meteorite

meteorite
meteoritic /mee'tee euh rit"ik/, meteoritical, meteorital /mee'tee euh ruyt"l/, adj.
/mee"tee euh ruyt'/, n.
1. a mass of stone or metal that has reached the earth from outer space; a fallen meteoroid.
2. a meteoroid.
[1815-25; METEOR + -ITE1]

* * *

Any interplanetary particle or chunk of stony or metallic matter (meteoroid) that survives passage through Earth's atmosphere and strikes the ground or that reaches the surface of another planet or moon.

On Earth the speed of entry
at least 7 mi/second (11 km/second)
generates enough friction with the air to vaporize part or all of the meteoroid and produce a streak of light (meteor). Though vast numbers of meteoroids enter Earth's atmosphere each year, only a few hundred reach the ground.

* * *

Introduction
  any fairly small natural object from interplanetary space—i.e., a meteoroid (meteor and meteoroid)—that survives its passage through Earth's atmosphere and lands on the surface. In modern usage the term is broadly applied to similar objects that land on the surface of other comparatively large bodies. For instance, meteorite fragments have been found in samples returned from the Moon, and at least one meteorite has been identified on the surface of Mars by the robotic rover Opportunity. The largest meteorite that has been identified on Earth was found in 1920 in Namibia. Named the Hoba meteorite, it measures 2.7 metres (9 feet) across, is estimated to weigh nearly 60 tons, and is made of an alloy of iron and nickel. The smallest meteorites, called micrometeorites, range in size from a few hundred micrometres (μm) to as small as about 10 μm and come from the population of tiny particles that fill interplanetary space (see interplanetary dust particle).

      Laboratory, astronomical, and theoretical studies show that most discrete meteorites found on Earth are fragments of asteroids (asteroid) that orbit in the inner portion of the main asteroid belt, between about 2.1 and 2.6 astronomical units (AU) from the Sun. (One astronomical unit is the average distance from Earth to the Sun—about 150 million km [93 million miles].) It is in this region that strong gravitational perturbations by the planets, especially Jupiter, can put meteoroids into Earth-crossing orbits. Not all meteoroids need to have formed in this region, however, as there are a number of processes that can cause their orbits to migrate over long time periods. Less than 1 percent of meteorites are thought to come from the Moon or Mars. On the other hand, there is good reason to believe that a significant fraction of the micrometeorites found drifting down through Earth's upper atmosphere come from comets. Although evidence from studies of meteors suggests that a small fraction of the cometary material that enters Earth's atmosphere in discrete chunks possesses sufficient strength to survive to reach the surface, it is not generally believed that any of this material exists in meteorite collections. For further discussion of the sources of meteorites and the processes by which they are brought to Earth, see meteor and meteoroid: Reservoirs of meteoroids in space (meteor and meteoroid) and Directing meteoroids to Earth (meteor and meteoroid).

      The principal driving force behind meteorite studies is the fact that small bodies (small body) such as asteroids and comets are most likely to preserve evidence of events that took place in the early solar system. There are at least two reasons to expect that this is the case. First, when the solar system began to form, it was composed of gas and fine-grained dust. The assembly of planet-size bodies from this dust almost certainly involved the coming together of smaller objects to make successively larger ones, beginning with dust balls and ending, in the inner solar system, with the rocky, or terrestrial, planets—Mercury, Venus, Earth, and Mars. In the outer solar system the formation of Jupiter, Saturn, and the other giant planets is thought to have involved more than simple aggregation, but their moons—and comets—probably did form by this basic mechanism. Available evidence indicates that asteroids and comets are leftovers of the intermediate stages of the aggregation mechanism. They are therefore representative of bodies that formed quite early in the history of the solar system. (See also solar system: Origin of the solar system (solar system); planetesimal.) Second, in the early solar system various processes were in operation that heated up solid bodies. The primary ones were decay of short-lived radioactive isotopes (isotope) within the bodies and collisions between the bodies as they grew. As a result, the interiors of larger bodies experienced substantial melting, with consequent physical and chemical changes to their constituents. Smaller bodies, on the other hand, generally radiated away this heat quite efficiently, which allowed their interiors to remain relatively cool. Consequently, they should preserve to some degree the dust and other material from which they formed. Indeed, certain meteorites do appear to preserve very ancient material, some of which predates the solar system.

Recovery of meteorites
 Meteorites traditionally are given the name of a geographic feature associated with the location where they are found. Until quite recently, there were no systematic efforts to recover them. This was largely because meteorites fall more or less uniformly over Earth's surface and because there was no obvious way to predict where they would fall or could be found. When a meteorite was seen to fall or when a person chanced upon an unusual-looking rock, the specimen was simply taken to a museum or a private collector.

      In the 1930s and '40s, enterprising meteorite collectors began crisscrossing the prairie regions of North America, asking farmers to bring them unusual rocks that they had found while plowing their fields. Prairie soil is largely derived from fine glacial loess and contains few large rocks. The collectors realized that there was a reasonable chance that any rocks the farmers unearthed would include meteorites.

      A better approach to finding meteorites than searching places with few rocks, however, is to search places where they can accumulate over time—i.e., where the surface is quite old and rates of weathering are low. Because meteorites contain minerals, such as iron metal, that are easily weathered, they do not normally last long on Earth's surface. Liquid water is one of the principal agents of weathering. In desert environments, where there is little water, meteorites survive much longer. Indeed, they tend to accumulate on the surface in arid regions if weathering rates are slower than the rates at which meteorites fall to Earth, provided that little windblown sand accumulates to bury them. Areas of the Sahara in North Africa and the Nullarbor Plain region in Australia have proved to be good places to look for meteorites. The most-successful collection efforts, however, have been in Antarctica.

