Accretion (astrophysics)

Last updated
ALMA image of HL Tauri, a protoplanetary disk HL Tau protoplanetary disk.jpg
ALMA image of HL Tauri, a protoplanetary disk

In astrophysics, accretion is the accumulation of particles into a massive object by gravitationally attracting more matter, typically gaseous matter, into an accretion disk. [1] [2] Most astronomical objects, such as galaxies, stars, and planets, are formed by accretion processes.

Contents

Overview

The accretion model that Earth and the other terrestrial planets formed from meteoric material was proposed in 1944 by Otto Schmidt, followed by the protoplanet theory of William McCrea (1960) and finally the capture theory of Michael Woolfson. [3] In 1978, Andrew Prentice resurrected the initial Laplacian ideas about planet formation and developed the modern Laplacian theory. [3] None of these models proved completely successful, and many of the proposed theories were descriptive.

The 1944 accretion model by Otto Schmidt was further developed in a quantitative way in 1969 by Viktor Safronov. [4] He calculated, in detail, the different stages of terrestrial planet formation. [5] [6] Since then, the model has been further developed using intensive numerical simulations to study planetesimal accumulation. It is now accepted that stars form by the gravitational collapse of interstellar gas. Prior to collapse, this gas is mostly in the form of molecular clouds, such as the Orion Nebula. As the cloud collapses, losing potential energy, it heats up, gaining kinetic energy, and the conservation of angular momentum ensures that the cloud forms a flattened disk—the accretion disk.

Accretion of galaxies

A few hundred thousand years after the Big Bang, the Universe cooled to the point where atoms could form. As the Universe continued to expand and cool, the atoms lost enough kinetic energy, and dark matter coalesced sufficiently, to form protogalaxies. As further accretion occurred, galaxies formed. [7] Indirect evidence is widespread. [7] Galaxies grow through mergers and smooth gas accretion. Accretion also occurs inside galaxies, forming stars.

Accretion of stars

The visible-light (left) and infrared (right) views of the Trifid Nebula, a giant star-forming cloud of gas and dust located 5,400 light-years (1,700 pc) away in the constellation Sagittarius Embryonic Stars in the Trifid Nebula.jpg
The visible-light (left) and infrared (right) views of the Trifid Nebula, a giant star-forming cloud of gas and dust located 5,400 light-years (1,700  pc ) away in the constellation Sagittarius

Stars are thought to form inside giant clouds of cold molecular hydrogengiant molecular clouds of roughly 300,000  M and 65 light-years (20  pc ) in diameter. [8] [9] Over millions of years, giant molecular clouds are prone to collapse and fragmentation. [10] These fragments then form small, dense cores, which in turn collapse into stars. [9] The cores range in mass from a fraction to several times that of the Sun and are called protostellar (protosolar) nebulae. [8] They possess diameters of 2,000–20,000 astronomical units (0.01–0.1  pc ) and a particle number density of roughly 10,000 to 100,000/cm3 (160,000 to 1,600,000/cu in). Compare it with the particle number density of the air at the sea level—2.8×1019/cm3 (4.6×1020/cu in). [9] [11]

The initial collapse of a solar-mass protostellar nebula takes around 100,000 years. [8] [9] Every nebula begins with a certain amount of angular momentum. Gas in the central part of the nebula, with relatively low angular momentum, undergoes fast compression and forms a hot hydrostatic (non-contracting) core containing a small fraction of the mass of the original nebula. This core forms the seed of what will become a star. [8] As the collapse continues, conservation of angular momentum dictates that the rotation of the infalling envelope accelerates, which eventually forms a disk.

Infrared image of the molecular outflow from an otherwise hidden newborn star HH 46/47 Embedded Outflow in Herbig-Haro object HH 46 47.jpg
Infrared image of the molecular outflow from an otherwise hidden newborn star HH 46/47

As the infall of material from the disk continues, the envelope eventually becomes thin and transparent and the young stellar object (YSO) becomes observable, initially in far-infrared light and later in the visible. [11] Around this time the protostar begins to fuse deuterium. If the protostar is sufficiently massive (above 80  MJ), hydrogen fusion follows. Otherwise, if its mass is too low, the object becomes a brown dwarf. [12] This birth of a new star occurs approximately 100,000 years after the collapse begins. [8] Objects at this stage are known as Class I protostars, which are also called young T Tauri stars, evolved protostars, or young stellar objects. By this time, the forming star has already accreted much of its mass; the total mass of the disk and remaining envelope does not exceed 10–20% of the mass of the central YSO. [11]

When the lower-mass star in a binary system enters an expansion phase, its outer atmosphere may fall onto the compact star, forming an accretion disk Accretion Disk Binary System.jpg
When the lower-mass star in a binary system enters an expansion phase, its outer atmosphere may fall onto the compact star, forming an accretion disk

