earth

earth
/errth/, n.
1. (often cap.) the planet third in order from the sun, having an equatorial diameter of 7926 mi. (12,755 km) and a polar diameter of 7900 mi. (12,714 km), a mean distance from the sun of 92.9 million mi. (149.6 million km), and a period of revolution of 365.26 days, and having one satellite. See table under planet.
2. the inhabitants of this planet, esp. the human inhabitants: The whole earth rejoiced.
3. this planet as the habitation of humans, often in contrast to heaven and hell: to create a hell on earth.
4. the surface of this planet: to fall to earth.
5. the solid matter of this planet; dry land; ground.
6. soil and dirt, as distinguished from rock and sand; the softer part of the land.
7. the hole of a burrowing animal; lair.
8. Chem. any of several metallic oxides that are difficult to reduce, as alumina, zirconia, and yttria. Cf. alkaline earth, rare earth.
9. Also called earth color. Fine Arts. any of various pigments consisting chiefly of iron oxides and tending toward brown in hue.
10. Chiefly Brit. Elect. a ground.
11. Archaic. a land or country.
12. move heaven and earth. See heaven (def. 7).
13. on earth, in the world: Where on earth have you been?
14. run to earth,
a. Hunting. to chase (an animal) into its hole or burrow: to run a fox to earth.
b. to search out; track down: They ran the fugitive to earth in Algiers.
v.t.
15. Chiefly Brit. Elect. to ground.
[bef. 950; ME erthe, OE eorthe; c. G Erde, D aarde, ON jorth, Dan jord, Goth airtha]
Syn. 3. EARTH, GLOBE, WORLD are terms applied to the planet on which we dwell. EARTH is used esp. in speaking of a condition of existence contrasted with that in heaven or hell: those who are yet on earth. GLOBE formerly emphasized merely the roundness of the earth: to circumnavigate the globe. It is now used more like WORLD, with especial application to the inhabitants of the earth and their activities, interests, and concerns. In this sense, both GLOBE and WORLD are more inclusive than EARTH and are used more abstractly: the politics of the globe; the future of the world; One World.

* * *

I
Third planet in distance outward from the Sun.

Believed to be about 4.6 billion years old, it is some 92,960,000 mi (149,600,000 km) from the Sun. It orbits the Sun at a speed of 18.5 mi (29.8 km) per second, making one complete revolution in 365.25 days. As it revolves, it spins on its axis, rotating once every 23 hours 56 minutes 4 seconds. The fifth largest planet of the solar system, it has an equatorial circumference of 24,902 mi (40,076 km). Its total surface area is roughly 197,000,000 sq mi (509,600,000 sq km), of which about 29% is land. Earth's atmosphere consists of a mixture of gases, chiefly nitrogen and oxygen. Its only natural satellite, the Moon, orbits the planet at a distance of about 238,860 mi (384,400 km). Earth's surface is traditionally subdivided into seven continental masses: Africa, Antarctica, Asia, Australia, Europe, North America, and South America. These continents are surrounded by the so-called World Ocean, which is broken down into three major bodies: the Atlantic, Pacific, and Indian oceans. Broadly speaking, Earth's interior consists of two regions: a core composed largely of molten, iron-rich metallic alloy; and a rocky shell of silicate minerals comprising both the mantle and crust (See also Moho; lithosphere). Fluid motions in the electrically conductive outer core generate a magnetic field around Earth that is responsible for the Van Allen radiation belts. According to the theory of plate tectonics, the crust and upper mantle are divided into a number of large and small plates that float on and travel independently of the lower mantle. Plate motions are responsible for continental drift and seafloor spreading and for most volcanic and seismic activity on Earth.
II
(as used in expressions)
Earth Resources Technology Satellites
Earth crossing asteroid
baked earth

* * *

planet
Introduction
 third planet from the Sun and the fifth in the solar system in terms of size and mass. Its single most outstanding feature is that its near-surface environments are the only places in the universe known to harbour life. It is designated by the symbol ♁. Earth's name in English, the international language of astronomy, derives from Old English and Germanic words for ground and earth, and it is the only name for a planet of the solar system that does not come from Greco-Roman mythology.

      Since the Copernican revolution of the 16th century, at which time the Polish astronomer Nicolaus Copernicus (Copernicus, Nicolaus) proposed a Sun-centred model (Copernican system) of the universe (see heliocentric system), enlightened thinkers have regarded Earth as a planet like the others of the solar system. Concurrent sea voyages provided practical proof that Earth is a globe, just as Galileo's use of his newly invented telescope in the early 17th century soon showed various other planets to be globes as well. It was only after the dawn of the space age, however, when photographs from rockets and orbiting spacecraft first captured the dramatic curvature of Earth's horizon, that the conception of Earth as a roughly spherical planet rather than as a flat entity was verified by direct human observation. Humans first witnessed Earth as a complete orb floating in the inky blackness of space in December 1968 when Apollo (Apollo program) 8 carried astronauts around the Moon. Robotic space probes on their way to destinations beyond Earth, such as the Galileo and the Near Earth Asteroid Rendezvous (NEAR) spacecraft in the 1990s, also looked back with their cameras to provide other unique portraits of the planet.

      Viewed from another planet in the solar system, Earth would appear bright and bluish in colour. Easiest to see through a large telescope would be its atmospheric features, chiefly the swirling white cloud patterns of midlatitude and tropical storms, ranged in roughly latitudinal belts around the planet. The polar regions also would appear a brilliant white, because of the clouds above and the snow and ice below. Beneath the changing patterns of clouds would appear the much darker blue-black oceans, interrupted by occasional tawny patches of desert lands. The green landscapes that harbour most human life would not be easily seen from space. Not only do they constitute a modest fraction of the land area, which itself is less than one-third of Earth's surface, but they are often obscured by clouds. Over the course of the seasons, some changes in the storm patterns and cloud belts on Earth would be observed. Also prominent would be the growth and recession of the winter snowcap across land areas of the Northern Hemisphere.

      Scientists have applied the full battery of modern instrumentation to studying Earth in ways that have not yet been possible for the other planets; thus, much more is known about its structure and composition. This detailed knowledge, in turn, provides deeper insight into the mechanisms by which planets in general cool down, by which their magnetic fields are generated, and by which the separation of lighter elements from heavier ones as planets develop their internal structure releases additional energy for geologic processes and alters crustal compositions.