      The Antarctic can be viewed as a cold desert. Annual snowfall is quite low over most of the interior, and the intense cold slows weathering rates considerably. Most meteorites that fall on the ice sheet become buried and are stored for 20,000–30,000 years, although some appear to have been in Antarctica for a million years or more. The ice of the Antarctic sheet gradually flows radially from the South Pole northward toward the coast. In places, the ice encounters an obstruction, such as a buried hill, that forces it to flow upward. Strong katabatic winds, which sweep down the gently sloping ice sheets from the centre of the continent, sandblast the upwelling ice with snow and ice particles, eroding it at rates as high as 5–10 cm (2–4 inches) per year and leaving the meteorites stranded on the surface. Areas of upwelling ice, called blue ice for its colour, can be recognized from aerial or satellite photographs, and on foot the dark meteorites are relatively easy to spot against the ice and snow. The drawback of collecting in Antarctica is the harsh conditions that the collection teams must endure for weeks to months while camping out on the ice. Since the 1970s several countries, notably the United States and Japan, have operated scientific collection programs. Some tens of thousands of meteorites have been brought back from Antarctica by the two countries' programs, increasing the number of meteorites available to researchers manyfold. These include half of all known Martian meteorites, 80 percent of known lunar meteorites, and numerous other rare or unique samples. Because large numbers of Antarctic meteorites are found within small areas, the traditional geographic naming system is not used for them; rather, an identifier is made up of an abbreviated name of some local landmark plus a number that identifies the year of recovery and the specific sample. (See also Antarctic meteorite.)

Types of meteorites
 Meteorites traditionally have been divided into three broad categories— stony meteorites (or stones), iron meteorites (irons), and stony iron meteorites (stony irons)—based on the proportions of rock-forming minerals and nickel-iron (also called iron-nickel) metal alloy they contain. Stony meteorites make up about 94 percent of all known meteorites, irons about 5 percent, and stony irons about 1 percent.There is considerable diversity within each category, leading to numerous subdivisions (classes, groups, etc.) based on variations in chemistry, mineralogy, and structure. It is important to realize that meteorite classification is based primarily on observable characteristics. Just because subdivisions belong to the same category, it does not necessarily follow that they all consist of meteorites that have the same or similar parent bodies. Indeed, more often than not, they are unrelated. Conversely, subdivisions from different categories may have a common origin. For instance, if a large asteroid were to melt, its denser metallic components would tend to sink to its centre (its core), while its less-dense rocky material would form a mantle around it, much like what happened to Earth. This separation process is known as geochemical differentiation. When the differentiated asteroid is later broken up by collisions, samples of its rocky mantle, iron core, and core-mantle interface might be represented in the three main categories. Thus, the challenge for researchers is to determine which types of meteorites are related and which are not, as well as to identify the processes that were responsible for the tremendous diversity that is seen among them.

Stony meteorites (stony meteorite)
Chondrites (chondrite)
 The most fundamental distinction between the various stony meteorites is between those that were once molten, the achondrites, and those that were not, the chondrites. Chondrites have been subdivided into three main classes—ordinary, carbonaceous, and enstatite chondrites—and these in turn have been divided into a number of groups.

      Chondrites are the most abundant meteorites (about 87 percent of stony meteorites) in collections. They also are arguably the most important. In terms of terrestrial rocks, these meteorites seem akin to sedimentary conglomerates—i.e., fragments of preexisting rock cemented together. They are a mechanical mixture of components that formed in the solar nebula or even earlier. Perhaps more remarkably, the compositions of chondrites are very similar to that of the Sun, except for the absence (in chondrites) of very volatile elements such as hydrogen and helium. The Sun contains more than 99 percent of the mass of the solar system. The composition of the Sun must therefore be very close to the average composition of the solar system when it formed. As a result, the Sun's composition can serve as a reference. Deviations in a meteorite's composition from this reference composition provide clues to the processes that influenced the formation of its parent body and the components in it.

Chondrules (chondrule)
 Meteorites are classified as chondrites based on the presence within them of small spherical bodies (typically about 1 mm [0.04 inch] in diameter) called chondrules (chondrule). From their shapes and the texture of the crystals in them, chondrules appear to have been free-floating molten droplets in the solar nebula. Simulation experiments show that chondrules formed by “flash” heating (to peak temperatures of 1,400–1,800 °C) and then rapid cooling (10–1,000 °C per hour). The sizes, compositions, and proportions of different types of chondrules vary from one chondrite meteorite to the next, which means that chondrule formation must have been a fairly localized process. There is also good evidence for its occurring many times. If chondrule abundance in chondrites is any guide, the chondrule-forming process was one of the most energetic and important in the solar nebula, at least in the region of the asteroid belt. Nevertheless, despite more than a century of study and speculation, scientists have yet to determine definitively what the process was.

Refractory inclusions
      Minor but important constituents of chondrites are refractory inclusions. They are so termed because they are highly enriched in the least-volatile, or refractory, elements. Because calcium and aluminum are two of the most abundant refractory elements in them, they are often called calcium-aluminum-rich inclusions, or CAIs. They range in shape from highly irregular to spherical and in size from tens of micrometres up to a centimetre or more. Like chondrules, they formed at high temperatures but appear to have been heated for more prolonged periods. Many but not all types of inclusions appear to have been formed from a molten state, which probably came about by the heating of preexisting solids. Others seemed to have formed as crystalline solids that condensed directly from a hot gas. Like chondrules, there is no consensus on the mechanism or mechanisms that formed refractory inclusions.

      The space between the chondrules and refractory inclusions is filled with a fine-grained matrix that cements the larger meteoritic components together. The matrix is richer in volatile elements than are chondrules and inclusions, suggesting that at least some fraction of it formed at a lower temperature. The matrix of many chondrites contains organic matter (up to about 2 percent by weight). The isotopic compositions of the hydrogen and nitrogen atoms in the organic matter are often very unusual. These compositions are best explained if at least some of the organic matter was produced in the interstellar molecular cloud from which the solar system formed. Other materials that predate the solar system survive in the matrix, albeit at much lower concentrations. Unlike the organic matter, these materials formed not in the interstellar medium but around stars that died millions to hundreds of millions of years before the solar system formed. The evidence that these tiny grains (a few nanometres to 10 micrometres in size) have circumstellar origins lies in their isotopic compositions. These are so different from the compositions of solar system materials that they could only have been produced by nucleosynthesis (formation of elements) in stars. For instance, the average ratio of carbon-12 to carbon-13 observed in solar system objects is about 89 to 1, with a range of about 85–94 to 1. For some material isolated from chondrites, the carbon-12/carbon-13 ratios of individual particles range from about 2 to 1 to about 7,000 to 1. Types of minerals of circumstellar origin that have been isolated from chondrites include diamond, graphite, silicon carbide, silicon nitride, olivine, corundum, spinel, chromite, and hibonite.