At the next stage, the envelope completely disappears, having been gathered up by the disk, and the protostar becomes a classical T Tauri star. [13] The latter have accretion disks and continue to accrete hot gas, which manifests itself by strong emission lines in their spectrum. The former do not possess accretion disks. Classical T Tauri stars evolve into weakly lined T Tauri stars. [14] This happens after about 1 million years. [8] The mass of the disk around a classical T Tauri star is about 1–3% of the stellar mass, and it is accreted at a rate of 10−7 to 10−9 M per year. [15] A pair of bipolar jets is usually present as well. The accretion explains all peculiar properties of classical T Tauri stars: strong flux in the emission lines (up to 100% of the intrinsic luminosity of the star), magnetic activity, photometric variability and jets. [16] The emission lines actually form as the accreted gas hits the "surface" of the star, which happens around its magnetic poles. [16] The jets are byproducts of accretion: they carry away excessive angular momentum. The classical T Tauri stage lasts about 10 million years [8] (there are only a few examples of so-called Peter Pan disks, where the accretion continues to persist for much longer periods, sometimes lasting for more than 40 million years [17] ). The disk eventually disappears due to accretion onto the central star, planet formation, ejection by jets, and photoevaporation by ultraviolet radiation from the central star and nearby stars. [18] As a result, the young star becomes a weakly lined T Tauri star, which, over hundreds of millions of years, evolves into an ordinary Sun-like star, dependent on its initial mass.

Accretion of planets

Artist's impression of a protoplanetary disk showing a young star at its center Artist's Impression of a Baby Star Still Surrounded by a Protoplanetary Disc.jpg
Artist's impression of a protoplanetary disk showing a young star at its center

Self-accretion of cosmic dust accelerates the growth of the particles into boulder-sized planetesimals. The more massive planetesimals accrete some smaller ones, while others shatter in collisions. Accretion disks are common around smaller stars, stellar remnants in a close binary, or black holes surrounded by material (such as those at the centers of galaxies). Some dynamics in the disk, such as dynamical friction, are necessary to allow orbiting gas to lose angular momentum and fall onto the central massive object. Occasionally, this can result in stellar surface fusion (see Bondi accretion).

In the formation of terrestrial planets or planetary cores, several stages can be considered. First, when gas and dust grains collide, they agglomerate by microphysical processes like van der Waals forces and electromagnetic forces, forming micrometer-sized particles. During this stage, accumulation mechanisms are largely non-gravitational in nature. [19] However, planetesimal formation in the centimeter-to-meter range is not well understood, and no convincing explanation is offered as to why such grains would accumulate rather than simply rebound. [19] :341 In particular, it is still not clear how these objects grow to become 0.1–1 km (0.06–0.6 mi) sized planetesimals; [5] [20] this problem is known as the "meter size barrier": [21] [22] As dust particles grow by coagulation, they acquire increasingly large relative velocities with respect to other particles in their vicinity, as well as a systematic inward drift velocity, that leads to destructive collisions, and thereby limit the growth of the aggregates to some maximum size. [23] Ward (1996) suggests that when slow moving grains collide, the very low, yet non-zero, gravity of colliding grains impedes their escape. [19] :341 It is also thought that grain fragmentation plays an important role replenishing small grains and keeping the disk thick, but also in maintaining a relatively high abundance of solids of all sizes. [23]

A number of mechanisms have been proposed for crossing the 'meter-sized' barrier. Local concentrations of pebbles may form, which then gravitationally collapse into planetesimals the size of large asteroids. These concentrations can occur passively due to the structure of the gas disk, for example, between eddies, at pressure bumps, at the edge of a gap created by a giant planet, or at the boundaries of turbulent regions of the disk. [24] Or, the particles may take an active role in their concentration via a feedback mechanism referred to as a streaming instability. In a streaming instability the interaction between the solids and the gas in the protoplanetary disk results in the growth of local concentrations, as new particles accumulate in the wake of small concentrations, causing them to grow into massive filaments. [24] Alternatively, if the grains that form due to the agglomeration of dust are highly porous their growth may continue until they become large enough to collapse due to their own gravity. The low density of these objects allows them to remain strongly coupled with the gas, thereby avoiding high velocity collisions which could result in their erosion or fragmentation. [25]

Grains eventually stick together to form mountain-size (or larger) bodies called planetesimals. Collisions and gravitational interactions between planetesimals combine to produce Moon-size planetary embryos (protoplanets) over roughly 0.1–1 million years. Finally, the planetary embryos collide to form planets over 10–100 million years. [20] The planetesimals are massive enough that mutual gravitational interactions are significant enough to be taken into account when computing their evolution. [5] Growth is aided by orbital decay of smaller bodies due to gas drag, which prevents them from being stranded between orbits of the embryos. [26] [27] Further collisions and accumulation lead to terrestrial planets or the core of giant planets.

