Venus

Venus
/vee"neuhs/, n., pl. Venuses for 2.
1. an ancient Italian goddess of gardens and spring, identified by the Romans with Aphrodite as the goddess of love and beauty.
2. an exceptionally beautiful woman.
3. (sometimes l.c.) Archaeol. a statuette of a female figure, usually carved of ivory and typically having exaggerated breasts, belly, or buttocks, often found in Upper Paleolithic cultures from Siberia to France.
4. Astron. the planet second in order from the sun, having an equatorial diameter of 7521 miles (12,104 km), a mean distance from the sun of 67.2 million miles (108.2 million km), a period of revolution of 224.68 days, and no moons. It is the most brilliant planet in the solar system. See table under planet.
5. Chem. Obs. copper.
[ < L Venus, s. Vener- orig. a neut. common n. meaning "physical desire, sexual appetite," hence "qualities exciting desire, seductiveness, charm," "a goddess personifying sexual attractiveness"; c. Skt vanah desire, akin to WISH; cf. VENERATE, VENOM]

* * *

I
Roman goddess of cultivated fields and gardens, later associated with Aphrodite.

She was the daughter of Jupiter and Dione, the wife of Vulcan, and the mother of Cupid. She was famous for her romantic intrigues and affairs with both gods and mortals, and she became associated with many aspects of femininity. The planet Venus, originally the star of Ishtar, came to be named for Venus through her association with Ishtar. She has been a favourite subject in art since ancient times, notably in the statue known as Venus de Milo (с 150 BC)and in Sandro Botticelli's painting The Birth of Venus (с 1485).
II
Second major planet from the Sun.

Named for the Roman goddess, Venus is, after the Moon, the most brilliant natural object in the night sky. It comes closer to Earth
about 26 million mi (42 million km)
than any other planet. Its orbit around the Sun is nearly circular at a distance of 67 million mi (108 million km) and takes 225 days; its rotation, in retrograde motion, takes even longer (243 days). As viewed from Earth, Venus undergoes phase changes similar to the Moon's, going through one cycle of phases in 584 days. It is seen only near sunrise or sunset and has long been known as both the morning and the evening star. Venus is a near twin of Earth in size and mass but is completely enveloped by thick clouds of concentrated sulfuric-acid droplets. Its surface gravity is about 90% that of Earth. Its atmosphere is over 96% carbon dioxide, with a pressure about 90 times Earth's. The dense atmosphere and thick cloud layers trap incoming solar energy so efficiently that Venus has the highest surface temperature of any of the Sun's planets, about 860 °F (460 °C). Radar imaging indicates that its surface is dry and rocky, consisting mostly of gently rolling plains, broad depressions, and two large elevated regions analogous to continents on Earth; it also has impact craters, extensive lava fields, and massive shield volcanos. The interior is thought to be similar to that of Earth, with a metal core, a dense rocky mantle, and a less-dense rocky crust. Unlike Earth, Venus has no intrinsic magnetic field.
III
(as used in expressions)
Venus's flytrap
Venus flytrap
Williams Venus and Serena

* * *

▪ Roman goddess
 ancient Italian goddess associated with cultivated fields and gardens and later identified by the Romans with the Greek goddess of love, Aphrodite.

      Venus had no worship in Rome in early times, as the scholar Marcus Terentius Varro (116–27 BC) shows, attesting that he could find no mention of her name in old records. This is corroborated by the absence of any festival for her in the oldest Roman calendar and by her lack of a flamen (special priest). Her cult among the Latins, however, seems to be immemorial, for she had apparently at least two ancient temples, one at Lavinium, the other at Ardea, at which festivals of the Latin cities were held. Hence, it was no long step to bring her to Rome, apparently from Ardea itself. But how she came to be identified with so important a deity as Aphrodite remains a puzzle.