Alteration processes
      Few if any chondrites have remained completely unaltered since they formed as part of their larger parent asteroids. Three processes have modified the chondrites to varying degrees: aqueous alteration, thermal metamorphism, and shock. Soon after the chondritic parent bodies were formed, they were all heated to some degree. In some bodies, temperatures were modest but high enough for liquid water to exist; reaction of the original minerals with water—aqueous alteration—transformed them to complex mixtures of minerals. Other chondritic parents were heated more intensely, and, if they once contained water, it was driven off. The temperatures achieved were high enough to induce changes in mineralogy and physical structure—thermal metamorphism—but insufficient to cause widespread melting. At an early stage, this heating resulted in an increasing uniformity of mineral composition and recrystallization of the matrix. Organic matter and circumstellar grains in the matrix were also destroyed at this stage. With more intense heating, even the chondrules recrystallized. In the most-metamorphosed chondrites examined, those whose parent bodies experienced temperatures of roughly 1,000 °C (about 1,800 °F or 1,270 K), the chondrules are quite difficult to see. The third modification process, shock, is caused by collisions of meteoritic parent bodies. Not just chondrites but all major types of meteorites exhibit shock features, which range from minor fracturing to localized melting. The processes of aqueous alteration and thermal metamorphism were probably finished within about 50 million years of the formation of the solar system. On the other hand, collisions of asteroids and their fragments continue to this day.

Classification systems
      As can be gathered from the preceding discussion, the features now seen in chondrites reflect processes from two distinct episodes—those that led to the formation of the chondritic parent bodies and those that later altered the material in the parent bodies. As a result, chondrites are classified in two complementary ways. Based on the concentrations of their major elements (iron, magnesium, silicon, calcium, and aluminum) and on their oxidation states, oxygen isotopic compositions, and petrology (e.g., abundance of chondrules and matrix, chondrule size, and mineralogy), chondrites naturally cluster into distinct classes and groups. It is generally believed that the defining characteristics of the classes and groups were determined by conditions prior to and during the formation of the meteorites' parent bodies and that each group comes from a different parent asteroid or set of asteroids.

 In addition, within each of the groups, the meteorites differ in the degree to which they were thermally metamorphosed or aqueously altered. These differences are referred to as petrologic types; they are broken down in the petrologic types table—>. . Types 2 and 1 represent increasing degrees of alteration by water, and types 3 through 6 (some researchers extend the types to 7) reflect increasing degrees of modification by heating. Thus, a meteorite that experienced extensive aqueous alteration would be classified type 1, and one that experienced temperatures just short of melting would be type 6 (or 7). A meteorite that remained completely unmodified by either process since its formation would lie at the boundary of types 2 and 3.

 As an example of how the two classification methods are applied, the carbonaceous chondrite known as the Allende meteorite, whose fall was witnessed in 1969, is classified CV3. This indicates that it belongs to the CV group and petrologic type 3 of the second table—>.

      Meteorites are also classified according to the severity of shock and the terrestrial weathering they have experienced, but these schemes are less commonly used. Still another way to distinguish meteorites is as “falls” or “finds,” depending on whether or not they were observed to fall to Earth.

CI carbonaceous chondrites
      Perhaps the most interesting type of chondrite is the CI group of carbonaceous chondrites (carbonaceous chondrite). Strictly speaking, it could be questioned why such meteorites are called chondrites at all, inasmuch as they do not contain chondrules. They are aqueously altered so heavily that, if they once contained chondrules, all evidence of them has been erased. When their elemental abundances are compared with those of the Sun, however, it turns out that the two are extremely similar. In fact, of all meteorite types, the CI chondrites most closely resemble the Sun in composition. Consequently, in devising a classification scheme, it makes sense to group them with the chondrites.

      Because CI chondrites are chemically so Sun-like—and thus so like the average composition of the forming solar system—some scientists have speculated that they are of cometary rather than of asteroidal origin. Comets are believed to represent the most unaltered material in the solar system. Although there are difficulties with this idea, scientific knowledge about the nature and origin of comets is still limited, which makes it unwise to entirely dismiss this intriguing possibility.

Achondrites (achondrite)
  achondrites, their name meaning “without chondrites,” are a relatively small but diverse group of meteorites. They exhibit a range of features that would be expected if their parent bodies experienced widespread melting: igneous features similar to those observed in terrestrial volcanic rocks, segregation of molten metal (possibly into a core) from molten silicate rock (magma), and magmatic segregation of silicate crystals and melt. Most achondrites collected on Earth are derived from asteroids, but one small group is thought to come from Mars and another from the Moon.

      The three most numerous asteroidal achondrite groups are the aubrites, the howardite-eucrite-diogenite association, and the ureilites. Aubrites are also known as enstatite achondrites. Like the enstatite class of chondrites, the aubrites derive from parent bodies that formed under highly chemically reducing conditions. As a result, they contain elements in the form of less-common compounds—for example, calcium as the sulfide mineral oldhamite (CaS) rather than in its more usual silicate and carbonate forms.

      The howardite, eucrite, and diogenite (HED) meteorites all seem to be related to one another and probably came from the same asteroidal body, tentatively identified as Vesta, the second largest member of the asteroid belt. They have also been linked to the mesosiderites, a group of stony iron meteorites (meteorite) (see below Association of meteorites with asteroids (meteorite)). The HED parent body seems to have had a complex history that included melting, segregation of metal into a core, crystallization, metamorphism, and impact brecciation (the process in which an impact shatters rock).

      The eucrites are subdivided into cumulate eucrites and basaltic eucrites. Cumulate eucrites are like terrestrial gabbros in that they seem to have formed at depth in their parent body and crystallized quite slowly. By contrast, basaltic eucrites are similar to terrestrial basalts, apparently having formed at or near the surface of their parent body and cooled relatively fast. The diogenites, composed predominantly of the mineral pyroxene, also seem to have formed at depth. The howardites are impact breccias composed of cemented fragments of diogenite and eucrite materials.

      The third main class of asteroid-derived achondrites, the ureilites, are carbon-bearing. They consist of a silicate rock, made primarily of the minerals olivine and pyroxene, that has dark veins running through it. The veins, which constitute as much as 10 percent of the meteorites, are composed of carbon (graphite and some diamond), nickel-iron metal, and sulfides. The silicates clearly crystallized from magma, but there is debate about how they formed. The carbon-rich veins seem to have formed by shock-induced redistribution of graphite that originally crystallized along with the silicates. In addition to the three main achondrite classes, there exist several minor classes and a collection of unique achondrite specimens, all of which reflect the variability of melting processes in the asteroids.