If the planetesimals formed via the gravitational collapse of local concentrations of pebbles, their growth into planetary embryos and the cores of giant planets is dominated by the further accretions of pebbles. Pebble accretion is aided by the gas drag felt by objects as they accelerate toward a massive body. Gas drag slows the pebbles below the escape velocity of the massive body causing them to spiral toward and to be accreted by it. Pebble accretion may accelerate the formation of planets by a factor of 1000 compared to the accretion of planetesimals, allowing giant planets to form before the dissipation of the gas disk. [28] [29] Yet, core growth via pebble accretion appears incompatible with the final masses and compositions of Uranus and Neptune. [30]

The formation of terrestrial planets differs from that of giant gas planets, also called Jovian planets. The particles that make up the terrestrial planets are made from metal and rock that condensed in the inner Solar System. However, Jovian planets began as large, icy planetesimals, which then captured hydrogen and helium gas from the solar nebula. [31] Differentiation between these two classes of planetesimals arise due to the frost line of the solar nebula. [32]

Accretion of asteroids

Chondrules in a chondrite meteorite. A millimeter scale is shown. Chondrules grassland 1.jpg
Chondrules in a chondrite meteorite. A millimeter scale is shown.

Meteorites contain a record of accretion and impacts during all stages of asteroid origin and evolution; however, the mechanism of asteroid accretion and growth is not well understood. [33] Evidence suggests the main growth of asteroids can result from gas-assisted accretion of chondrules, which are millimeter-sized spherules that form as molten (or partially molten) droplets in space before being accreted to their parent asteroids. [33] In the inner Solar System, chondrules appear to have been crucial for initiating accretion. [34] The tiny mass of asteroids may be partly due to inefficient chondrule formation beyond 2 AU, or less-efficient delivery of chondrules from near the protostar. [34] Also, impacts controlled the formation and destruction of asteroids, and are thought to be a major factor in their geological evolution. [34]

Chondrules, metal grains, and other components likely formed in the solar nebula. These accreted together to form parent asteroids. Some of these bodies subsequently melted, forming metallic cores and olivine-rich mantles; others were aqueously altered. [34] After the asteroids had cooled, they were eroded by impacts for 4.5 billion years, or disrupted. [35]

For accretion to occur, impact velocities must be less than about twice the escape velocity, which is about 140  m/s (460  ft/s ) for a 100 km (60 mi) radius asteroid. [34] Simple models for accretion in the asteroid belt generally assume micrometer-sized dust grains sticking together and settling to the midplane of the nebula to form a dense layer of dust, which, because of gravitational forces, was converted into a disk of kilometer-sized planetesimals. But, several arguments[ which? ] suggest that asteroids may not have accreted this way. [34]

Accretion of comets

486958 Arrokoth, a Kuiper belt object which is thought to represent the original planetesimals from which the planets grew UltimaThule CA06 color vertical (rotated).png
486958 Arrokoth, a Kuiper belt object which is thought to represent the original planetesimals from which the planets grew

Comets, or their precursors, formed in the outer Solar System, possibly millions of years before planet formation. [36] How and when comets formed is debated, with distinct implications for Solar System formation, dynamics, and geology. Three-dimensional computer simulations indicate the major structural features observed on cometary nuclei can be explained by pairwise low velocity accretion of weak cometesimals. [37] [38] The currently favored formation mechanism is that of the nebular hypothesis, which states that comets are probably a remnant of the original planetesimal "building blocks" from which the planets grew. [39] [40] [41]

Astronomers think that comets originate in both the Oort cloud and the scattered disk. [42] The scattered disk was created when Neptune migrated outward into the proto-Kuiper belt, which at the time was much closer to the Sun, and left in its wake a population of dynamically stable objects that could never be affected by its orbit (the Kuiper belt proper), and a population whose perihelia are close enough that Neptune can still disturb them as it travels around the Sun (the scattered disk). Because the scattered disk is dynamically active and the Kuiper belt relatively dynamically stable, the scattered disk is now seen as the most likely point of origin for periodic comets. [42] The classic Oort cloud theory states that the Oort cloud, a sphere measuring about 50,000 AU (0.24 pc) in radius, formed at the same time as the solar nebula and occasionally releases comets into the inner Solar System as a giant planet or star passes nearby and causes gravitational disruptions. [43] Examples of such comet clouds may already have been seen in the Helix Nebula. [44]

The Rosetta mission to comet 67P/Churyumov–Gerasimenko determined in 2015 that when Sun's heat penetrates the surface, it triggers evaporation (sublimation) of buried ice. While some of the resulting water vapour may escape from the nucleus, 80% of it recondenses in layers beneath the surface. [45] This observation implies that the thin ice-rich layers exposed close to the surface may be a consequence of cometary activity and evolution, and that global layering does not necessarily occur early in the comet's formation history. [45] [46] While most scientists thought that all the evidence indicated that the structure of nuclei of comets is processed rubble piles of smaller ice planetesimals of a previous generation, [47] the Rosetta mission dispelled the idea that comets are "rubble piles" of disparate material. [48] [49] The Rosetta mission confirmed the idea that comets are "rubble piles" of disparate material. [50] [51] Comets appear to have formed as ~100-km bodies, then overwhelmingly ground/recontacted into their present states. [52]