      That Venus' identification with Aphrodite took place fairly early is certain. A contributory reason for it is perhaps the date (August 19) of the foundation of one of her Roman temples. August 19 is the Vinalia Rustica, a festival of Jupiter; hence, he and Venus came to be associated, and this facilitated their equation, as father and daughter, with the Greek deities Zeus and Aphrodite. She was, therefore, also a daughter of Dione, was the wife of Vulcan, and was the mother of Cupid. In myth and legend she was famous for her romantic intrigues and affairs with both gods and mortals, and she became associated with many aspects, both positive and negative, of femininity. As Venus Verticordia, she was charged with the protection of chastity in women and girls. But the most important cause of the identification was the reception into Rome of the famous cult of Venus Erycina—i.e., of Aphrodite of Eryx (Erice) in Sicily—this cult itself resulting from the identification of an Oriental mother-goddess with the Greek deity. This reception took place during and shortly after the Second Punic War. A temple was dedicated to Venus Erycina on the Capitol in 215 BC and a second outside the Colline gate in 181 BC. The latter developed in a way reminiscent of the temple at Eryx with its harlots, becoming the place of worship of Roman courtesans, hence the title of dies meretricum (“prostitutes' day”) attached to April 23, the day of its foundation.

      The importance of the worship of Venus-Aphrodite was increased by the political ambitions of the gens Iulia, the clan of Julius Caesar and, by adoption, of Augustus. They claimed descent from Iulus (Ascanius), the son of Aeneas; Aeneas was the alleged founder of the temple of Eryx and, in some legends, of the city of Rome also. From the time of Homer onward, he was made the son of Aphrodite, so that his descent gave the Iulii divine origin. Others than the Iulii sought to connect themselves with a deity grown so popular and important, notably Gnaeus Pompeius, the triumvir. He dedicated a temple to Venus as Victrix (“Bringer of Victory”) in 55 BC. Julius Caesar's own temple (46 BC), however, was dedicated to Venus Genetrix, and as Genetrix (“Begetting Mother”) she was best known until the death of Nero in AD 68. But despite the extinction of the Julio-Claudian line, she remained popular, even with the emperors; Hadrian completed a temple of Venus at Rome in AD 135.

 As a native Italian deity, Venus had no myths of her own. She therefore took over those of Aphrodite and, through her, became identified with various foreign goddesses. The most noteworthy result of this development is perhaps the acquisition by the planet Venus of that name. The planet was at first the star of the Babylonian goddess Ishtar and thence of Aphrodite. Because of her association with love and with feminine beauty, the goddess Venus has been a favourite subject in art since ancient times; notable representations include the statue known as the “Venus de Milo” (c. 150 BC) and the painting “The Birth of Venus” (c. 1485; see photograph—>) by Sandro Botticelli.
 

planet
Introduction
 second planet from the Sun and sixth in the solar system in size and mass. No planet approaches closer to Earth than Venus; at its nearest it is the closest large body to Earth other than the Moon. Because Venus's orbit is nearer the Sun than Earth's, the planet is always roughly in the same direction in the sky as the Sun and can be seen only in the hours near sunrise or sunset. When it is visible, it is the most brilliant planet in the sky. Venus is designated by the symbol ♀.

 Venus was one of the five planets—along with Mercury, Mars, Jupiter, and Saturn—known in ancient times, and its motions were observed and studied for centuries prior to the invention of advanced astronomical instruments. Its appearances were recorded by the Babylonians, who equated it with the goddess Ishtar, about 3000 BC, and it also is mentioned prominently in the astronomical records of other ancient civilizations, including those of China, Central America, Egypt, and Greece. Like the planet Mercury, Venus was known in ancient Greece by two different names—Phosphorus (see Lucifer) when it appeared as a morning star and Hesperus when it appeared as an evening star. Its modern name comes from the Roman goddess of love and beauty (the Greek equivalent being Aphrodite), perhaps because of the planet's luminous jewel-like appearance.