 About three dozen meteorites have been identified as having come from Mars. All are volcanic rocks. All but one of these belong to one of three classes—shergottites, nakhlites, and chassignites—which were established well before a Martian origin was suspected. The three groups are often referred to collectively as SNCs. One piece of evidence for a planetary origin of the SNCs is their young age, between 150 million and 1.3 billion years. To retain enough heat so that volcanic activity could continue until just 1.3 billion years ago, let alone more recently, required a planet-sized parent body. Because there is considerable geochemical evidence that the rocks did not originate on Earth, the only likely candidates that remain are Venus and Mars, both of which appear to have experienced recent volcanic activity. The most convincing evidence for a Martian origin comes from an Antarctic meteorite, an SNC named EETA79001. This meteorite contains trapped gases (noble gases (noble gas), nitrogen, and carbon dioxide) whose relative abundances and isotopic compositions are almost identical to those of the Martian atmosphere as measured by the two Viking landers. Scientists believe that the Martian meteorites are fragments of the planet's near surface that were launched into space by large impacts and that eventually found their way to Earth. In the case of EETA79001, atmospheric gases apparently became trapped in glasses produced during the violent shock event that excavated the rock from Mars. As the only samples of Mars available to scientists on Earth, Martian meteorites provide a unique window into the evolution of this enigmatic planet.

      Several Martian meteorites have been aqueously altered to some degree, which is in line with other evidence that liquid water was present at least periodically on Mars at some time in the past. The most unique Martian meteorite is another Antarctic specimen, ALH84001. This rock, an orthopyroxenite, has a crystallization age of about 4.5 billion years, which is roughly the same age as asteroidal meteorites (see below The ages of meteorites and their components (meteorite)), but several of its properties clearly tie it to the other Martian meteorites. About 3.9 billion years ago, aqueous fluids passed through it, precipitating carbonate-magnetite-sulfide mineral assemblages. Some researchers interpreted these rather unusual assemblages as evidence for life on Mars. They also reported features in the meteorite that they interpreted as fossilized bacteria. These claims created considerable controversy, but they also generated important debate on how life might originate and how it might be recognized even if it is unlike the life known on Earth.

      A number of lunar meteorites have been found in Antarctica and hot deserts on Earth. They probably would not have been recognized as having come from the Moon were it not for the lunar samples brought back by the manned Apollo (Apollo program) and robotic Luna missions. The meteorites, which likely are fragments blasted off the Moon by large impacts, resemble the various rock types represented in the lunar samples (e.g., mare basalts, highland regolith breccias, and highland impact-melt breccias), but they almost certainly came from areas that were not sampled by the various missions. Therefore, like the Martian meteorites, they are an important source of new information on the formation and evolution of their parent body.

Iron meteorites
  iron meteorites are pieces of denser metal that segregated from the less-dense silicates when their parent bodies were at least partially melted. They most probably came from the cores of their parent asteroids, although some researchers have suggested that metal, rather than forming a single repository, may have pooled more locally, producing a structure resembling raisin bread, with metal chunks as the “raisins.” The latter would have been likely to occur if the asteroid underwent localized shock melting rather than melting of the entire body.

      Iron meteorites are principally composed of two nickel-iron minerals, nickel-poor kamacite and nickel-rich taenite. The abundances of these two minerals strongly influence the structure of iron meteorites. At one extreme is the class known as hexahedrites, which are composed almost entirely of kamacite. Being nearly of a single mineral, hexahedrites are essentially structureless except for shock features. At the other extreme is the class known as ataxites, which are made up primarily of taenite. Ataxites are the rarest class and can contain up to about 60 percent nickel by weight. Again, because they are nearly monomineralic, they are almost featureless structurally. Between these two classes are the octahedrites. In these meteorites, kamacite crystals form as interlocking plates in an octahedral arrangement, with taenite filling the interstices. This interlocking arrangement, called the Widmanstätten pattern, is revealed when a cut and polished surface of the meteorite is etched with dilute acid. The pattern is an indication that octahedrites formed at relatively low pressure, as would be expected if they formed in asteroid-sized bodies.

      At one time iron meteorites were classified in terms of nickel content and Widmanstätten structure, but this has been largely superseded by a chemical classification based on gallium, germanium, and nickel content. The most-common classes have the rather uninspiring names IAB, IIAB, IIIAB, IVA, and IVB. There are numerous other smaller classes and unique iron meteorites. On the assumption that most iron meteorites formed in the cores of their parent asteroids, variations in the composition and properties of iron meteorites in a given class reflect the changing conditions during solidification of these cores. Gallium and germanium abundances in molten nickel-iron metal are relatively unaffected by the process of crystallization, but they are sensitive to the conditions under which the parent asteroid formed. Thus, iron meteorites with similar gallium and germanium abundances are probably related to one another, either because they came from the same asteroid or because their parent asteroids formed at similar times and places. Nickel abundances, on the other hand, are influenced by crystallization because nickel tends to concentrate in those portions of nickel-iron metal that are still molten. As a result, nickel abundances can be used to determine the sequence of crystallization within iron meteorite classes.

      The IAB, IIICD, and IIE iron meteorites exhibit geochemical characteristics that are distinct from those of the other classes of irons. Their origin remains uncertain, but they may have been produced by impact processes.

Stony iron meteorites (stony iron meteorite)
       stony iron meteorites contain roughly equal amounts of silicate minerals and nickel-iron metal. They fall into two groups: pallasites and mesosiderites. Pallasites are composed of a network of nickel-iron metal in which are set crystals of the silicate mineral olivine. Olivine crystals are typically about 0.5 cm across. The centres of large areas of metal exhibit the Widmanstätten structure. Pallasites formed at the interface between regions of molten nickel-iron metal and molten silicates. The molten nickel-iron metal regions could have been the outer cores of asteroids or, less likely, nuggets in the asteroids where the metal had collected. Similarly, the molten silicate regions could have been the deepest layers of the silicate mantle.

      As discussed above in the section Achondrites (meteorite), mesosiderites are probably related to the three classes of achondrites collectively called HEDs. Like one of the HED classes, howardites, mesosiderites are impact breccias containing fragments belonging to the other two classes, eucrites and diogenites. In addition, however, the mesosiderites contain a large amount of dispersed nickel-iron metal. The origin of the metal is not known for certain, but it may be from the core of the body that collided with and brecciated the mesosiderite parent body.

