213/aa70ea . S2CID   119336960.
  50. 1 2 Chang, K. (10 May 2017). "The Harmony That Keeps Trappist-1's 7 Earth-size Worlds From Colliding". The New York Times . Retrieved 26 June 2017.
  51. Carter, J. A.; Agol, E.; Chaplin, W. J.; et al. (21 June 2012). "Kepler-36: A Pair of Planets with Neighboring Orbits and Dissimilar Densities". Science. 337 (6094): 556–559. arXiv: 1206.4718 . Bibcode:2012Sci...337..556C. doi:10.1126/science.1223269. PMID   22722249. S2CID   40245894.
  52. Barclay, T.; et al. (2013). "A sub-Mercury-sized exoplanet". Nature . 494 (7438): 452–454. arXiv: 1305.5587 . Bibcode:2013Natur.494..452B. doi:10.1038/nature11914. PMID   23426260. S2CID   205232792.
  53. Lissauer, J. J.; Marcy, G. W.; Bryson, S. T.; Rowe, J. F.; Jontof-Hutter, D.; Agol, E.; Borucki, W. J.; Carter, J. A.; Ford, E. B.; Gilliland, R. L.; Kolbl, R.; Star, K. M.; Steffen, J. H.; Torres, G. (25 February 2014). "Validation of Kepler's Multiple Planet Candidates. II: Refined Statistical Framework and Descriptions of Systems of Special Interest". The Astrophysical Journal. 784 (1): 44. arXiv: 1402.6352 . Bibcode:2014ApJ...784...44L. doi:10.1088/0004-637X/784/1/44. S2CID   119108651.
  54. Cabrera, J.; Csizmadia, Sz.; Lehmann, H.; Dvorak, R.; Gandolfi, D.; Rauer, H.; Erikson, A.; Dreyer, C.; Eigmüller, Ph.; Hatzes, A. (31 December 2013). "The Planetary System to KIC 11442793: A Compact Analogue to the Solar System". The Astrophysical Journal. 781 (1): 18. arXiv: 1310.6248 . Bibcode:2014ApJ...781...18C. doi:10.1088/0004-637X/781/1/18. S2CID   118875825.
  55. Hands, T. O.; Alexander, R. D. (13 January 2016). "There might be giants: unseen Jupiter-mass planets as sculptors of tightly packed planetary systems". Monthly Notices of the Royal Astronomical Society. 456 (4): 4121–4127. arXiv: 1512.02649 . Bibcode:2016MNRAS.456.4121H. doi:10.1093/mnras/stv2897. S2CID   55175754.
  56. Jenkins, J. S.; Tuomi, M.; Brasser, R.; Ivanyuk, O.; Murgas, F. (2013). "Two Super-Earths Orbiting the Solar Analog HD 41248 on the Edge of a 7:5 Mean Motion Resonance". The Astrophysical Journal . 771 (1): 41. arXiv: 1304.7374 . Bibcode:2013ApJ...771...41J. doi:10.1088/0004-637X/771/1/41. S2CID   14827197.
  57. Christiansen, Jessie L.; Crossfield, Ian J. M.; Barentsen, G.; Lintott, C. J.; Barclay, T.; Simmons, B. D.; Petigura, E.; Schlieder, J. E.; Dressing, C. D.; Vanderburg, A.; Allen, C. (11 January 2018). "The K2-138 System: A Near-resonant Chain of Five Sub-Neptune Planets Discovered by Citizen Scientists". The Astronomical Journal. 155 (2): 57. arXiv: 1801.03874 . Bibcode:2018AJ....155...57C. doi: 10.3847/1538-3881/aa9be0 . S2CID   52971376.