  Planetary data for Venus Planetary data for VenusVenus has been called Earth's twin because of the similarities in their masses, sizes, and densities and their similar relative locations in the solar system. Because they presumably formed in the solar nebula from the same kind of rocky planetary building blocks, they also likely have similar overall chemical compositions. Early telescopic observations of the planet revealed a perpetual veil of clouds, suggestive of a substantial atmosphere and leading to popular speculation that Venus was a warm, wet world, perhaps similar to Earth during its prehistoric age of swampy carboniferous forests and abundant life. Scientists now know, however, that Venus and Earth have evolved surface conditions that could hardly be more different. Venus is extremely hot, dry, and in other ways so forbidding that it is improbable that life as it is understood on Earth could have developed there. One of scientists' major goals in studying Venus is to understand how its harsh conditions came about, which may hold important lessons about the causes of environmental change on Earth.

Basic astronomical data
      Viewed through a telescope, Venus presents a brilliant yellow-white, essentially featureless face to the observer. Its obscured appearance results from the surface of the planet being hidden from sight by a continuous and permanent cover of clouds (cloud). Features in the clouds are difficult to see in visible light. When observed at ultraviolet wavelengths, the clouds exhibit distinctive dark markings, with complex swirling patterns near the equator and global-scale bright and dark bands that are V-shaped and open toward the west. Because of the all-enveloping clouds, little was known about Venus's surface, atmosphere, and evolution before the early 1960s, when the first radar observations were undertaken and spacecraft made the first flybys of the planet.

      Venus orbits the Sun at a mean distance of 108 million km (67 million miles), which is about 0.7 times Earth's distance from the Sun. It has the least eccentric orbit of any planet, with a deviation from a perfect circle of only about 1 part in 150. Consequently, its distances at perihelion and aphelion (i.e., when it is nearest and farthest from the Sun, respectively) vary little from the mean distance. The period of its orbit—that is, the length of the Venusian year—is 224.7 Earth days. As Venus and Earth revolve around the Sun, the distance between them varies from a minimum of about 42 million km (26 million miles) to a maximum of about 257 million km (160 million miles).

      Because Venus's orbit lies within Earth's, the planet exhibits phases (phase) like those of the Moon when viewed from Earth. In fact, the discovery of these phases by the Italian scientist Galileo in 1610 was one of the most important in the history of astronomy. In Galileo's day the prevailing model of the universe was based on the assertion by the Greek astronomer Ptolemy almost 15 centuries earlier that all celestial objects revolve around Earth (geocentric system) (see Ptolemaic system). Observation of the phases of Venus was inconsistent with this view but was consistent with the Polish astronomer Nicolaus Copernicus (Copernicus, Nicolaus)'s idea that the solar system is centred on the Sun (heliocentric system). Galileo's observation of the phases of Venus provided the first direct observational evidence for Copernican theory (Copernican system).

      The rotation of Venus on its axis is unusual in both its direction and its speed. The Sun and most of the planets in the solar system rotate in a counterclockwise direction when viewed from above their north poles; this direction is called direct, or prograde. Venus, however, rotates in the opposite, or retrograde (retrograde motion), direction. Were it not for the planet's clouds, an observer on Venus's surface would see the Sun rise in the west and set in the east. Venus spins very slowly, taking about 243 Earth days to complete one rotation with respect to the stars—the length of its sidereal day. Venus's spin and orbital periods are very nearly synchronized with Earth's orbit such that, when the two planets are at their closest, Venus presents almost the same face toward Earth. The reasons for this are complex and have to do with the gravitational interactions of Venus, Earth, and the Sun, as well as the effects of Venus's massive rotating atmosphere. Because Venus's spin axis is tilted only about 3° toward the plane of its orbit, the planet does not have appreciable seasons. Astronomers as yet have no satisfactory explanation for Venus's peculiar rotational characteristics. The idea cited most often is that, when Venus was forming from the accretion of planetary building blocks (planetesimals), one of the largest of these bodies collided with the proto-Venus in such a way as to tip it over and possibly slow its spin as well.