      It is convenient to consider separate parts of Earth in terms of concentric, roughly spherical layers. Extending from the interior outward, these are the core, the mantle, the crust (including the rocky surface), the hydrosphere (predominantly the oceans, which fill in low places in the crust), the atmosphere (itself divided into spherical zones such as the troposphere, where weather occurs, and the stratosphere, where lies the ozone layer that shields Earth's surface and its organisms against the Sun's ultraviolet rays), and the magnetosphere (an enormous region in space where Earth's magnetic field dominates the behaviour of electrically charged particles coming from the Sun).

      Knowledge about these divisions is summarized in this astronomically oriented overview. The discussion complements other treatments oriented to the Earth sciences and life sciences. Earth's figure and dimensions are discussed in the article geodesy. Its magnetic field is treated in the article geomagnetic field. The early evolution of the solid Earth and its atmosphere and oceans is covered in geologic history of Earth (Earth, geologic history of). The geologic and biological development of Earth, including its surface features and the processes by which they are created and modified, are discussed in geochronology, continental landform, and plate tectonics. The behaviour of the atmosphere and of its tenuous, ionized outer reaches is treated in atmosphere (atmosphere, evolution of), while the water cycle and major hydrologic features are described in hydrosphere, ocean, and river. The solid Earth as a field of study is covered in geologic sciences (geology), the methods and instruments employed to investigate Earth's surface and interior are discussed in Earth exploration, and the history of the study of Earth from antiquity to modern times is surveyed in Earth sciences. The global ecosystem of living organisms and their life-supporting stratum are detailed in biosphere.

Basic planetary data
      The mean distance of Earth from the Sun is about 150 million km (93 million miles). The planet orbits the Sun in a path that is presently more nearly a circle (less eccentric) than are the orbits of all but two of the other planets, Venus and Neptune. Earth makes one revolution, or one complete orbit of the Sun, in about 365.25 days. The direction of revolution—counterclockwise as viewed down from the north—is in the same sense, or direction, as the rotation of the Sun; Earth's spin, or rotation about its axis, is also in the same sense, which is called direct or prograde. The rotation period, or length of a sidereal day (see day and sidereal time)—23 hours, 56 minutes, and 4 seconds—is similar to that of Mars. Jupiter and most asteroids have days less than half as long, while Mercury and Venus have days more nearly comparable to their orbital periods. The 23.5° tilt, or inclination, of Earth's axis to its orbital plane, also typical, results in greater heating and more hours of daylight in one hemisphere or the other over the course of a year and so is responsible for the cyclic change of seasons.

      With an equatorial radius of 6,378 km (3,963 miles), Earth is the largest of the four inner, terrestrial (rocky) planets, but it is considerably smaller than the gas giants of the outer solar system. Earth has a single satellite, the moon, which orbits the planet at a mean distance of about 384,400 km (238,900 miles). The Moon is one of the bigger natural satellites in the solar system; only the giant planets have moons comparable or larger in size. Some planetary astronomers consider the Earth-Moon system a double planet, with some similarity in that regard to the dwarf planet Pluto and its largest moon, Charon.

      Earth's gravitational field (see gravitation) is manifested as the attractive force acting on a free body at rest, causing it to accelerate in the general direction of the centre of the planet. Departures from the spherical shape and the effect of Earth's rotation cause gravity to vary with latitude over the terrestrial surface. The average gravitational acceleration at sea level is about 980 cm/sec2 (32.2 feet/sec2).

       Planetary data for EarthEarth's gravity keeps the Moon in its orbit around the planet and also generates tides (tide) in the solid body of the Moon. Such deformations are manifested in the form of slight bulges at the lunar surface, detectable only by sensitive instruments. In turn, the Moon's mass—relatively large for a natural satellite—exerts a gravitational force that causes tides on Earth. The Sun, much more distant but vastly more massive, also raises tides on Earth. (See celestial mechanics: Tidal evolution (celestial mechanics).) The tides are most apparent as the twice-daily and daily rises and falls of the ocean water, although tidal deformations (Earth tide) occur in the solid Earth and in the atmosphere as well (see tide). The movement of the water throughout the ocean basins as a result of the tides (as well as, to a lesser extent, the tidal distortion of the solid Earth) dissipates orbital kinetic energy as heat, producing a gradual slowing of Earth's rotation and a spiraling outward of the Moon's orbit. Currently this slowing lengthens the day by a few thousandths of a second per century, but the rate of slowing varies with time as plate tectonics and sea-level changes alter the areas covered by inland bays and shallow seas. (For additional orbital and physical data, see the table (Planetary data for Earth).)

The atmosphere and hydrosphere
      The blankets of volatile gases and liquids near and above the surface of Earth are, along with solar energy, of prime importance to the sustenance of life on Earth. They are distributed and recycled throughout the atmosphere and hydrosphere of the planet.

The atmosphere
      Earth is surrounded by a relatively thin atmosphere (commonly called air) consisting of a mixture of gases, primarily molecular nitrogen (78 percent) and molecular oxygen (21 percent). Also present are much smaller amounts of gases such as argon (nearly 1 percent), water vapour (averaging 1 percent but highly variable in time and location), carbon dioxide (0.037 percent [370 parts per million] and presently rising), methane (0.00015 percent [1.5 parts per million]), and others, along with minute solid and liquid particles in suspension.

      Because Earth has a weak gravitational field (by virtue of its size) and warm atmospheric temperatures (due to its proximity to the Sun) compared with the giant planets, it lacks the most common gases in the universe that they possess: hydrogen and helium. Whereas both the Sun and Jupiter are composed predominantly of these two elements, they could not be retained long on early Earth and rapidly evaporated into interplanetary space. The high oxygen content of Earth's atmosphere is out of the ordinary. Oxygen is a highly reactive gas that, under most planetary conditions, would be combined with other chemicals in the atmosphere, surface, and crust. It is in fact supplied continuously by biological processes; without life, there would be virtually no free oxygen. The 1.5 parts per million of methane in the atmosphere is also far out of chemical equilibrium with the atmosphere and crust: it, too, is of biological origin, with the contribution by human activities far outweighing others.

      The gases of the atmosphere extend from the surface of Earth to heights of thousands of kilometres, eventually merging with the solar wind—a stream of charged particles that flows outward from the outermost regions of the Sun. The composition of the atmosphere is more or less constant with height to an altitude of about 100 km (60 miles), with particular exceptions being water vapour and ozone.