Association of meteorites with asteroids
 If meteoritic material comes from specific regions of the asteroid belt, then the asteroids in such regions should have the chemical and mineralogical composition observed in the meteorites. The surface mineralogical composition of asteroids, in principle, can be determined directly by observations from Earth of the fraction of sunlight they reflect ( albedo) and the spectrum of the reflected light (reflectance spectrum). A number of processes conspire, however, to make the association of certain asteroids with the various meteorite groups much more difficult than might be expected.

      Although no two asteroidal reflectance spectra are exactly alike in detail, most asteroids fall into one of two general groups, the S class and the C class. S class asteroids (e.g., Gaspra and Ida, observed by the Galileo spacecraft, and Eros, visited by the NEAR Shoemaker spacecraft) have moderate albedos and contain mixtures of olivine, pyroxene, and metallic iron. These are the same minerals found in ordinary chondrites, but they also are present in a number of other meteorite types. The C class asteroids (e.g., Mathilde, observed by NEAR Shoemaker) have low albedos, and their more featureless spectra indicate the presence of light-absorbing materials, although at least half have a spectral feature associated with iron-bearing hydrous silicates. It is plausible to consider the C class asteroids as candidate sources for certain groups of carbonaceous chondrite meteorites. Their low albedos and spectral evidence of hydrous silicates, however, make them unlikely sources of ordinary chondrites.

      When the S class asteroids are considered in more detail, there are difficulties in identifying them all as sources of ordinary chondrites. Largely because of their apparent range of mineralogies—specifically their ratios of olivine to pyroxene—the S class asteroids have been divided into seven subclasses. In light of this, it is possible that the S class actually represents a number of unrelated groups of asteroids. In addition, some research has linked the S class asteroids to several groups of achondrites. On the other hand, if most S type asteroids are not related to the ordinary chondrites, scientists would be challenged to explain how an uncommon and unidentified class of asteroid is supplying most of the meteorites to Earth.

      The asteroids in the S(IV) subclass seem to have mineralogies that best match those of the ordinary chondrites. This is supported by measurements made by NEAR Shoemaker of the elemental composition of the surface of Eros, which is classified as an S(IV) asteroid. With the notable exception of a low sulfur content, the composition of Eros was found to be consistent with that of an ordinary chondrite.

      Scientists have come to recognize relatively recently that the surfaces of asteroids and other solid bodies are not necessarily representative of what lies just a short distance beneath those surfaces. Both Eros's low-sulfur measurement and the fact that, overall, the spectra of S class asteroids do not exactly match those of ordinary chondrites may be due, at least partially, to the effects of a poorly understood set of processes collectively called space weathering. Important component processes of space weathering are thought to be the impacts of meteorites and micrometeorites and the impingement of energetic solar wind particles and solar radiation on surface materials. Over time these processes act to modify the chemical and physical surface properties of airless bodies such as Mercury, the Moon and some other planetary satellites, and asteroids and comets. Comparisons of younger surfaces around craters with older terrains on Eros by NEAR Shoemaker, and on Gaspra and Ida by Galileo, support the idea that space weathering occurs on asteroids.

      Asteroids are thought to be covered by a layer of pulverized rock, called regolith, produced by bombardment with meteorites of all sizes over millions to billions of years. The regolith need only be as thin as a few sheets of paper to completely mask the underlying material from reflectance spectroscopy, although on most asteroids it is probably much thicker. Unfortunately, because it is so loosely bound together, this regolith material does not survive entry into Earth's atmosphere in pieces that are large enough to identify as meteorites and analyze. Consequently, scientists do not have samples of regolith that can be compared with meteorites or asteroids directly. On the Moon, however, systematic changes are observed in the mineralogy and reflectance properties of the surface material as a result of this collisional grinding and other space weathering processes. Thus, although it seems likely that ordinary chondrites do come from S class asteroids, space weathering may be making it difficult to determine with certainty which S class asteroids are the parent bodies of these meteorites and which are unrelated but have a grossly similar mineralogy.

      Space weathering must also affect the spectra of the asteroidal sources of the other meteorite groups. Nevertheless, a number of more-or-less convincing associations between groups of meteorites and types of asteroids have been made. It has been proposed that the CV and CO groups of carbonaceous chondrites come from the K class asteroids. As mentioned above, a number of lines of evidence, including spectral measurements, point to the asteroid Vesta being the source of the howardite-eucrite-diogenite association and the mesosiderites. The most likely source of the iron meteorites is the M class of asteroids, but enstatite chondrites and mesosiderites have also been linked to them. The pallasites may come from A class asteroids. For additional discussion of asteroid classes and their compositions, see asteroid: Composition (asteroid).

The ages of meteorites and their components
      When the planets and asteroids formed, they contained a number of different radioactive isotopes, or radionuclides. Radionuclides decay at characteristic rates. The time it takes for half of the atoms of a quantity of of a radionuclide to decay, the half-life, is a common way of representing its decay rate. Many radionuclides have half-lives that are similar to or longer than the age of the solar system; for this reason they are often called long-lived radionuclides. As a result of their longevity, they are still present in meteorites and on Earth, and they are commonly used for dating rocks and meteorites.

      Scientists typically determine the age of a rock or meteorite using the isochron method. For purposes of illustration, consider the rubidium-strontium decay system. In this system, the radioactive parent rubidium-87 (87Rb) decays to the stable daughter isotope strontium-87 (87Sr). The half-life for 87Rb decay is 48.8 billion years. Strontium has a number of other stable isotopes, including strontium-86 (86Sr), which is often used as a reference. When a rock forms, the minerals within it have identical strontium isotopic compositions (e.g., 87Sr/86Sr ratios) but often have different rubidium/strontium ratios (e.g., 87Rb/86Sr ratios). In this case, as 87Rb decays, the 87Sr/86Sr ratios in the minerals all increase with time but at different rates—the 87Sr/86Sr ratios increase more rapidly in minerals with higher initial 87Rb/86Sr ratios. If the minerals' 87Sr/86Sr ratios as they exist now are plotted on a graph against their 87Rb/86Sr ratios, the data points form a straight line, called an isochron. The slope of the line is proportional to the time since the minerals formed, and the point where the line intercepts the 87Sr/86Sr axis (i.e., when 87Rb/86Sr is zero) gives the initial ratio when the minerals formed.