       Planetary data for VenusVenus's mean radius is 6,051.8 km (3,760.4 miles), or about 95 percent of Earth's at the Equator, while its mass is 4.87 × 1024 kg, or 81.5 percent that of Earth. The similarities to Earth in size and mass produce a similarity in density—5.25 grams per cubic centimetre for Venus, compared with 5.52 for Earth. They also result in a comparable surface gravity—humans standing on Venus would possess nearly 90 percent of their weight on Earth. Venus is more nearly spherical than most planets. A planet's rotation generally causes a bulging at the equator and a slight flattening at the poles, but Venus's very slow spin allows it to maintain its highly spherical shape. For additional orbital and physical data, see the table (Planetary data for Venus).

 Venus has the most massive atmosphere of the terrestrial planets, which includes Mercury, Earth, and Mars. Its gaseous envelope is composed of more than 96 percent carbon dioxide and 3.5 percent molecular nitrogen. Trace amounts of other gases are present, including carbon monoxide, sulfur dioxide, water vapour, argon, and helium. The atmospheric pressure at the planet's surface varies with surface elevation; at the elevation of the planet's mean radius it is about 95 bars, or 95 times the atmospheric pressure at Earth's surface. This is the same pressure found at a depth of about 1 km (0.6 mile) in Earth's oceans.

      Venus's upper atmosphere extends from the fringes of space down to about 100 km (60 miles) above the surface. There the temperature varies considerably, reaching a maximum of about 300–310 kelvins (K; 80–98 °F, 27–37 °C) in the daytime and dropping to a minimum of 100–130 K (−280 to −226 °F, −173 to −143 °C) at night. In the middle atmosphere the temperature increases smoothly with decreasing altitude, from about 173 K (−148 °F, −100 °C) at 100 km above the surface to roughly 263 K (14 °F, −10 °C) at the top of the continuous cloud deck, which lies at an altitude of more than 60 km (37 miles). Below the cloud tops the temperature continues to increase sharply through the lower atmosphere, or troposphere, reaching 737 K (867 °F, 464 °C) at the surface at the planet's mean radius. This temperature is higher than the melting point of lead or zinc.

      The clouds that enshroud Venus are enormously thick. The main cloud deck rises from about 48 km (30 miles) in altitude to 68 km (42 miles). In addition, thin hazes exist above and below the main clouds, extending as low as 32 km (20 miles) and as high as 90 km (56 miles) above the surface. The upper haze is somewhat thicker near the poles than in other regions.

      The main cloud deck is formed of three layers. All of them are quite tenuous—an observer in even the densest cloud regions would be able to see objects at distances of several kilometres. The opacity of the clouds varies rapidly with space and time, which suggests a high level of meteorologic activity. Whether lightning occurs within the clouds—a possibility indicated by data from some spacecraft missions to Venus—is a subject of debate. The clouds are bright and yellowish when viewed from above, reflecting roughly 85 percent of the sunlight striking them. The material responsible for the yellowish colour has not been confidently identified.

      The microscopic particles that make up the Venusian clouds consist of liquid droplets and perhaps also solid crystals. The dominant material is highly concentrated sulfuric acid. Other materials that may exist there include solid sulfur, nitrosylsulfuric acid, and phosphoric acid. Cloud particles range in size from less than 0.5 micrometres (0.00002 inch) in the hazes to a few micrometres in the densest layers.

      The reasons that some cloud-top regions appear dark when viewed in ultraviolet light are not fully known. Materials that may be present in minute quantities above the cloud tops and that may be responsible for absorbing ultraviolet light in some regions include sulfur dioxide, solid sulfur, chlorine, and iron(III) chloride.