      The atmosphere is commonly described in terms of distinct layers, or regions. Most of the atmosphere is concentrated in the troposphere, which extends from the surface to an altitude of about 10–15 km (6–9 miles), depending on latitude and season. The behaviour of the gases in this layer is controlled by convection. This process involves the turbulent, overturning motions resulting from buoyancy of near-surface air that is warmed by the Sun. Convection maintains a decreasing vertical temperature gradient—i.e., a temperature decline with altitude—of roughly 6 °C (10.8 °F) per km through the troposphere. At the top of the troposphere, which is called the tropopause, temperatures have fallen to about −80 °C (−112 °F). The troposphere is the region where nearly all water vapour exists and essentially all weather occurs.

      The dry, tenuous stratosphere lies above the troposphere and extends to an altitude of about 50 km (30 miles). Convective motions are weak or absent in the stratosphere; motions instead tend to be horizontally oriented. The temperature in this layer increases with altitude.

      In the upper stratospheric regions, absorption of ultraviolet light (ultraviolet radiation) from the Sun breaks down molecular oxygen (O2); recombination of single oxygen atoms with O2 molecules into ozone (O3) creates the shielding ozone layer (ozonosphere).

      Above the relatively warm stratopause is the even more tenuous mesosphere, in which temperatures again decline with altitude to 80–90 km (50–56 miles) above the surface, where the mesopause is defined. The minimum temperature attained there is extremely variable with season. Temperatures then rise with increasing height through the overlying layer known as the thermosphere. Also above about 80–90 km there is an increasing fraction of charged, or ionized, particles, which from this altitude upward defines the ionosphere (ionosphere and magnetosphere). Spectacular visible auroras are generated in this region, particularly along approximately circular zones around the poles, by the interaction of nitrogen and oxygen atoms in the atmosphere with episodic bursts of energetic particles originating from the Sun.

      Earth's general atmospheric circulation is driven by the energy of sunlight (solar radiation), which is more abundant in equatorial latitudes. Movement of this heat toward the poles is strongly affected by Earth's rapid rotation and the associated Coriolis force at latitudes away from the Equator (which adds an east-west component to the direction of the winds), resulting in multiple cells of circulating air in each hemisphere. Instabilities (perturbations in the atmospheric flow that grow with time) produce the characteristic high-pressure areas and low-pressure storms of the midlatitudes as well as the fast, eastward-moving jet streams of the upper troposphere that guide the paths of storms. The oceans are massive reservoirs of heat that act largely to smooth out variations in Earth's global temperatures, but their slowly changing currents and temperatures also influence weather and climate, as in the El Niño/Southern Oscillation weather phenomenon (see climate: El Niño/Southern Oscillation (climate) and ocean: El Niño/Southern Oscillation and climatic change (ocean)).

      Earth's atmosphere is not a static feature of the environment. Rather its composition has evolved over geologic time in concert with life and is changing more rapidly today in response to human activities. Roughly halfway through the history of Earth, the atmosphere's unusually high abundance of free oxygen began to develop, through photosynthesis by cyanobacteria (see blue-green algae) and saturation of natural surface sinks of oxygen (e.g., relatively oxygen-poor minerals and hydrogen-rich gases exuded from volcanoes). Accumulation of oxygen made it possible for complex cells, which consume oxygen during metabolism and of which all plants and animals are composed, to develop (see eukaryote).

      Earth's climate at any location varies with the seasons, but there are also longer-term variations in global climate. Volcanic explosions, such as the 1991 eruption of Mount Pinatubo (Pinatubo, Mount) in the Philippines, can inject great quantities of dust particles into the stratosphere, which remain suspended for years, decreasing atmospheric transparency and resulting in measurable cooling worldwide. Much rarer, giant impacts of asteroids and comets can produce even more profound effects, including severe reductions in sunlight for months or years, such as many scientists believe led to the mass extinction of living species at the end of the Cretaceous Period, 65 million years ago. (For additional information on the risks posed by cosmic impacts and the chances of their occurrence, see Earth impact hazard.) The dominant climate variations observed in the recent geologic record are the ice ages, which are linked to variations in Earth's tilt and its orbital geometry with respect to the Sun.

      The physics of hydrogen fusion leads astronomers to conclude that the Sun was 30 percent less luminous during the earliest history of Earth than it is today. Hence, all else being equal, the oceans should have been frozen. Observations of Earth's planetary neighbours, Mars and Venus, and estimates of the carbon locked in Earth's crust at present suggest that there was much more carbon dioxide in Earth's atmosphere during earlier periods. This would have enhanced warming of the surface via the greenhouse effect and so allowed the oceans to remain liquid.

      Today there is 100,000 times more carbon dioxide buried in carbonate rocks (carbonate mineral) in Earth's crust than in the atmosphere, in sharp contrast to Venus, whose atmospheric evolution followed a different course. On Earth, the formation of carbonate shells by marine life is the principal mechanism for transforming carbon dioxide to carbonates; abiotic processes involving liquid water also produce carbonates, albeit more slowly. On Venus, however, life never had the chance to arise and to generate carbonates. Because of the planet's location in the solar system, early Venus received 10–20 percent more sunlight than falls on Earth even today, despite the fainter young Sun at the time. Most planetary scientists believe that the elevated surface temperature that resulted kept water from condensing to a liquid. Instead, it remained in the atmosphere as water vapour, which, like carbon dioxide, is an efficient greenhouse gas. Together the two gases caused surface temperatures to rise even higher so that massive amounts of water escaped to the stratosphere, where it was dissociated by solar ultraviolet radiation. With conditions now too hot and dry to permit abiotic carbonate formation, most or all of the planet's inventory of carbon remained in the atmosphere as carbon dioxide. Models predict that Earth may suffer the same fate in a billion years, when the Sun exceeds its present brightness by 10–20 percent.

      Between the late 1950s and the end of the 20th century, the amount of carbon dioxide in Earth's atmosphere increased by more than 15 percent because of the burning of fossil fuels (e.g., coal, oil, and natural gas) and the destruction of tropical rainforests, such as that of the Amazon River basin. Computer models predict that a net doubling of carbon dioxide by the middle of the 21st century could lead to a global warming of 1.5–4.5 °C (2.7–8.1 °F) averaged over the planet, which would have profound effects on sea level and agriculture. Although this conclusion has been criticized by some on the basis that the warming observed so far has not kept pace with the projection, analyses of ocean temperature data have suggested that much of the warming during the 20th century actually occurred in the oceans themselves—and will eventually appear in the atmosphere.