      In this illustration, the minerals within a single rock are used to date it, and the line on the graph is called an internal isochron. The same principle can be applied if one uses numerous rocks that formed at the same time and place but which had different initial 87Rb/86Sr ratios. The result is called a whole-rock isochron. In practice, an isochron is ambiguous in that it dates the time either when the minerals or rocks formed or when they were last heated and the strontium isotopes in them rehomogenized. Consequently, other evidence about a rock or suite of rocks is needed to determine what the isochron is actually dating. If the data points for minerals or rocks do not fall on a line, it indicates that the system has been disturbed and cannot be used for dating. Shock is the most common cause of disturbed systems in meteorites.

      In addition to the long-lived radionuclides, a number of short-lived radionuclides were present in the early solar system. Most of these have half-lives of only a few million years or less. They will have decayed away long ago and cannot be used to obtain absolute ages directly. However, their original abundances in some objects can still be determined by the isochron method. By comparing the original abundances of a short-lived radionuclide in different objects, scientists can determine their relative ages. If one or more of these objects also have had their absolute ages determined using long-lived radionuclides, the relative ages can be converted into absolute ones. Trying to establish absolute ages for relative ages that have been determined from various short-lived radionuclides has been the focus of much modern research, but it has proved to be difficult. This is because the short-lived radionuclides typically behave chemically quite differently from one another and from the long-lived isotopes. Nevertheless, given the antiquity of meteorites, scientists have developed a remarkably accurate picture of the timing of events in the early solar system.

      The oldest objects in meteorites, with ages of approximately 4,567,000,000 years, are refractory inclusions. With a few exceptions, these are also the objects with the highest abundances of short-lived radionuclides. The absolute ages of chondrules have not been accurately measured. The abundances of the short-lived radionuclide aluminum-26 in chondrules from ordinary and carbonaceous chondrites have been interpreted to indicate that they formed over an extended period from 1 million to at least 3 and perhaps as long as 10 million years after the refractory inclusions. There is some debate, however, over whether these ages, particularly the later ones, really date when chondrules formed or, rather, when their isotopes were reset by later processes. Metamorphism in the ordinary chondrites ended between 5 and 55 million years after refractory inclusions formed, and in enstatite chondrites between 9 and 34 million years after. This age span probably reflects both the size of the chondrite parent bodies and how deeply within their parent bodies the meteoritic materials were located. Larger bodies cool more slowly, as do more deeply buried regions of a body.

      The formation ages of ordinary and enstatite chondrites are uncertain, but, given the age ranges established for the end of metamorphism, they can be no more than 5 and 9 million years after the formation of refractory inclusions, respectively. There is some evidence that enstatite chondrites formed about 2 million years after refractory inclusions. The formation ages of carbonaceous chondrites are also not known, but dating of minerals produced during their alteration by liquid water indicates they must have formed within 3–7 million years, and possibly less than 1 million years, after the formation of refractory inclusions. The crystallization ages of achondrites from their magmas range from about 4,558,000,000 to roughly 4,399,000,000 years. There is some indication that the parent body of the HED meteorites started melting about 4,565,000,000 years ago. Iron and stony iron meteorites crystallized within 10–20 million years of refractory inclusions, while relatively recent evidence suggests that metal-silicate differentiation of their parent asteroids occurred less than 1.5 million years after the formation of refractory inclusions. This again demonstrates the rapidity with which many asteroids melted, differentiated, and solidified.

Cosmic-ray (cosmic ray) exposure ages of meteorites
      The time it takes for a meteoroid to reach Earth from the asteroid belt is an important constraint when trying to identify the mechanism or mechanisms responsible for delivering meteoroids to Earth. The time cannot be measured directly, but an indication of it can be found from cosmic-ray exposure ages of meteorites. This age measures how long a meteorite existed as a small meteoroid (less than a few metres across) in space or near the surface (within a few metres) within a larger body.

      High-energy galactic cosmic rays (cosmic ray)—primarily protons (proton)—have a range of penetration on the order of a few metres in meteoroidal material. Any meteoroid of smaller dimensions will be irradiated throughout by this proton bombardment. The high-energy protons knock protons and neutrons out of the atomic nuclei of various elements present in the meteoroid (see spallation). As a consequence, a large number of otherwise rare isotopic (isotope) species, both stable and radioactive, are produced. They include the stable noble gas isotopes helium-3, neon-21, argon-38, and krypton-83 and various short- and moderately long-lived radioactive isotopes, including beryllium-10 (half-life 1.6 × 106 years), aluminium-26 (7.3 × 105 years), chlorine-36 (3 × 105 years), calcium-41 (105 years), manganese-53 (3.7 × 106 years), and krypton-81 (2.1 × 105 years). The concentration of the radioactive isotopes can be used to monitor the cosmic-ray bombardment rate, and the accumulation of the stable species (e.g., neon-21) measures the total time since this bombardment began—i.e., the time since the meteoroid was excavated by collisions from an object that was large enough to shield it from cosmic rays.

      The vast majority of meteorites have exposure ages that are greater than one million years. For chondritic meteorites, the number of meteorites with a given cosmic-ray exposure age drops off quite quickly as the age increases. Most ordinary chondrites have exposure ages of less than 50 million years, and most carbonaceous chondrites less than 20 million years. Achondrites have ages that cluster between 20 and 30 million years. Iron meteorites have a much broader range of exposure ages, which extend up to about two billion years. There are often peaks in the exposure age distributions of meteorite groups; these probably reflect major impact events that disrupted larger bodies.