  58. 1 2 Lopez, T. A.; Barros, S. C. C.; Santerne, A.; Deleuil, M.; Adibekyan, V.; Almenara, J.-M.; Armstrong, D. J.; Brugger, B.; Barrado, D.; Bayliss, D.; Boisse, I.; Bonomo, A. S.; Bouchy, F.; Brown, D. J. A.; Carli, E.; Demangeon, O.; Dumusque, X.; Díaz, R. F.; Faria, J. P.; Figueira, P.; Foxell, E.; Giles, H.; Hébrard, G.; Hojjatpanah, S.; Kirk, J.; Lillo-Box, J.; Lovis, C.; Mousis, O.; da Nóbrega, H. J.; Nielsen, L. D.; Neal, J. J.; Osborn, H. P.; Pepe, F.; Pollacco, D.; Santos, N. C.; Sousa, S. G.; Udry, S.; Vigan, A.; Wheatley, P. J. (1 November 2019). "Exoplanet characterisation in the longest known resonant chain: the K2-138 system seen by HARPS". Astronomy & Astrophysics. 631: A90. arXiv: 1909.13527 . Bibcode:2019A&A...631A..90L. doi:10.1051/0004-6361/201936267. S2CID   203593804.
  59. Leleu, Adrien; Coleman, Gavin A. L.; Ataiee, S. (1 November 2019). "Stability of the co-orbital resonance under dissipation – Application to its evolution in protoplanetary discs". Astronomy & Astrophysics. 631: A6. arXiv: 1901.07640 . Bibcode:2019A&A...631A...6L. doi:10.1051/0004-6361/201834486. S2CID   219840769.
  60. "K2-138 System Diagram". jpl.nasa.gov. Retrieved 20 November 2019.
  61. Hardegree-Ullman, K.; Christiansen, J. (January 2019). "K2-138 g: Spitzer Spots a Sixth Sub-Neptune for the Citizen Science System". American Astronomical Society Meeting Abstracts #233. 233: 164.07. Bibcode:2019AAS...23316407H.
  62. "AO-1 Programmes – CHEOPS Guest Observers Programme – Cosmos". cosmos.esa.int. Retrieved 20 November 2019.
  63. Heller, René; Rodenbeck, Kai; Hippke, Michael (1 May 2019). "Transit least-squares survey – I. Discovery and validation of an Earth-sized planet in the four-planet system K2-32 near the 1:2:5:7 resonance". Astronomy & Astrophysics. 625: A31. arXiv: 1904.00651 . Bibcode:2019A&A...625A..31H. doi:10.1051/0004-6361/201935276. ISSN   0004-6361. S2CID   90259349.
  64. David, Trevor J.; Petigura, Erik A.; Luger, Rodrigo; Foreman-Mackey, Daniel; Livingston, John H.; Mamajek, Eric E.; Hillenbrand, Lynne A. (29 October 2019). "Four Newborn Planets Transiting the Young Solar Analog V1298 Tau". The Astrophysical Journal. 885 (1): L12. arXiv: 1910.04563 . Bibcode:2019ApJ...885L..12D. doi: 10.3847/2041-8213/ab4c99 . ISSN   2041-8213. S2CID   204008446.
  65. Hara, N. C.; Bouchy, F.; Stalport, M.; Boisse, I.; Rodrigues, J.; Delisle, J.- B.; Santerne, A.; Henry, G. W.; Arnold, L.; Astudillo-Defru, N.; Borgniet, S. (2020). "The SOPHIE search for northern extrasolar planets. XVII. A compact planetary system in a near 3:2 mean motion resonance chain". Astronomy & Astrophysics. 636: L6. arXiv: 1911.13296 . doi:10.1051/0004-6361/201937254. S2CID   208512859.
  66. Vanderburg, A.; Rowden, P.; Bryson, S.; Coughlin, J.; Batalha, N.; Collins, K.A.; Latham, D.W.; Mullally, S.E.; Colón, K.D.; Henze, C.; Huang, C.X.; Quinn, S.N. (2020). "A Habitable-zone Earth-sized Planet Rescued from False Positive Status". The Astrophysical Journal. 893 (1): L27. arXiv: 2004.06725 . Bibcode:2020ApJ...893L..27V. doi: 10.3847/2041-8213/ab84e5 . S2CID   215768850.