      Another present concern regarding the atmosphere is the impact of human activities on the stratospheric ozone layer (ozonosphere). Complex chemical reactions involving traces of man-made chlorofluorocarbons (CFCs) were found in the mid-1980s to be creating temporary holes in the ozone layer, particularly over Antarctica, during polar spring. Yet more disturbing was the discovery of a growing depletion of ozone over the highly populated temperate latitudes, since the short-wavelength ultraviolet radiation that the ozone layer effectively absorbs has been found to cause skin cancer. International agreements in place to halt the production of the most egregious ozone-destroying CFCs will eventually halt and reverse the depletion, but only by the middle of the 21st century, because of the long residence time of these chemicals in the stratosphere.

 Earth's hydrosphere is a discontinuous layer of water at or near the planet's surface; it includes all liquid and frozen surface waters, groundwater held in soil and rock, and atmospheric water vapour. Unique within the solar system, the hydrosphere is essential to all life as it is presently understood. Almost 71 percent of Earth's surface is covered by saltwater oceans, with a volume of about 1.4 billion cubic km (336 million cubic miles) and an average temperature of about 4 °C (39.2 °F), not far above the freezing point of water. The oceans contain about 97 percent of the planet's water volume. The remainder occurs as fresh water, three-quarters of which is locked up in the form of ice at polar latitudes. Most of the remaining fresh water is groundwater held in soils and rocks; less than 1 percent of it occurs in lakes and rivers. In terms of percentage, atmospheric water vapour is negligible, but the transport of water evaporated from the oceans onto land surfaces is an integral part of the hydrologic cycle that renews and sustains life.

      The hydrologic cycle involves the transfer of water from the oceans through the atmosphere to the continents and back to the oceans over and beneath the land surface. The cycle includes processes such as precipitation, evaporation, transpiration, infiltration, percolation, and runoff. These processes operate throughout the entire hydrosphere, which extends from about 15 km (9 miles) into the atmosphere to roughly 5 km (3 miles) into the crust.

      About one-third of the solar energy that reaches Earth's surface is expended on evaporating ocean water. The resulting atmospheric moisture and humidity condense into clouds, rain, snow, and dew. Moisture is a crucial factor in determining weather. It is the driving force behind storms and is responsible for separating electrical charge, which is the cause of lightning and thus of natural wildland fires, which have an important role in some ecosystems. Moisture wets the land, replenishes subterranean aquifers, chemically weathers (weathering) the rocks, erodes (erosion) the landscape, nourishes life, and fills the rivers, which carry dissolved chemicals and sediments back into the oceans.

      Water also plays a vital role in the carbon (carbon dioxide) dioxide cycle (a part of the more inclusive carbon cycle). Under the action of water and dissolved carbon dioxide, calcium is weathered from continental rocks and carried to the oceans, where it combines to form calcium carbonates (including shells of marine life). Eventually the carbonates are deposited on the seafloor and are lithified to form limestones. Some of these carbonate rocks are later dragged deep into Earth's interior by the global process of plate tectonics (see below The outer shell (Earth)) and melted, resulting in a rerelease of carbon dioxide (from volcanoes, for example) into the atmosphere. Cyclic processing of water, carbon dioxide, and oxygen through geologic and biological systems on Earth has been fundamental to maintaining the habitability of the planet with time and to shaping the erosion and weathering of the continents, and it contrasts sharply with the lack of such processes on Venus. (Evidence of past episodes of liquid water erosion—and possibly limited amounts of such erosion today—has been found on Mars.)

The outer shell
      Earth's outermost, rigid, rocky layer is called the crust. It is composed of low-density, easily melted rocks; the continental crust is predominantly granitic rock (see granite), while composition of the oceanic crust corresponds mainly to that of basalt and gabbro. Analyses of seismic waves, generated by earthquakes within Earth's interior, show that the crust extends about 50 km (30 miles) beneath the continents but only 5–10 km (3–6 miles) beneath the ocean floors.

      At the base of the crust, a sharp change in the observed behaviour of seismic waves marks the interface with the mantle. The mantle is composed of denser rocks, on which the rocks of the crust float. On geologic timescales, the mantle behaves as a very viscous fluid and responds to stress by flowing. Together the uppermost mantle and the crust act mechanically as a single rigid layer, called the lithosphere.

 The lithospheric outer shell of Earth is not one continuous piece but is broken, like a slightly cracked eggshell, into about a dozen major separate rigid blocks, or plates. There are two types of plates, oceanic and continental. An example of an oceanic plate is the Pacific Plate, which extends from the East Pacific Rise to the deep-sea trenches (deep-sea trench) bordering the western part of the Pacific basin. A continental plate is exemplified by the North American Plate, which includes North America as well as the oceanic crust between it and a portion of the Mid-Atlantic Ridge. The latter is an enormous submarine mountain chain that extends down the axis of the Atlantic basin, passing midway between Africa and North and South America.

 The lithospheric plates are about 60 km (35 miles) thick beneath the oceans and 100–200 km (60–120 miles) beneath the continents. (It should be noted that these thicknesses are defined by the mechanical rigidity of the lithospheric material. They do not correspond to the thickness of the crust, which is defined at its base by the discontinuity in seismic wave behaviour, as cited above.) They ride on a weak, perhaps partially molten, layer of the upper mantle called the asthenosphere. Slow convection currents deep within the mantle generated by radioactive heating of the interior drive lateral movements of the plates (and the continents on top of them) at a rate of several centimetres per year. The plates interact along their margins, and these boundaries are classified into three general types on the basis of the relative motions of the adjacent plates: divergent, convergent, and transform (or strike-slip).

      In areas of divergence, two plates move away from each other. Buoyant upwelling motions in the mantle force the plates apart at rift (rift valley) zones (such as along the middle of the Atlantic Ocean floor), where magmas from the underlying mantle rise to form new oceanic crustal rocks.