      The ranges of exposure ages relate both to the dynamic evolution of meteoroid orbits and to the collisional lifetime of the meteoroids. The almost total absence of meteorites with exposure ages of less than a million years suggests that meteoroid orbits cannot become Earth-crossing in much less than a million years. Numerical simulations on computers are consistent with this, but they also predict that orbital lifetimes should fall off much faster than do the cosmic-ray exposure ages. This has prompted the suggestion that meteorites spend a significant fraction of their time as small meteoroids migrating within the asteroid belt until their orbits intersect a resonance—i.e., a region in the belt where they experience strong gravitational perturbations by the planets, particularly Jupiter—that puts the meteoroids in Earth-crossing orbits. The general drop-off in the frequency of meteorites with older exposure ages and the upper limit for most stony meteorites of 50 million years are consistent with estimates that half of any given meteoroid population is eliminated by collisions in 5–10 million years. The longer exposure ages of iron meteorites suggest that their greater strength allows them to survive longer in space. (For a detailed discussion of the resonance mechanisms that eject meteoroids from the asteroid belt, see meteor and meteoroid: Directing meteoroids to Earth. (meteor and meteoroid)

Meteorites and the formation of the early solar (solar nebula) system
      As mentioned above, scientists study meteorites for insights into the events that took place surrounding the birth and early evolution of the solar system. They know from astronomical observations that all stars form by gravitational collapse of dense regions in interstellar molecular clouds. This is almost certainly how the solar nebula formed, and the presence of preserved circumstellar and interstellar material in meteorites is consistent with this idea. Less clear is what precipitated the gravitational collapse of the region of the molecular cloud that became the solar system.

      Gravitational collapse can occur spontaneously—i.e., through random fluctuations of density. Another possibility, however, is suggested by the finding in meteorites (particularly in their refractory inclusions) of short-lived radionuclides that were present at formation (as opposed to the later production of radionuclides by recent cosmic-ray irradiation). The shortest-lived of the radionuclides found to date, calcium-41, has a half-life of only about 100,000 years. This radionuclide must have been made and incorporated into refractory inclusions within just a few half-lives (less than a million years), or its abundance would have been too low to detect. This is remarkably short by astronomical standards. Because the other short-lived radionuclides have longer half-lives, they do not put such stringent time constraints on the interval between their synthesis and formation of refractory inclusions. Nevertheless, the absolute and relative abundances of the short-lived radionuclides can be compared with the values predicted for likely sources of the radionuclides.

      One potential source of radionuclides is nucleosynthesis in stars ending their lives in catastrophic explosions called supernovas and in a class of dying stars known as asymptotic giant branch (AGB) stars. Both supernovas and AGB stars produce massive, fast-moving winds (flows of matter) that are rich in short-lived radionuclides. Numerical simulations show that under some conditions, when these winds hit an interstellar molecular cloud that cannot collapse spontaneously, they compress it to the point that it becomes gravitationally unstable and collapses. The simulations also show that some of the wind material with its complement of short-lived radionuclides is mixed into the collapsing cloud. Thus, in this scenario, the radionuclides are fingerprints of the stellar wind responsible for triggering the collapse of the molecular cloud that evolved into the Sun and planets.

      An alternative explanation for the short-lived radionuclides in meteorites has not been ruled out—their synthesis in the solar nebula by intense radiation from an early active Sun. This concept has proved somewhat less successful than the stellar-wind idea at explaining the absolute and relative abundances of the short-lived radionuclides. Nevertheless, no model incorporating either of these explanations has been completely successful in this regard.

      After collapse was initiated, the first solids known to have formed in the solar nebula were the refractory inclusions, which apparently were made in relatively short-lived heating events about 4,567,000,000 years ago. The gradation of planetary compositions from dry, rocky, metal-rich Mercury to gas-rich Jupiter and its icy moons suggests that there was a temperature gradient in the inner solar system (see also solar system: Differentiation into inner and outer planets (solar system)). Astrophysical models predict such gradients, although the absolute value of the gradient varies with the conditions assumed for a given model. One of many ideas for producing the refractory inclusions is that they formed in convection currents circulating at the edge of the hottest region of the inner nebula.

      The ambient environment in the asteroid belt at the time the asteroids were being assembled must have been thermally a rather tranquil one. The fact that presolar material is preserved in meteorites argues against widespread heating of the asteroidal region, as do the presence of water-bearing minerals and the relatively high content of volatile elements in many chondrites. Again, this is consistent with most current astrophysical models. Despite the evidence for an overall low temperature in this region of the solar system, the abundance of chondrules (chondrule) in all chondritic meteorites except the CI chondrites attest to local, transient episodes of very high temperatures.

      If chondrules were relatively rare in meteorites, their formation could be regarded as of secondary importance in the early solar system. Chondrules and their fragments, however, make up most of the mass of the most abundant class of meteorites, the ordinary chondrites, and a major portion of other chondrites, which indicates that their formation must have been of central importance. Even if the parent bodies of ordinary chondrites formed only within a restricted region of the asteroid belt adjacent to the major resonance that is thought to put chondritic material into Earth-crossing orbits (see meteor and meteoroid: Directing meteoroids to Earth (meteor and meteoroid)), this region still represents about 10 percent of the asteroid belt. It also is likely that the parent asteroids of other chondrule-bearing meteorites formed outside this region, even though they may be in that region today.

      Ideas abound for how chondrules formed—e.g., electrical discharges, shock waves, collisions between molten asteroids, and outflows associated with the early active Sun—but none has gained general acceptance. The ages of chondrules are crucial to distinguishing between some of these ideas. If they really formed over a period of 1–10 million years after refractory inclusions, this would be problematic for certain models. Fewer problems would arise if it turned out that the measured ages of most chondrules reflect when they were reheated or altered in their parent body.

      Asteroidal bodies began to form perhaps as early as one million years after refractory inclusions. Certainly, within 5–10 million years they were being heated, aqueously altered, and melted. Volcanic activity on some asteroidal bodies, presumably the larger ones, continued for as long as about 170 million years. The process responsible for this heating remains to be clearly identified. The short-lived radioactive isotopes aluminum-26 and iron-60 appear to be the most likely heat sources, but heat from electric currents induced by early solar activity and from the release of gravitational potential energy as asteroids formed also may have contributed.

      Asteroid-sized bodies presumably were forming not just in the asteroid belt but everywhere in the solar system. Concurrently, they would have begun aggregating into larger bodies in a process that eventually produced the rocky inner planets. This process was remarkably rapid. The Moon probably formed by an impact of a Mars-sized body with the growing Earth (see Moon: Origin and evolution (Moon)). The oldest Moon rocks that have been dated are about 4.44 billion years old, but there is evidence that the Moon actually formed within 30 million years of the refractory inclusions. Similarly, the oldest meteoritic material from Mars is about 4.5 billion years old, but there is evidence that Mars itself formed about 13 million years after the refractory inclusions. Thus, within as little as 30 million years of the appearance of the first solids, the aggregation process that started with tiny particles had produced the rocky inner planets.