  67. Weiss, L.M.; Fabrycky, D.C.; Agol, E.; Mills, S.M.; Howard, A.W.; Isaacson, H.; Petigura, E.A.; Fulton, B.; Hirsch, L.; Sinukoff, E. (2020). "The Discovery of the Long-Period, Eccentric Planet Kepler-88 d and System Characterization with Radial Velocities and Photodynamical Analysis" (PDF). The Astronomical Journal. 159 (5): 242. arXiv: 1909.02427 . Bibcode:2020AJ....159..242W. doi: 10.3847/1538-3881/ab88ca . S2CID   202539420.
  68. Klesman, Alison (29 November 2023). "'Shocked and delighted': Astronomers find six planets orbiting in resonance". Astronomy Magazine. Retrieved 23 December 2023.
  69. 1 2 Wright, J. T.; Fakhouri, O.; Marcy, G. W.; Han, E.; Feng, Y.; Johnson, J. A.; Howard, A. W.; Fischer, D. A.; Valenti, J. A.; Anderson, J.; Piskunov, N. (2011). "The Exoplanet Orbit Database". Publications of the Astronomical Society of the Pacific . 123 (902): 412–42. arXiv: 1012.5676 . Bibcode:2011PASP..123..412W. doi:10.1086/659427. S2CID   51769219.
  70. Terquem, C.; Papaloizou, J. C. B. (2007). "Migration and the Formation of Systems of Hot Super-Earths and Neptunes". The Astrophysical Journal . 654 (2): 1110–1120. arXiv: astro-ph/0609779 . Bibcode:2007ApJ...654.1110T. doi:10.1086/509497. S2CID   14034512.
  71. 1 2 3 Langford, P.M. (12 March 2012). "Transits of Venus". Astronomical Society of the Channel Island of Guernsey. Archived from the original on 11 January 2012. Retrieved 15 January 2016.
  72. Bazsó, A.; Eybl, V.; Dvorak, R.; Pilat-Lohinger, E.; Lhotka, C. (2010). "A survey of near-mean-motion resonances between Venus and Earth". Celestial Mechanics and Dynamical Astronomy . 107 (1): 63–76. arXiv: 0911.2357 . Bibcode:2010CeMDA.107...63B. doi:10.1007/s10569-010-9266-6. S2CID   117795811.
  73. 1 2 Shortt, D. (22 May 2012). "Some Details About Transits of Venus". The Planetary Society . Retrieved 22 May 2012.
  74. Rosenblatt, P.; Charnoz, S.; Dunseath, K.M.; Terao-Dunseath, M.; Trinh, A.; Hyodo, R.; et al. (4 July 2016). "Accretion of Phobos and Deimos in an extended debris disc stirred by transient moons" (PDF). Nature Geoscience. 9 (8): 581–583. Bibcode:2016NatGe...9..581R. doi:10.1038/ngeo2742. S2CID   133174714.
  75. Goffin, E. (2001). "New determination of the mass of Pallas". Astronomy and Astrophysics . 365 (3): 627–630. Bibcode:2001A&A...365..627G. doi: 10.1051/0004-6361:20000023 .
  76. Kovacevic, A.B. (2012). "Determination of the mass of Ceres based on the most gravitationally efficient close encounters". Monthly Notices of the Royal Astronomical Society . 419 (3): 2725–2736. arXiv: 1109.6455 . Bibcode:2012MNRAS.419.2725K. doi:10.1111/j.1365-2966.2011.19919.x.
  77. Taylor, D. B. (1982). "The secular motion of Pallas". Monthly Notices of the Royal Astronomical Society . 199 (2): 255–265. Bibcode:1982MNRAS.199..255T. doi: 10.1093/mnras/199.2.255 .
  78. Goldreich, P. (1965). "An explanation of the frequent occurrence of commensurable mean motions in the solar system". Monthly Notices of the Royal Astronomical Society . 130 (3): 159–181. Bibcode:1965MNRAS.130..159G. doi: 10.1093/mnras/130.3.159 .
  79. 1 2 Chen, E. M. A.; Nimmo, F. (2008). "Thermal and Orbital Evolution of Tethys as Constrained by Surface Observations" (PDF). Lunar and Planetary Science XXXIX. Lunar and Planetary Institute. #1968. Retrieved 14 March 2008.