See also

Related Research Articles

<span class="mw-page-title-main">Star formation</span> Process by which dense regions of molecular clouds in interstellar space collapse to form stars

Star formation is the process by which dense regions within molecular clouds in interstellar space, sometimes referred to as "stellar nurseries" or "star-forming regions", collapse and form stars. As a branch of astronomy, star formation includes the study of the interstellar medium (ISM) and giant molecular clouds (GMC) as precursors to the star formation process, and the study of protostars and young stellar objects as its immediate products. It is closely related to planet formation, another branch of astronomy. Star formation theory, as well as accounting for the formation of a single star, must also account for the statistics of binary stars and the initial mass function. Most stars do not form in isolation but as part of a group of stars referred as star clusters or stellar associations.

<span class="mw-page-title-main">Planetesimal</span> Solid objects in protoplanetary disks and debris disks

Planetesimals are solid objects thought to exist in protoplanetary disks and debris disks. Per the Chamberlin–Moulton planetesimal hypothesis, they are believed to form out of cosmic dust grains. Believed to have formed in the Solar System about 4.6 billion years ago, they aid study of its formation.

<span class="mw-page-title-main">Nebular hypothesis</span> Astronomical theory about the Solar System

The nebular hypothesis is the most widely accepted model in the field of cosmogony to explain the formation and evolution of the Solar System. It suggests the Solar System is formed from gas and dust orbiting the Sun which clumped up together to form the planets. The theory was developed by Immanuel Kant and published in his Universal Natural History and Theory of the Heavens (1755) and then modified in 1796 by Pierre Laplace. Originally applied to the Solar System, the process of planetary system formation is now thought to be at work throughout the universe. The widely accepted modern variant of the nebular theory is the solar nebular disk model (SNDM) or solar nebular model. It offered explanations for a variety of properties of the Solar System, including the nearly circular and coplanar orbits of the planets, and their motion in the same direction as the Sun's rotation. Some elements of the original nebular theory are echoed in modern theories of planetary formation, but most elements have been superseded.

<span class="mw-page-title-main">Protoplanetary disk</span> Gas and dust surrounding a newly formed star

A protoplanetary disk is a rotating circumstellar disc of dense gas and dust surrounding a young newly formed star, a T Tauri star, or Herbig Ae/Be star. The protoplanetary disk may also be considered an accretion disk for the star itself, because gases or other material may be falling from the inner edge of the disk onto the surface of the star. This process should not be confused with the accretion process thought to build up the planets themselves. Externally illuminated photo-evaporating protoplanetary disks are called proplyds.

<span class="mw-page-title-main">Proplyd</span> Dust ring surrounding large stars thousands of solar radii wide

A proplyd, short for ionized protoplanetary disk, is an externally illuminated photoevaporating protoplanetary disk around a young star. Nearly 180 proplyds have been discovered in the Orion Nebula. Images of proplyds in other star-forming regions are rare, while Orion is the only region with a large known sample due to its relative proximity to Earth.

<span class="mw-page-title-main">Protoplanet</span> Large planetary embryo

A protoplanet is a large planetary embryo that originated within a protoplanetary disk and has undergone internal melting to produce a differentiated interior. Protoplanets are thought to form out of kilometer-sized planetesimals that gravitationally perturb each other's orbits and collide, gradually coalescing into the dominant planets.

<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.

<span class="mw-page-title-main">Ice giant</span> Giant planet primarily consisting of compounds with freezing points exceeding 100°K

An ice giant is a giant planet composed mainly of elements heavier than hydrogen and helium, such as oxygen, carbon, nitrogen, and sulfur. There are two ice giants in the Solar System: Uranus and Neptune.

In astronomy or planetary science, the frost line, also known as the snow line or ice line, is the minimum distance from the central protostar of a solar nebula where the temperature is low enough for volatile compounds such as water, ammonia, methane, carbon dioxide and carbon monoxide to condense into solid grains, which will allow their accretion into planetesimals. Beyond the line, otherwise gaseous compounds can be quite easily condensed to allow formation of gas and ice giants; while within it, only heavier compounds can be accreted to form the typically much smaller rocky planets.

<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.

<span class="mw-page-title-main">History of Solar System formation and evolution hypotheses</span>

The history of scientific thought about the formation and evolution of the Solar System began with the Copernican Revolution. The first recorded use of the term "Solar System" dates from 1704. Since the seventeenth century, philosophers and scientists have been forming hypotheses concerning the origins of our Solar System and the Moon and attempting to predict how the Solar System would change in the future. René Descartes was the first to hypothesize on the beginning of the Solar System; however, more scientists joined the discussion in the eighteenth century, forming the groundwork for later hypotheses on the topic. Later, particularly in the twentieth century, a variety of hypotheses began to build up, including the now-commonly accepted nebular hypothesis.