      The circulation of Venus's atmosphere is quite remarkable and is unique among the planets. Although the planet rotates only three times in two Earth years, the cloud features in the atmosphere circle Venus completely in about four days. The wind at the cloud tops blows from east to west at a velocity of about 100 metres per second (360 km [220 miles] per hour). This enormous velocity decreases markedly with decreasing height such that winds at the planet's surface are quite sluggish—typically no more than 1 metre per second (less than 4 km [2.5 miles] per hour). Much of the detailed nature of the westward flow above the cloud tops can be attributed to tidal (tide) motions induced by solar heating. Nevertheless, the fundamental cause of this “superrotation” of Venus's dense atmosphere is unknown, and it remains one of the more intriguing mysteries in planetary science.

 Most information about wind directions at the planet's surface comes from observations of wind-blown materials. Despite low surface-wind velocities, the great density of Venus's atmosphere enables these winds to move loose fine-grained materials, producing surface features that have been seen in radar images. Some features resemble sand dunes, while others are “wind streaks” produced by preferential deposition or erosion downwind from topographic features. The directions assumed by the wind-related features suggest that in both hemispheres the surface winds blow predominantly toward the equator. This pattern is consistent with the idea that simple hemispheric-scale circulation systems called Hadley cells (Hadley cell) exist in the Venusian atmosphere. According to this model, atmospheric gases rise upward as they are heated by solar energy at the planet's equator, flow at high altitude toward the poles, sink to the surface as they cool at higher latitudes, and flow toward the equator along the planet's surface until they warm and rise again. Some deviations from the equatorward flow pattern are observed on regional scales. They may be caused by the influence of topography on wind circulation.

      A major consequence of Venus's massive atmosphere is that it produces an enormous greenhouse effect, which intensely heats the planet's surface. Because of its bright continuous cloud cover, Venus actually absorbs less of the Sun's light than does Earth. Nevertheless, the sunlight that does penetrate the clouds is absorbed both in the lower atmosphere and at the surface. The surface and the gases of the lower atmosphere, which are heated by the absorbed light, reradiate this energy at infrared wavelengths. On Earth most reradiated infrared radiation escapes back into space, which allows Earth to maintain a reasonably cool surface temperature. On Venus, in contrast, the dense carbon dioxide atmosphere and the thick cloud layers trap much of the infrared radiation. The trapped radiation heats the lower atmosphere further, ultimately raising the surface temperature by hundreds of degrees. Study of the Venusian greenhouse effect has led to an improved understanding of the more subtle but very important influence of greenhouse gases in Earth's atmosphere and a greater appreciation of the effects of energy use and of other human activities on Earth's energy balance.

      Above the main body of the Venusian atmosphere lies the ionosphere (ionosphere and magnetosphere). As its name implies, the ionosphere is composed of ions (ion), or charged particles, produced both by absorption of ultraviolet solar radiation and by the impact of the solar wind—the flow of charged particles streaming outward from the Sun—on the upper atmosphere. The primary ions in the Venusian ionosphere are forms of oxygen (O+ and O2+) and carbon dioxide (CO2+).

Interaction with the solar wind
      Unlike most planets, including Earth, Venus does not exhibit an intrinsic magnetic field (see geomagnetic field). Sensitive measurements by orbiting spacecraft have shown that any dipole field originating from within Venus must be no more than 1/8,000 that of Earth's. The lack of a magnetic field may be related in part to the planet's slow rotation because, according to the dynamo theory that explains the origin of planetary magnetic fields, rotation helps to drive the fluid motions within the planet's interior that produce the field. It is also possible that Venus may lack a magnetic field because its core is fluid but does not circulate or simply because the core is solid and hence is incapable of supporting a dynamo.