  80. Murray, C.D.; Thompson, R.P. (1990). "Orbits of shepherd satellites deduced from the structure of the rings of Uranus". Nature . 348 (6301): 499–502. Bibcode:1990Natur.348..499M. doi:10.1038/348499a0. S2CID   4320268.
  81. 1 2 Tittemore, W. C.; Wisdom, J. (1990). "Tidal evolution of the Uranian satellites: III. Evolution through the Miranda-Umbriel 3:1, Miranda-Ariel 5:3, and Ariel-Umbriel 2:1 mean-motion commensurabilities". Icarus . 85 (2): 394–443. Bibcode:1990Icar...85..394T. doi:10.1016/0019-1035(90)90125-S. hdl: 1721.1/57632 .
  82. 1 2 Tittemore, W. C.; Wisdom, J. (1988). "Tidal Evolution of the Uranian Satellites I. Passage of Ariel and Umbriel through the 5:3 Mean-Motion Commensurability". Icarus . 74 (2): 172–230. Bibcode:1988Icar...74..172T. doi:10.1016/0019-1035(88)90038-3. hdl: 1721.1/57632 .
  83. 1 2 Zhang, K.; Hamilton, D.P. (2007). "Orbital resonances in the inner Neptunian system: I. The 2:1 Proteus–Larissa mean-motion resonance". Icarus . 188 (2): 386–399. Bibcode:2007Icar..188..386Z. doi:10.1016/j.icarus.2006.12.002.
  84. 1 2 Zhang, K.; Hamilton, D.P. (2008). "Orbital resonances in the inner Neptunian system: II. Resonant history of Proteus, Larissa, Galatea, and Despina". Icarus . 193 (1): 267–282. Bibcode:2008Icar..193..267Z. doi:10.1016/j.icarus.2007.08.024.
  85. 1 2 3 4 Matson, J. (11 July 2012). "New moon for Pluto: Hubble Telescope spots a 5th Plutonian satellite". Scientific American . Retrieved 12 July 2012.
  86. 1 2 Ward, W.R.; Canup, R.M. (2006). "Forced resonant migration of Pluto's outer satellites by Charon". Science . 313 (5790): 1107–1109. Bibcode:2006Sci...313.1107W. doi:10.1126/science.1127293. PMID   16825533. S2CID   36703085.
  87. 1 2 3 Ragozzine, D.; Brown, M.E. (2009). "Orbits and masses of the satellites of the dwarf planet Haumea = 2003 EL61". The Astronomical Journal . 137 (6): 4766–4776. arXiv: 0903.4213 . Bibcode:2009AJ....137.4766R. doi:10.1088/0004-6256/137/6/4766. S2CID   15310444.
  88. Hansen, K. (7 June 2004). "Orbital shuffle for early solar system". Geotimes . Retrieved 26 August 2007.
  89. Tittemore, W. C. (1990). "Tidal heating of Ariel". Icarus . 87 (1): 110–139. Bibcode:1990Icar...87..110T. doi:10.1016/0019-1035(90)90024-4.
  90. Tittemore, W. C.; Wisdom, J. (1989). "Tidal Evolution of the Uranian Satellites II. An Explanation of the Anomalously High Orbital Inclination of Miranda" (PDF). Icarus . 78 (1): 63–89. Bibcode:1989Icar...78...63T. doi:10.1016/0019-1035(89)90070-5. hdl: 1721.1/57632 .
  91. Malhotra, R.; Dermott, S. F (1990). "The Role of Secondary Resonances in the Orbital History of Miranda". Icarus . 85 (2): 444–480. Bibcode:1990Icar...85..444M. doi: 10.1016/0019-1035(90)90126-T .
  92. Burns, J. A.; Simonelli, D. P.; Showalter, M. R.; Hamilton, D. P.; Porco, Carolyn C.; Esposito, L. W.; Throop, H. (2004). "Jupiter's Ring-Moon System" (PDF). In Bagenal, Fran; Dowling, Timothy E.; McKinnon, William B. (eds.). Jupiter: The Planet, Satellites and Magnetosphere. Cambridge University Press. ISBN   978-0-521-03545-3.