      Lithospheric plates move toward each other along convergent boundaries. When a continental plate and an oceanic plate come together, the leading edge of the oceanic plate is forced beneath the continental plate and down into the asthenosphere—a process called subduction. Only the thinner, denser slabs of oceanic crust will subduct, however. When two thicker, more buoyant continents come together at convergent zones, they resist subduction and tend to buckle, producing great mountain (orogeny) ranges. The Himalayas, along with the adjacent Plateau of Tibet (Tibet, Plateau of), were formed during such a continent-continent collision, when India was carried into the Eurasian Plate by relative motion of the Indian-Australian Plate.

      At the third type of plate boundary, the transform variety, two plates slide parallel to one another in opposite directions. These areas are often associated with high seismicity, as stresses that build up in the sliding crustal slabs are released at intervals to generate earthquakes (earthquake). The San Andreas Fault in California is an example of this type of boundary, which is also known as a fault or fracture zone (see submarine fracture zone).

      Most of Earth's active tectonic (diastrophism) processes, including nearly all earthquakes, occur near plate margins. Volcanoes (volcano) form along zones of subduction (subduction zone), because the oceanic crust tends to be remelted as it descends into the hot mantle and then rises to the surface as lava. Chains of active, often explosive volcanoes are thus formed in such places as the western Pacific and the west coasts of the Americas. Older mountain ranges, eroded by weathering and runoff, mark zones of earlier plate-margin activity. The oldest, most geologically stable parts of Earth are the central cores of some continents (such as Australia, parts of Africa, and northern North America). Called continental shields, they are regions where mountain building, faulting, and other tectonic processes are diminished compared with the activity that occurs at the boundaries between plates. Because of their stability, erosion has had the time to flatten the topography of continental shields. It is also on the shields that geologic evidence of crater scars from ancient impacts of asteroids and comets is better-preserved. Even there, however, tectonic processes and the action of water have erased many ancient features. In contrast, much of the oceanic crust is substantially younger (tens of millions of years old), and none dates back more than 200 million years.

      This conceptual framework in which scientists now understand the evolution of Earth's lithosphere—termed plate tectonics—is almost universally accepted, although many details remain to be worked out. For example, scientists have yet to reach a general agreement as to when the original continental (continent) cores formed or how long ago modern plate-tectonic processes began to operate. Certainly the processes of internal convection, segregation of minerals by partial melting and recrystallization, and basaltic volcanism were operating more vigorously in the first billion years of Earth's history, when the planet's interior was much hotter than it is today; nevertheless, how the surface landmasses were formed and were dispersed may have been different.

      Once major continental shields grew, plate tectonics was characterized by the cyclic assembly and breakup of supercontinents created by the amalgamation of many smaller continental cores and island arcs. Scientists have identified two such cycles in the geologic record. A supercontinent began breaking up about 700 million years ago, in late Precambrian time, into several major continents, but by about 250 million years ago, near the beginning of the Triassic Period, the continued drifting of these continents resulted in their fusion again into a single supercontinental landmass called Pangea. Some 70 million years later, Pangea began to fragment, gradually giving rise to today's continental configuration. The distribution is still asymmetric, with continents predominantly located in the Northern Hemisphere opposite the Pacific basin.

      Startlingly, of the four terrestrial planets, only Earth shows evidence of long-term, pervasive plate tectonics. Both Venus and Mars exhibit geology dominated by basaltic volcanism on a largely immovable crust, with only faint hints of possibly limited episodes of horizontal plate motion. Mercury is intrinsically much denser than the other terrestrial planets, which implies a larger metallic core; its surface is mostly covered with impact craters, but it also shows a global pattern of scarps suggesting shrinkage of the planet, associated perhaps with interior cooling. Apparently essential to the kind of plate tectonics that occurs on Earth are large planetary size (hence, high heat flow and thin crust), which eliminates Mars, and pervasive crustal water to soften the rock, which Venus lost very early in its history. Although Earth is indeed geologically active and hence possesses a youthful surface, Venus's surface may have been completely renewed by global basaltic volcanism within the past billion years, and small portions of Mars's surface may have experienced very recent erosion from liquid water or landslides.

The interior
      More than 90 percent of Earth's mass is composed of iron, oxygen, silicon, and magnesium, elements that can form the crystalline minerals known as silicate (silicate mineral)s. Nevertheless, in chemical and mineralogical composition, as in physical properties, Earth is far from homogeneous. Apart from the superficial lateral differences near the surface (i.e., in the compositions of the continental and oceanic crusts), Earth's principal differences vary with distance toward the centre. This is due to increasing temperatures and pressures and to the original segregation of materials, soon after Earth accreted from the solar nebula about 4.56 billion years ago, into a metal-rich core, a silicate-rich mantle, and the more highly refined crustal rocks. Earth is geochemically differentiated to a great extent (see below Planetary differentiation (Earth)). Crustal rocks contain several times as much of the rock-forming element aluminum as does the rest of the solid Earth and many dozens of times as much uranium. On the other hand, the crust, which accounts for a mere 0.4 percent of Earth's mass, contains less than 0.1 percent of its iron. Between 85 and 90 percent of Earth's iron is concentrated in the core.

      The increasing pressure with depth causes phase changes in crustal rocks at depths between 5 and 50 km (3 and 30 miles), which marks the top of the upper mantle, as mentioned above. This transition area is called the Mohoroviić discontinuity, or Moho. Most basaltic magmas are generated in the upper mantle at depths of hundreds of kilometres. The upper mantle, which is rich in the olivine, pyroxene, and silicate perovskite minerals, shows significant lateral differences in composition. A large fraction of Earth's interior, from a depth of about 650 km (400 miles) down to 2,900 km (1,800 miles), consists of the lower mantle, which is composed chiefly of magnesium- and iron-bearing silicates, including the high-pressure equivalents of olivine and pyroxene.

      The mantle is not static but rather churns slowly in convective motions, with hotter material rising up and cooler material sinking; through this process, Earth gradually loses its internal heat. In addition to being the driving force of horizontal plate motion, mantle convection is manifested in the occurrence of temporary superplumes—huge, rising jets of hot, partially molten rock—which may originate from a deep layer near the core-mantle interface. Much larger than ordinary thermal plumes, such as that associated with the Hawaiian island chain in the central Pacific (see volcano: Intraplate volcanism (volcano)), superplumes may have had profound effects on Earth's geologic history and even on its climate. One outburst of global volcanism about 65 million years ago, which created the vast flood basalt deposits known as the Deccan Traps on the Indian subcontinent (see plateau), may have been associated with a superplume, though this model is far from universally accepted.