<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">Retrograde and prograde motion</span> Relative directions of orbit or rotation

Retrograde motion in astronomy is, in general, orbital or rotational motion of an object in the direction opposite the rotation of its primary, that is, the central object. It may also describe other motions such as precession or nutation of an object's rotational axis. Prograde or direct motion is more normal motion in the same direction as the primary rotates. However, "retrograde" and "prograde" can also refer to an object other than the primary if so described. The direction of rotation is determined by an inertial frame of reference, such as distant fixed stars.

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.

<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">Circumstellar disc</span> Accumulation of matter around a star

A circumstellar disc is a torus, pancake or ring-shaped accretion disk of matter composed of gas, dust, planetesimals, asteroids, or collision fragments in orbit around a star. Around the youngest stars, they are the reservoirs of material out of which planets may form. Around mature stars, they indicate that planetesimal formation has taken place, and around white dwarfs, they indicate that planetary material survived the whole of stellar evolution. Such a disc can manifest itself in various ways.

<span class="mw-page-title-main">Pebble accretion</span>

Pebble accretion is the accumulation of particles, ranging from centimeters up to meters in diameter, into planetesimals in a protoplanetary disk that is enhanced by aerodynamic drag from the gas present in the disk. This drag reduces the relative velocity of pebbles as they pass by larger bodies, preventing some from escaping the body's gravity. These pebbles are then accreted by the body after spiraling or settling toward its surface. This process increases the cross section over which the large bodies can accrete material, accelerating their growth. The rapid growth of the planetesimals via pebble accretion allows for the formation of giant planet cores in the outer Solar System before the dispersal of the gas disk. A reduction in the size of pebbles as they lose water ice after crossing the ice line and a declining density of gas with distance from the sun slow the rates of pebble accretion in the inner Solar System resulting in smaller terrestrial planets, a small mass of Mars and a low mass asteroid belt.

In planetary science a streaming instability is a hypothetical mechanism for the formation of planetesimals in which the drag felt by solid particles orbiting in a gas disk leads to their spontaneous concentration into clumps which can gravitationally collapse. Small initial clumps increase the orbital velocity of the gas, slowing radial drift locally, leading to their growth as they are joined by faster drifting isolated particles. Massive filaments form that reach densities sufficient for the gravitational collapse into planetesimals the size of large asteroids, bypassing a number of barriers to the traditional formation mechanisms. The formation of streaming instabilities requires solids that are moderately coupled to the gas and a local solid to gas ratio of one or greater. The growth of solids large enough to become moderately coupled to the gas is more likely outside the ice line and in regions with limited turbulence. An initial concentration of solids with respect to the gas is necessary to suppress turbulence sufficiently to allow the solid to gas ratio to reach greater than one at the mid-plane. A wide variety of mechanisms to selectively remove gas or to concentrate solids have been proposed. In the inner Solar System the formation of streaming instabilities requires a greater initial concentration of solids or the growth of solid beyond the size of chondrules.

<span class="mw-page-title-main">Dust astronomy</span> Branch of astronomy

Dust astronomy is a subfield of astronomy that uses the information contained in individual cosmic dust particles ranging from their dynamical state to its isotopic, elemental, molecular, and mineralogical composition in order to obtain information on the astronomical objects occurring in outer space. Dust astronomy overlaps with the fields of Planetary science, Cosmochemistry, and Astrobiology.