      As the solar wind bombards a planet at supersonic speeds, it generally forms a bow shock (bow wave) on the planet's sunward side—that is, a standing wave of plasma that slows down, heats, and deflects the flow around the planet. For some planets the bow shock lies at a considerable distance from the surface, held off by the planet's magnetic field. For example, because of Jupiter's enormous magnetic field, the bow shock exists about 3,000,000 km (1,900,000 miles) from the planet; for Earth the distance is about 65,000 km (40,000 miles). Because Venus lacks a detectable field, however, its bow shock lies just a few thousand kilometres above the surface, held off only by the planet's ionosphere. This closeness of the bow shock to the surface leads to particularly intense interactions between the solar wind and Venus's atmosphere. In fact, the top of the ionosphere, known as the ionopause, lies at a much lower altitude on the dayside of Venus than on the nightside owing to the pressure exerted by the solar wind. The density of the ionosphere is also far greater on the dayside of the planet than on the nightside.

      Venus's interaction with the solar wind results in a gradual, continuous loss to space of hydrogen and oxygen from the planet's upper ionosphere. This process is equivalent to a gradual loss of water from the planet. Over the course of Venus's history, the total amount of water lost via this mechanism could have been as much as a few percent of a world ocean the size of Earth's.

Character of the surface
 The high atmospheric pressure, the low wind velocities, and, in particular, the extremely high temperatures create a surface environment on Venus that is markedly different from any other in the solar system. A series of landings by robotic Soviet (Venera) spacecraft in the 1970s and early '80s provided detailed data on surface composition and appearance. Views of the Venusian landscape, typified in colour images obtained by the Soviet Union's Venera 13 lander in 1982 (see figure—>), show rocky plains that stretch toward the horizon. Despite the heavy cloud cover, the surface is well illuminated by the yellow-orange light that filters through the clouds.

      The most striking characteristic of the surface at the Venera 13 site and most other Venera landing sites is the flat, slabby, layered nature of the rocks (rock). Both volcanic and sedimentary rocks on Earth can develop such an appearance under appropriate conditions, but the reason that the Venusian rocks have done so is not known with certainty. Also present among the rocks is a darker, fine-grained soil. The grain size of the soil is unknown, but some of it was fine enough to be lifted briefly into the atmosphere by the touchdown of the Venera lander, which suggests that some grains are no more than a few tens of micrometres in diameter. Scattered throughout the soil and atop the rocks are pebble-size particles that could be either small rocks or clods of soil.

      The general surface appearance at the Venera landing sites is probably common on Venus, but it is likely not representative of all locations on the planet. Radar data from the U.S. Magellan spacecraft, which studied Venus from orbit in the early 1990s, provided global information about the roughness of the Venusian surface at scales of metres to tens of metres. Although much of the planet is indeed covered by lowland plains that appear smooth to radar, some terrains were found to be very much rougher. These include areas covered by ejecta (the material expelled from impact craters and extending around them), steep slopes associated with tectonic activity, and some lava flows. How such terrains would appear from a lander's perspective is not known, but large boulders and other sorts of angular blocks presumably would be more common than at the Venera sites.

Surface composition
      A number of the Soviet landers carried instruments to analyze the chemical composition of the surface materials of Venus. Because only the relative proportions of a few elements were measured, no definitive information exists concerning the rock types or minerals present. Two techniques were used to measure the abundances of various elements. Gamma-ray spectrometers, which were carried on Veneras (Venera) 8, 9, and 10 and the landers of the Soviet Vega 1 and 2 missions, measured the concentrations of naturally radioactive isotopes of the elements uranium, potassium, and thorium. X-ray fluorescence instruments, carried on Veneras 13 and 14 and Vega 2, measured the concentrations of a number of major elements.

      The Venera 8 site gave indications that the rock composition may be similar to that of granite or other igneous rocks that compose Earth's continents. This inference, however, was based only on rather uncertain measurements of the concentrations of a few radioactive elements. Measurements of radioactive elements at the Venera 9 and 10 and Vega 1 and 2 landing sites suggested that the compositions there resemble those of basalt rocks found on Earth's ocean floors and in some volcanic regions such as Hawaii and Iceland. The Venera 13 and 14 and Vega 2 X-ray instruments measured concentrations of silicon, aluminum, magnesium, iron, calcium, potassium, titanium, manganese, and sulfur. Although some differences in composition were seen among the three sites, on the whole the elemental compositions measured by all three landers were similar to those of basalts on Earth.