      With a radius of almost 3,500 km (2,200 miles), Earth's core is about the size of the entire planet Mars. About one-third of Earth's mass is contained in the core, most of which is liquid iron alloyed with some lighter, cosmically abundant components (e.g., sulfur, oxygen, and, controversially, even hydrogen). Its liquid nature is revealed by the failure of shear-type seismic waves to penetrate the core. A small, central part of the core, however, below a depth of about 5,100 km (3,200 miles), is solid. Temperatures (temperature) in the core are extremely hot, ranging from 4,000–5,000 K (roughly 6,700–8,500 °F; 3,700–4,700 °C) at the outer part of the core to 5,000–7,000 K (8,500–12,100 °F; 4,700–6,700 °C) in the centre, comparable to the surface of the Sun. Large uncertainties in temperature arise from questions as to which compounds form alloys with iron in the core, and more recent data favour the lower end of the temperature estimates for the inner core. The core's reservoir of heat may contribute as much as one-fifth of all the internal heat that ultimately flows to the surface of Earth. The basic structure of Earth—crust, mantle, and core—appears to be replicated on the other terrestrial planets, though with substantial variations in the relative size of each region.

The geomagnetic field and magnetosphere
      Helical fluid motions in Earth's electrically conducting liquid outer core have an electromagnetic dynamo effect, giving rise to the geomagnetic field. The planet's sizable, hot core, along with its rapid spin, probably accounts for the exceptional strength of the magnetic field of Earth compared with those of the other terrestrial planets. Venus, for example, which has a metallic core that may be similar to Earth's in size, rotates very slowly and has no detected intrinsic magnetic field. Mercury and Mars have only small intrinsic magnetic fields.

      Earth's main magnetic field permeates the planet and an enormous volume of space surrounding it. A great teardrop-shaped region of space called the magnetosphere is formed by the interaction of Earth's field with the solar wind. At a distance of about 65,000 km (40,000 miles) outward toward the Sun, the pressure of the solar wind is balanced by the geomagnetic field. This serves as an obstacle to the solar wind, and the flow of charged particles, or plasma, is deflected around Earth by the resulting bow shock. The magnetosphere so produced streams out into an elongated magnetotail that stretches several million kilometres downstream from Earth away from the Sun.

      Plasma particles from the solar wind can leak through the magnetopause, the sunward boundary of the magnetosphere, and populate its interior; charged particles from the Earth's ionosphere also enter the magnetosphere. The magnetotail can store for hours an enormous amount of energy—several billion megajoules, which is roughly equivalent to the yearly electricity production of many smaller countries). This occurs through a process called reconnection, in which the Sun's magnetic field, dragged into interplanetary space by the solar wind, becomes linked with the magnetic field in Earth's magnetosphere. The energy is released in dynamic structural reconfigurations of the magnetosphere, called geomagnetic substorms, which often result in the precipitation of energetic particles into the ionosphere, giving rise to fluorescing auroral (aurora) displays.

      Converging magnetic field lines fairly close to Earth can trap highly energetic particles so that they gyrate between the Northern and Southern hemispheres and slowly drift longitudinally around the planet in two concentric doughnut-shaped zones known as the Van Allen radiation belts. Many of the charged particles trapped in these belts are produced when energetic cosmic rays strike Earth's upper atmosphere, producing neutrons that then decay into electrons, which are negatively charged, and protons, which are positively charged. Others come from the solar wind or Earth's atmosphere. The inner radiation belt was detected in 1958 by the American physicist James Van Allen (Van Allen, James A.) and colleagues, using a Geiger-Müller counter aboard the first U.S. satellite, Explorer 1; the outer belt was distinguished by other U.S. and Soviet spacecraft launched the same year. Earth's magnetosphere has been extensively studied ever since, and space physicists have extended their studies of plasma processes to the vicinities of comets and other planets. (For additional information on the interaction of the Sun and Earth's charged particles and magnetic fields, see plasma: Solar-terrestrial forms (plasma).)

      An important characteristic of Earth's magnetic field is polarity reversal. In this process the direction of the dipole component reverses—i.e., the north magnetic pole becomes the south magnetic pole and vice versa. From studying the direction of magnetization of many rocks, geologists know that such reversals occur, without a discernible pattern, at intervals that range from tens of thousands of years to millions of years, though they are still uncertain about the mechanisms responsible. It is likely that during the changeover, which is believed to take a few thousand years, a nondipolar field remains, at a small fraction of the strength of the normal field. In the temporary absence of the dipole component, the solar wind would approach much closer to Earth, allowing particles that are normally deflected by the field or are trapped in its outer portions to reach the surface. The increase in particle radiation could lead to increased rates of genetic damage and thus of mutations or sterility in plants and animals, leading to the disappearance of some species. Scientists have looked for evidence of such changes in the fossil record at times of past field reversals, but the results have been inconclusive.

Clark R. Chapman Jonathan I. Lunine

Development of Earth's structure and composition
      The origin of Earth in its present form has long been the subject of intellectual interest, but since the mid-20th century scientists have made particularly significant advances both in concepts and in measurements. Analysis of the isotopes in meteorites and, in particular, of rocks brought back from the Moon by U.S. Apollo (Apollo program) astronauts have produced some of the major contributions. Other gains have come from geochemical research on terrestrial samples combined with the new understanding of internal processes brought on by the recognition of plate tectonics, study of the terrestrial planets as a group, and advances in numerical modeling of the physical processes that lead to planetary formation.

      The starting point in tracing planetary evolution (planet) is nucleosynthesis, the formation of the chemical elements on a cosmic scale. This includes the nuclear processes by which the lightest elements—mostly hydrogen and helium—were produced at the explosive birth of the universe (see big-bang model), some 14 billion years ago, and the subsequent formation of the heavier elements within stars (see chemical element: Origin of the elements (chemical element)). By analogy with what astronomers presently observe to happen in regions of star formation, it is thought that the solar system began as a cloud of gas and dust comprising such preexisting elements. Under its own gravitational attraction, the cloud collapsed into a rotating disk of matter, called the solar nebula. The collapse could have been initiated by a shock wave emanating from a nearby supernova, a violently exploding star, or by random density fluctuations in the cloud itself. Once sufficiently high pressures and densities were achieved in the compacted nebular core, nuclear fusion reactions within it could begin, giving birth to a star. The outer part of the rotating disk—the matter not incorporated into the new Sun—became the raw material for the planets and other orbiting bodies of the solar system. The birth of the Sun, which makes up more than 99.9 percent of the mass of the entire solar system, is taken to be the time at which the planets started to form, approximately 4.56 billion years ago.