References

  1. "Science with the VLTI". European Southern Observatory. 8 August 2008. Archived from the original on 24 May 2011. Retrieved 11 April 2011.
  2. Masters, Harris (26 August 2010). "Transcript of The Accretion of Galaxies and Stars". Prezi. Retrieved 8 January 2016.
  3. 1 2 Woolfson, M. M. (March 1993). "The Solar System – its Origin and Evolution". Quarterly Journal of the Royal Astronomical Society. 34: 1–20. Bibcode:1993QJRAS..34....1W.
    For details of Kant's position, see Palmquist, Stephen (September 1987). "Kant's Cosmogony Re-evaluated". Studies in History and Philosophy of Science. 18 (3): 255–269. Bibcode:1987SHPS...18..255P. doi:10.1016/0039-3681(87)90021-5.
  4. Henbest, Nigel (24 August 1991). "Birth of the planets: The Earth and its fellow planets may be survivors from a time when planets ricocheted around the Sun like ball bearings on a pinball table". New Scientist . Retrieved 18 April 2008.
  5. 1 2 3 Papaloizou, John C. B.; Terquem, Caroline (28 November 2005). "Planet formation and migration" (PDF). CERN . Retrieved 21 October 2015.
  6. Safronov, Viktor S. (1972) [1969]. Evolution of the Protoplanetary Cloud and Formation of the Earth and the Planets. Jerusalem: Israel Program for Scientific Translations. hdl:2027/uc1.b4387676. ISBN   0-7065-1225-1. NASA Technical Translation F-677.
  7. 1 2 Kereš, Dušan; Davé, Romeel; Fardal, Mark; Faucher-Giguere, C.-A.; Hernquist, Lars; et al. (2010). Gas Accretion in Galaxies (PDF). Massive Galaxies Over Cosmic Time 3. 8–10 November 2010. Tucson, Arizona. National Optical Astronomy Observatory.
  8. 1 2 3 4 5 6 7 Montmerle, Thierry; Augereau, Jean-Charles; Chaussidon, Marc; Counelle, Mathieu; Marty, Bernard; et al. (June 2006). "Solar System Formation and Early Evolution: the First 100 Million Years". Earth, Moon, and Planets. 98 (1–4): 39–95. Bibcode:2006EM&P...98...39M. doi:10.1007/s11038-006-9087-5. S2CID   120504344.
  9. 1 2 3 4 Pudritz, Ralph E. (January 2002). "Clustered Star Formation and the Origin of Stellar Masses". Science . 295 (5552): 68–75. Bibcode:2002Sci...295...68P. doi:10.1126/science.1068298. PMID   11778037. S2CID   33585808.
  10. Clark, Paul C.; Bonnell, Ian A. (July 2005). "The onset of collapse in turbulently supported molecular clouds". Monthly Notices of the Royal Astronomical Society. 361 (1): 2–16. Bibcode:2005MNRAS.361....2C. doi: 10.1111/j.1365-2966.2005.09105.x .
  11. 1 2 3 Motte, F.; Andre, P.; Neri, R. (August 1998). "The initial conditions of star formation in the ρ Ophiuchi main cloud: wide-field millimeter continuum mapping". Astronomy and Astrophysics. 336: 150–172. Bibcode:1998A&A...336..150M.
  12. Stahler, Steven W. (September 1988). "Deuterium and the Stellar Birthline". The Astrophysical Journal. 332: 804–825. Bibcode:1988ApJ...332..804S. doi:10.1086/166694.
  13. Mohanty, Subhanjoy; Jayawardhana, Ray; Basri, Gibor (June 2005). "The T Tauri Phase down to Nearly Planetary Masses: Echelle Spectra of 82 Very Low Mass Stars and Brown Dwarfs". The Astrophysical Journal. 626 (1): 498–522. arXiv: astro-ph/0502155 . Bibcode:2005ApJ...626..498M. doi:10.1086/429794. S2CID   8462683.
  14. Martin, E. L.; Rebolo, R.; Magazzu, A.; Pavlenko, Ya. V. (February 1994). "Pre-main sequence lithium burning". Astronomy and Astrophysics. 282: 503–517. arXiv: astro-ph/9308047 . Bibcode:1994A&A...282..503M.
  15. Hartmann, Lee; Calvet, Nuria; Gullbring, Eric; D'Alessio, Paula (March 1998). "Accretion and the evolution of T Tauri disks". The Astrophysical Journal. 495 (1): 385–400. Bibcode:1998ApJ...495..385H. doi: 10.1086/305277 .
  16. 1 2 Muzerolle, James; Calvet, Nuria; Hartmann, Lee (April 2001). "Emission-line diagnostics of T Tauri magnetospheric accretion. II. Improved model tests and insights into accretion physics". The Astrophysical Journal. 550 (2): 944–961. Bibcode:2001ApJ...550..944M. doi: 10.1086/319779 .
  17. Silverberg, Steven M.; Wisniewski, John P.; Kuchner, Marc J.; Lawson, Kellen D.; Bans, Alissa S.; Debes, John H.; Biggs, Joseph R.; Bosch, Milton K. D.; Doll, Katharina; Luca, Hugo A. Durantini; Enachioaie, Alexandru (14 January 2020). "Peter Pan Disks: Long-lived Accretion Disks Around Young M Stars". The Astrophysical Journal. 890 (2): 106. arXiv: 2001.05030 . Bibcode:2020ApJ...890..106S. doi: 10.3847/1538-4357/ab68e6 . S2CID   210718358.