      In order for asteroids to have formed and developed at all on the timescale of a few million years, theoretical calculations suggest that the density of matter required was more like that in the regions occupied by the giant planets. The quantity of material observed in the asteroid belt today, however, is quite small, perhaps as little as 1/10,000 of that originally present. Some natural process must have removed almost all the material in this region of the solar system after the formation of the asteroidal bodies.

      Although the details are not yet fully understood, it seems most likely that the formation of the giant planets, particularly Jupiter, quickly resulted in the evacuation of most of the matter from this region of the solar system (see asteroid: Origin and evolution of the asteroids (asteroid)). The mineralogical and chemical record in meteorites is not compatible with their ever having been part of a planet even as large as the Moon. This implies that Jupiter formed rapidly, before bodies in the asteroid belt had grown to become full-fledged planets. In the period before Jupiter approached its present mass, the asteroids would have moved in nearly circular orbits. During the final formation of Jupiter (and Saturn), the changing mass distribution in the outer solar system caused waves of resonant gravitational perturbations to sweep through the asteroid belt, increasing the eccentricities and inclinations of the asteroids to the moderate values observed today. Given what is known of the ages of asteroids and given the upper limit to the size of the present-day asteroids, this means that proto-Jupiter within about one million years of the formation of the solar system had already begun to capture the massive quantities of hydrogen and helium from the solar nebula that constitute most of the giant planet today. Such rapid growth of Jupiter apparently required a more complex formation mechanism than that for the rocky planets (see Jupiter: Origin of the Jovian system (Jupiter)).

      The foregoing scenario of early solar system evolution is likely to be wrong in some, and perhaps many, of the details. Nevertheless, without the samples of asteroids and primitive solar system materials provided by meteorites, there would be little observational basis at all for formulating models of this kind. For good reason, meteorites have been dubbed “poor man's space probes.” Until spacecraft missions bring back a variety of samples from asteroids and comets, the most precise and detailed data for the evolution of the solar system will come from meteorites.

Conel M.O'D. Alexander George W. Wetherill

Additional Reading
Introductory information can be found in Brigitte Zanda and Monica Rotaru (eds.), Meteorites: Their Impact on Science and History (2001; originally published in French, 1996); Alex Bevan and John de Laeter, Meteorites: A Journey through Space and Time (2002); Harry Y. McSween, Jr., Meteorites and Their Parent Planets, 2nd ed. (1999); O. Richard Norton, Rocks from Space: Meteorites and Meteorite Hunters, 2nd ed. (1998); and Fritz Heide and Frank Wlotzka, Meteorites: Messengers from Space (1995). More-advanced treatments are John T. Wasson, Meteorites: Their Record of Early Solar-System History (1985); V.A. Bronshten, Physics of Meteoric Phenomena (1983; originally published in Russian, 1981); and Robert Hutchison, Meteorites: A Petrologic, Chemical, and Isotopic Synthesis (2004). Dante S. Lauretta and Marvin Killgore, A Color Atlas of Meteorites in Thin Section (2005), provides a comprehensive collection of often-striking photographs of petrologic thin sections of meteorites. A descriptive and historical treatment of iron meteorites, including beautiful photographs, is Vagn F. Buchwald, Handbook of Iron Meteorites, Their Distribution, Composition, and Structure, 3 vol. (1975). H.H. Nininger, Out of the Sky: An Introduction to Meteorites (1952, reprinted 1959), provides firsthand experiences of fall phenomena on a nontechnical level. A catalog of known meteorites to the end of 1999, including data regarding their fall, is M.M. Grady (ed.), Catalogue of Meteorites, 5th ed. (2000).Conel M.O'D. Alexander George W. Wetherill

* * *


Universalium. 2010.

Игры ⚽ Нужно сделать НИР?

Look at other dictionaries:

  • Meteorite — Météorite Pour les articles homonymes, voir Météorite (homonymie). Météorite de Willamette. Une météorite est un corps matériel extraterrestre …   Wikipédia en Français

  • météorite — [ meteɔrit ] n. m. ou f. • 1830; de météore ♦ Astron. Fragment de corps céleste qui tombe sur la Terre ou sur un astre quelconque. ⇒ aérolithe. Bombardement de météorites. « Un jour, elle m apporte une pierre couleur de fer, lisse et lourde. C… …   Encyclopédie Universelle

  • Meteorite — Me te*or*ite, n. [Cf. F. m[ e]t[ e]orite.] (Min.) A mass of stone or iron which has fallen to the earth from space; an a[ e]rolite. [1913 Webster] Note: Meteorites usually show a pitted surface with a fused crust, caused by the heat developed in… …   The Collaborative International Dictionary of English

  • Meteorite — Meteorite, Meteorsteine, Mondsteine, Aerolithe, kosmische Körper, welche in Klumpen oder Staubform, zumeist unter starkem Geräusch und in erhöhter Temperatur auf die Erde niederfallen. Ihre Beschaffenheit ähnelt vielfach irdischen Gesteinen,… …   Lexikon der gesamten Technik

  • meteorite — s.f. o m. [der. di meteora ]. (astron.) [corpo solido costituito prevalentemente da materiali a base di silicati e da leghe ferro nichel, che, provenendo dagli spazi interplanetari, raggiunge la superficie della Terra] ▶◀ ‖ bolide, meteora …   Enciclopedia Italiana

  • meteorite — (n.) rock that falls to earth, after streaking across the sky as a meteor, 1818, from METEOR (Cf. meteor) + ITE (Cf. ite) …   Etymology dictionary

  • meteorite — s. f. [Astronomia] O mesmo que meteorito.   ‣ Etimologia: meteoro + ite …   Dicionário da Língua Portuguesa

  • meteorite — ► NOUN ▪ a piece of rock or metal that has fallen to the earth from space …   English terms dictionary

  • meteorite — [mēt′ē ər īt΄] n. that part of a relatively large meteoroid that survives passage through the atmosphere and falls to the surface of a planet or moon as a mass of metal or stone meteoritic [mēt′mə ərit′ik] adj …   English World dictionary

  • Meteorite — This article is about debris from space that survives impact with the ground. For other uses of Meteor and Meteors , see Meteor (disambiguation). For popular applications, see Falling star. For the fictional superhero in the Marvel Comics… …   Wikipedia

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”