Accretion of the early Earth
      As the gas making up the solar nebula beyond the Sun cooled with time, mineral grains are thought to have condensed and aggregated to form the earliest meteoritic material. In addition, as is suggested by the finding of anomalous concentrations of isotopes (isotope) in a few meteorites, solid material from outside the solar system, apparently existing prior to the formation of the Sun, was occasionally incorporated into these developing small bodies.

      The concentrations of isotopes that decay radioactively and of isotopes that are produced by radioactive decay (radioactivity) provide scientists the information required to determine when meteorites and the planets formed. For example, the concentrations of rubidium-87 (rubidium-strontium dating) and the strontium-87 into which it decays, or those of samarium-147 and its decay product neodymium-143, indicate that the oldest meteorites formed some 4.56 billion years ago. Other isotope studies demonstrate that Earth formed within, at most, a few tens of millions of years after the birth of the Sun.

      The most abundant elements in the Sun, hydrogen and helium, are severely lacking in the inner, terrestrial planets but are still abundant constituents of the large, gaseous, outer planets such as Jupiter and Saturn. It is thought that only at the distance of Jupiter and beyond—in the colder regions of the solar system, including the zone beyond Neptune in which comets originated (see Kuiper belt and Oort cloud)—could the more volatile substances, which also include water, carbon dioxide, and ammonia, condense and be retained in appreciable amounts during the formation of the planets. Nevertheless, when the relative abundances of the less volatile elements are compared for the Sun, for a class of primitive, largely chemically unaltered meteorites called CI carbonaceous chondrites (considered by many researchers the most pristine samples of original solar system material), and for the estimated composition of Earth, their values are all in close agreement. This is the basis for the chondritic model, which holds that Earth (and presumably the other terrestrial planets) was essentially built up from bodies made of such meteoritic material. This idea is corroborated by isotopic studies of rocks derived from interior regions of Earth considered to be little changed throughout the planet's history. Thus, it appears that the composition of Earth is roughly what would be expected given the observed elemental abundances in the Sun and accounting for the loss of the more volatile elements.

 The dust and grains that condensed out of the cooling nebular gas aggregated gravitationally to form larger fragments of rock. The chondritic meteorites (see chondrite) observed today are basically just such collections of grains and fragments that were compacted together into larger pieces. Through continued accretion, the smaller pieces formed boulders and asteroid-size bodies ( planetesimals) and, ultimately, bodies the size of the Moon and Mars. The larger the planetesimals grew, the greater their gravitational attraction and the more effectively they swept up additional particles and rock fragments while circling the Sun. Growth slowed when most bodies were lunar- and Mars-size because they were limited in number and hence effectively isolated one from another in their orbits. As Jupiter increased to its giant size, its powerful gravity perturbed these “embryos” of the terrestrial planets, elongating their orbits and allowing their incremental growth to approximately the mass of Earth to proceed over tens of millions of years.

      Stony meteorites and iron meteorites (those composed largely of iron alloyed with nickel and sulfur) both fall on Earth today, and both types are thought to have been present during the formation of the planetesimals that would accrete to become Earth. In other words, Earth seems to have accreted only after most, if not all, solid matter had already condensed. Thus, a wide range of minerals was included in the grains, the larger fragments, and even the planetesimals that were accumulated by the growing planet. Apparently, such an aggregation of dense metallic fragments and less dense rocky fragments is not very stable. Calculations based on the measured strengths of rocks indicate that the metallic fragments probably sank downward as Earth grew. Although the planet was relatively cold at this stage—less than 500 K (440 °F; 230 °C)—the rock was weak. This is an important point because it leads to the conclusion that Earth's metallic core began to form during accretion of the planet and probably before the planet had grown to one-fifth of its present volume.

Effects of planetesimal impacts
      During its accretion, Earth is thought to have been shock-heated by the impacts of meteorite-size bodies and larger planetesimals. For a meteorite collision, the heating is concentrated near the surface where the impact occurs, which allows the heat to radiate back into space. A planetesimal, however, can penetrate sufficiently deeply on impact to produce heating well beneath the surface. In addition, the debris formed on impact can blanket the planetary surface, which helps to retain heat inside the planet. Some scientists have suggested that, in this way, Earth may have become hot enough to begin melting after growing to less than 15 percent of its final volume.

      Among the planetesimals striking the forming Earth, at least one is considered to have been comparable in size to Mars. Although the details are not well understood, there is good evidence that the impact of such a large planetesimal created the Moon. Among the more persuasive indications is that the relative abundances of many trace elements in rocks from the Moon are close to the values obtained for Earth's mantle. Unless this is a fortuitous coincidence, it points to the Moon having been derived from the mantle. Computer simulations have shown that a glancing collision of a Mars-size planetary body could have been sufficient to excavate from Earth's interior the material that would form the Moon. Again, the evidence for such large collisions suggests that Earth was very effectively heated during accretion.

      It is apparent, then, that many processes contributing to the early development of Earth occurred almost simultaneously, within tens to hundreds of million of years after the Sun was formed. Meteorites and Earth were formed within this time, and the Moon, which has been dated at more than four billion years in age, apparently was formed in the same time period. Simultaneously, Earth's core was accumulating and may have been completely formed during the planet's growth period. In addition to the possible accretional heating caused by planetesimal impacts, the sinking of metal to form the core released enough gravitational energy to heat the entire planet by 1,000 K (1,800 °F; 1,000 °C) or more. Thus, once core formation began, Earth's interior became sufficiently hot to convect. Although it is not known whether or in what form plate tectonics was active at the surface, it seems quite possible that the underlying mantle convection began even before the planet had grown to its final dimensions. Only later in Earth's development did radioactivity become an important heat source as well.

Planetary differentiation
      Once hot, Earth's interior could begin its chemical evolution. For example, outgassing of a fraction of volatile substances that had been trapped in small amounts within the accreting planet probably formed the earliest atmosphere. Outgassing of water to Earth's surface began before 4.3 billion years ago, a time based on analysis of ancient zircons that show the effects of alteration by liquid water. In Earth's deepest interior, chemical reactions between the mantle and the core became possible. Perhaps the most important event for Earth's surface, however, was the formation of the earliest crust by partial melting of the interior. This chemical separation by partial melting and outgassing of volatiles is termed differentiation. As the interior differentiated, less-dense liquids rose from the melt toward the surface and crystallized to form crust.