  18. Adams, Fred C.; Hollenbach, David; Laughlin, Gregory; Gorti, Uma (August 2004). "Photoevaporation of circumstellar disks due to external far-ultraviolet radiation in stellar aggregates". The Astrophysical Journal. 611 (1): 360–379. arXiv: astro-ph/0404383 . Bibcode:2004ApJ...611..360A. doi:10.1086/421989. S2CID   16093937.
  19. 1 2 3 Ward, William R. (1996). "Planetary Accretion". ASP Conference Series. Completing the Inventory of the Solar System. 107: 337–361. Bibcode:1996ASPC..107..337W.
  20. 1 2 Chambers, John E. (July 2004). "Planetary accretion in the inner Solar System". Earth and Planetary Science Letters. 233 (3–4): 241–252. Bibcode:2004E&PSL.223..241C. doi:10.1016/j.epsl.2004.04.031.
  21. Küffmeier, Michael (3 April 2015). "What is the meter size barrier?". Astrobites. Retrieved 15 January 2015.
  22. Grishin, Evgeni; et al. (August 2019). "Planet seeding through gas-assisted capture of interstellar objects". Monthly Notices of the Royal Astronomical Society. 487 (3): 3324–3332. arXiv: 1804.09716 . Bibcode:2019MNRAS.487.3324G. doi:10.1093/mnras/stz1505.
  23. 1 2 Birnstiel, T.; Dullemond, C. P.; Brauer, F. (August 2009). "Dust retention in protoplanetary disks". Astronomy and Astrophysics. 503 (1): L5–L8. arXiv: 0907.0985 . Bibcode:2009A&A...503L...5B. doi:10.1051/0004-6361/200912452. S2CID   12932274.
  24. 1 2 Johansen, A.; Blum, J.; Tanaka, H.; Ormel, C.; Bizzarro, M.; Rickman, H. (2014). "The Multifaceted Planetesimal Formation Process". In Beuther, H.; Klessen, R. S.; Dullemond, C. P.; Henning, T. (eds.). Protostars and Planets VI. University of Arizona Press. pp. 547–570. arXiv: 1402.1344 . Bibcode:2014prpl.conf..547J. doi:10.2458/azu_uapress_9780816531240-ch024. ISBN   978-0-8165-3124-0. S2CID   119300087.
  25. Johansen, A.; Jacquet, E.; Cuzzi, J. N.; Morbidelli, A.; Gounelle, M. (2015). "New Paradigms For Asteroid Formation". In Michel, P.; DeMeo, F.; Bottke, W. (eds.). Asteroids IV. Space Science Series. University of Arizona Press. p. 471. arXiv: 1505.02941 . Bibcode:2015aste.book..471J. doi:10.2458/azu_uapress_9780816532131-ch025. ISBN   978-0-8165-3213-1. S2CID   118709894.
  26. Weidenschilling, S. J.; Spaute, D.; Davis, D. R.; Marzari, F.; Ohtsuki, K. (August 1997). "Accretional Evolution of a Planetesimal Swarm". Icarus . 128 (2): 429–455. Bibcode:1997Icar..128..429W. doi:10.1006/icar.1997.5747.
  27. Kary, David M.; Lissauer, Jack; Greenzweig, Yuval (November 1993). "Nebular Gas Drag and Planetary Accretion". Icarus . 106 (1): 288–307. Bibcode:1993Icar..106..288K. doi:10.1006/icar.1993.1172.
  28. Lewin, Sarah (19 August 2015). "To Build a Gas Giant Planet, Just Add Pebbles". Space.com. Retrieved 22 November 2015.
  29. Lambrechts, M.; Johansen, A. (August 2012). "Rapid growth of gas-giant cores by pebble accretion". Astronomy & Astrophysics. 544: A32. arXiv: 1205.3030 . Bibcode:2012A&A...544A..32L. doi:10.1051/0004-6361/201219127. S2CID   53961588.
  30. Helled, Ravit; Bodenheimer, Peter (July 2014). "The Formation of Uranus and Neptune: Challenges and Implications for Intermediate-mass Exoplanets". The Astrophysical Journal. 789 (1). 69. arXiv: 1404.5018 . Bibcode:2014ApJ...789...69H. doi:10.1088/0004-637X/789/1/69. S2CID   118878865.
  31. D'Angelo, Gennaro; Durisen, Richard H.; Lissauer, Jack J. (December 2010). "Giant Planet Formation". In Seager, Sara (ed.). Exoplanets. University of Arizona Press. pp. 319–346. arXiv: 1006.5486 . Bibcode:2010exop.book..319D. ISBN   978-0-8165-2945-2.
  32. Bennett, Jeffrey; Donahue, Megan; Schneider, Nicholas; Voit, Mark (2014). "Formation of the Solar System". The Cosmic Perspective (7th ed.). San Francisco: Pearson. pp. 136–169. ISBN   978-0-321-89384-0.
  33. 1 2 Johansen, Anders (April 2015). "Growth of asteroids, planetary embryos, and Kuiper belt objects by chondrule accretion". Science Advances. 1 (3): e1500109. arXiv: 1503.07347 . Bibcode:2015SciA....1E0109J. doi:10.1126/sciadv.1500109. PMC   4640629 . PMID   26601169.