      Uncertainty exists over when and how the continental crust began to grow, because the record of the first 600 million years has not been found. The oldest known rocks date to only about 4 billion years. Because these are metamorphic rocks—i.e., because they were changed by heat and pressure from preexisting crustal rocks at the time of their dated age—it can be inferred that crust was present earlier in Earth's history. In fact, two tiny grains of zircon from Australia have been dated at 4.28 billion and 4.4 billion years, but their relation to the formation of continental crust is uncertain.

      Although direct evidence is not available, indirect evidence derived from the compositions of rocks indicates that continental crust formed early. Isotopic analyses suggests that the average age of the present continental crust is about 2.5 billion years. Thus, in all probability, repeated partial melting of the upper mantle formed successively more refined, continent-like crustal rocks starting before 4 billion years ago. Over the first billion years, however, much of the continental crust that was formed appears to have been reincorporated into the mantle—the isotopic data infers that on average about one-third of the continental crust was recycled every billion years. As a result, only a few fragments of crust older than 3.5 billion years remain, virtually none older than 4 billion years.

      The process of partial melting and formation of crust, especially continental crust, leads to a depletion of certain elements (e.g., silicon and aluminum) from the mantle. Undepleted and thus relatively primitive regions still exist, making up about one-third to one-half of the mantle, according to the isotopic models. The distribution of depleted and undepleted regions, however, is uncertain. Although much (perhaps all) of the upper mantle appears to be depleted, it is not known whether depleted rocks also exist in the lower mantle.

      What is recognized is that Earth is still differentiating into chemically distinct layers or regions. This is most evident in the processes of plate tectonics that involve ongoing production of crust at divergent plate boundaries such as the midocean ridges. As this material is cycled back down into the mantle at subduction zones and then upward again, it continues to undergo chemical processing from basaltic to andesitic and eventually to granitic (continental) composition. Thus, chemical and thermal evolution of the interior, intimately connected through mantle convection, is still vigorously in progress some 4.56 billion years after the formation of the planet.

Raymond Jeanloz Jonathan I. Lunine

Additional Reading
Peter D. Ward and Don Brownlee, Rare Earth: Why Complex Life is Uncommon in the Universe, 2nd rev. ed. (2003), lays out for general readers a case for the uncommon nature of planet Earth. Jonathan I. Lunine, Earth: Evolution of a Habitable World (1999); and John J.W. Rogers, A History of the Earth (1993), provide comprehensive introductions to Earth from, respectively, planetological and geologic perspectives. William K. Hartmann, The History of Earth: An Illustrated Chronicle of an Evolving Planet (1991), is a lavishly illustrated introduction. A graduate-level text comprising excellent chapters on its subject matter is Robin M. Canup and Kevin Righter (eds.), Origin of the Earth and Moon (2000). A popular and balanced account of the physical causes of mass extinctions is given in Charles Frankel, The End of the Dinosaurs: Chicxulub Crater and Mass Extinctions (1999; originally published in French, 1996). The periodical Scientific American (monthly) is an excellent and accessible source for the latest thinking on Earth and planetary processes; for example, a good description of plate tectonics is found in Ian W.D. Dalziel, “Earth Before Pangea,” 272(1):58–63 (January 1995). The State of the World report (annual), published by the Worldwatch Institute, provides authoritative updates on Earth's vital signs for the general reader.Jonathan I. Lunine

* * *


Universalium. 2010.

Игры ⚽ Поможем сделать НИР
Synonyms:

Look at other dictionaries:

  • Earth — ([ e]rth), n. [AS. eor[eth]e; akin to OS. ertha, OFries. irthe, D. aarde, OHG. erda, G. erde, Icel. j[ o]r[eth], Sw. & Dan. jord, Goth. a[=i]r[thorn]a, OHG. ero, Gr. ?, adv., to earth, and perh. to E. ear to plow.] 1. The globe or planet which we …   The Collaborative International Dictionary of English

  • Earth — Earth, Wind Fire Saltar a navegación, búsqueda Earth, Wind Fire Información personal Origen Chicago, Illinois Información artística Género …   Wikipedia Español

  • earth — n Earth, world, universe, cosmos, macrocosm are comparable when they mean the entire area or extent of space in which man thinks of himself and of his fellow men as living and acting. Earth applies, however, only to part of what he knows by sight …   New Dictionary of Synonyms

  • earth — [ʉrth] n. [ME erthe < OE eorthe, akin to Ger erde < IE base * er > Gr era, earth, Welsh erw, field] 1. [often E ] the planet that we live on; terrestrial globe: it is the fifth largest planet of the solar system and the third in distance …   English World dictionary

  • Earth-X — Éditeur Marvel Comics Format Série limitée Date(s) de publication 1997 2000 Numéros 15 + sketchbook 4 tomes pour l édition française Scénariste(s) …   Wikipédia en Français

  • Earth X — Éditeur Marvel Comics Format série limitée Date(s) de publication 1997 2000 Numéros …   Wikipédia en Français

  • earth — ► NOUN 1) (also Earth) the planet on which we live. 2) the substance of the land surface; soil. 3) Brit. electrical connection to the ground, regarded as having zero electrical potential. 4) the underground lair of a badger or fox. 5) one of the… …   English terms dictionary

  • Earth — (engl. ‚Erde‘) bezeichnet: Earth (Band), eine Musikgruppe aus einem Subgenre des Metal einen früheren Namen der Band Black Sabbath Earth (Computerspiel), eine Computerspiel Reihe aus dem Genre der Echtzeitstrategiespiele Filme: Earth (1986),… …   Deutsch Wikipedia

  • Earth 2 — may refer to:* Earth Two, a parallel world in the DC Multiverse; the home of DC s Golden Age heroes and Earth 2 the current version. * Earth 2 (TV series), a science fiction TV series that aired in 1994 ndash;1995 * Earth 2 (album), an album by… …   Wikipedia

  • Earth — Основная информация …   Википедия

  • earth|y — «UR thee», adjective, earth|i|er, earth|i|est. 1. of earth or soil: »Potatoes have an earthy smell. 2. like earth or soil (in texture …   Useful english dictionary

Share the article and excerpts

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