  34. 1 2 3 4 5 6 Scott, Edward R. D. (2002). "Meteorite Evidence for the Accretion and Collisional Evolution of Asteroids" (PDF). In Bottke Jr., W. F.; Cellino, A.; Paolicchi, P.; Binzel, R. P. (eds.). Asteroids III. University of Arizona Press. pp. 697–709. Bibcode:2002aste.book..697S. ISBN   978-0-8165-2281-1.
  35. Shukolyukov, A.; Lugmair, G. W. (2002). "Chronology of Asteroid Accretion and Differentiation" (PDF). In Bottke Jr., W. F.; Cellino, A.; Paolicchi, P.; Binzel, R. P. (eds.). Asteroids III. University of Arizona Press. pp. 687–695. Bibcode:2002aste.book..687S. ISBN   978-0-8165-2281-1.
  36. "How comets were assembled". University of Bern, via Phys.org. 29 May 2015. Retrieved 8 January 2016.
  37. Jutzi, M.; Asphaug, E. (June 2015). "The shape and structure of cometary nuclei as a result of low-velocity accretion". Science . 348 (6241): 1355–1358. Bibcode:2015Sci...348.1355J. doi: 10.1126/science.aaa4747 . PMID   26022415. S2CID   36638785.
  38. Weidenschilling, S. J. (June 1997). "The Origin of Comets in the Solar Nebula: A Unified Model". Icarus. 127 (2): 290–306. Bibcode:1997Icar..127..290W. doi:10.1006/icar.1997.5712.
  39. Choi, Charles Q. (15 November 2014). "Comets: Facts About The 'Dirty Snowballs' of Space". Space.com. Retrieved 8 January 2016.
  40. Nuth, Joseph A.; Hill, Hugh G. M.; Kletetschka, Gunther (20 July 2000). "Determining the ages of comets from the fraction of crystalline dust". Nature . 406 (6793): 275–276. Bibcode:2000Natur.406..275N. doi:10.1038/35018516. PMID   10917522. S2CID   4430764.
  41. "How Asteroids and Comets Formed". Science Clarified. Retrieved 16 January 2016.
  42. 1 2 Levison, Harold F.; Donnes, Luke (2007). "Comet Populations and Cometary Dynamics". In McFadden, Lucy-Ann Adams; Weissman, Paul Robert; Johnson, Torrence V. (eds.). Encyclopedia of the Solar System (2nd ed.). Amsterdam: Academic Press. pp.  575–588. ISBN   978-0-12-088589-3.
  43. Greenberg, Richard (1985). "The Origin of Comets among the Accreting Outer Planets". In Carusi, Andrea; Valsecchi, Giovanni B. (eds.). Dynamics of Comets: Their Origin and Evolution. Astrophysics and Space Science Library, Volume 115. Vol. 115. Springer Netherlands. pp. 3–10. Bibcode:1985ASSL..115....3G. doi:10.1007/978-94-009-5400-7_1. ISBN   978-94-010-8884-8. S2CID   209834532.
  44. "Evaporation and Accretion of Extrasolar Comets Following White Dwarf Kicks". Cornell University Department of Astronomy. 2014. Retrieved 22 January 2016.
  45. 1 2 Filacchione, Gianrico; Capaccioni, Fabrizio; Taylor, Matt; Bauer, Markus (13 January 2016). "Exposed ice on Rosetta's comet confirmed as water" (Press release). European Space Agency. Archived from the original on 18 January 2016. Retrieved 14 January 2016.
  46. Filacchione, G.; de Sanctis, M. C.; Capaccioni, F.; Raponi, A.; Tosi, F.; et al. (13 January 2016). "Exposed water ice on the nucleus of comet 67P/Churyumov–Gerasimenko". Nature . 529 (7586): 368–372. Bibcode:2016Natur.529..368F. doi:10.1038/nature16190. PMID   26760209. S2CID   4446724.
  47. Krishna Swamy, K. S. (May 1997). Physics of Comets. World Scientific Series in Astronomy and Astrophysics, Volume 2 (2nd ed.). World Scientific. p. 364. ISBN   981-02-2632-2.
  48. Khan, Amina (31 July 2015). "After a bounce, Rosetta". Los Angeles Times . Retrieved 22 January 2016.
  49. "Rosetta's frequently asked questions". European Space Agency. 2015. Retrieved 22 January 2016.
  50. Rickman, H; Marchi, S; AHearn, M; Barbieri, C; El-Maarry, M; Güttler, C; Ip, W (2015). "Comet 67P/Churyumov-Gerasimenko: Constraints on its origin from OSIRIS observations". Astronomy & Astrophysics. 583: Article 44. arXiv: 1505.07021 . Bibcode:2015A&A...583A..44R. doi:10.1051/0004-6361/201526093. S2CID   118394879.
  51. Michel, P.; Schwartz, S.; Jutzi, M.; Marchi, S.; Zhang, Y.; Richardson, D. C. (2018). Catastrophic Disruptions As The Origin Of 67PC-G And Small Bilobate Comets. 42nd COSPAR Scientific Assembly. p. B1.1-0002-18.
  52. Marschall, R.; Morbidelli, A.; Bottke, W. F.; Vokrouhlicky, D.; Nesvorny, D.; Deienno, R. (May 2023). "Comets are Fragments: What the Kuiper Belt Size Distribution Tells Us About its Collisional Evolution". Asteroids, Comets, Meteors 2023: 2470.