      A surprising result of orbital radar observations of Venus is that the highest elevations on the planet exhibit anomalously high radar reflectivity. The best interpretation seems to be that the highest elevations are coated with a thin layer of some semiconducting material. Its composition is unknown, but it could be an iron-containing mineral such as pyrite or magnetite, which formed at cooler, higher elevations from low concentrations of atmospheric iron(II) chloride vapour in the atmosphere.

Surface features
 Earth-based observatories and Venus-orbiting spacecraft have provided global-scale information on the nature of the planet's surface. All have used radar systems to penetrate the thick Venusian clouds.

      The entire surface of the planet is dry and rocky. Because there is no sea level in the literal sense, elevation is commonly expressed as a planetary radius—i.e., as the distance from the centre of the planet to the surface at a given location. Another method, in which elevation is expressed as the distance above or below the planet's mean radius, is also used. Most of the planet consists of gently rolling plains. In some areas the elevations change by only a few hundred metres over distances of hundreds of kilometres. Globally, more than 80 percent of the surface deviates less than 1 km (0.6 mile) from the mean radius. At several locations on the plains are broad, gently sloping topographic depressions, or lowlands, that may reach several thousand kilometres across; they include Atalanta Planitia, Guinevere Planitia, and Lavinia Planitia.

 Two striking features are the continent-sized highland areas, or terrae— Ishtar Terra in the northern hemisphere and Aphrodite Terra along the equator. Ishtar is roughly the size of Australia, while Aphrodite is comparable in area to South America. Ishtar possesses the most spectacular topography on Venus. Much of its interior is a high plateau, called Lakshmi Planum, that resembles in configuration the Plateau of Tibet on Earth. Lakshmi is bounded by mountains on most sides, the largest range being the enormous Maxwell Montes on the east. These mountains soar about 11 km (7 miles) above the mean radius of Venus. The topography of Aphrodite, more complex than that of Ishtar, is characterized by a number of distinct mountain ranges and several deep, narrow troughs. In addition to the two main terrae are several smaller elevated regions, including Alpha Regio, Beta Regio, and Phoebe Regio.

      Many of the surface features on Venus can be attributed to tectonic activity—that is, to deformational motions within the crust. These include mountain belts, plains deformation belts, rifts, coronae, and tesserae, which are discussed in turn below (see also tectonic landform).

Mountain belts
 Found in the terrae, Venus's mountain belts are in some ways similar to ones on Earth such as the Himalayas of Asia and the Andes of South America. Among the best examples are those that encircle Lakshmi Planum, which in addition to Maxwell Montes include Freyja, Akna, and Danu Montes. Maxwell Montes is particularly broad and comparable in size to the Himalayas.

      Venus's mountain belts typically consist of parallel ridges and troughs with spacings of 5–10 km (3–6 miles). They probably developed when broad bands of the lithosphere were compressed from the sides and became thickened, folding and thrusting surface materials upward. Their formation in some respects thus resembles the building of many mountain ranges on Earth. On the other hand, because of the lack of liquid water or ice on Venus, their appearance differs in major ways from their counterparts on Earth. Without the flow of rivers or glaciers to wear them down, Venusian mountain belts have acquired steep slopes as a result of folding (fold) and faulting (fault). In some places the slopes have become so steep that they have collapsed under their own weight. The erosional forms common in mountainous regions on Earth are absent.

Plains deformation belts
      Although plains deformation belts are similar in some ways to mountain belts, they display less pronounced relief and are found primarily in low-lying areas of the planet, such as Lavinia Planitia and Atalanta Planitia. Like mountain belts, they show strong evidence for parallel folding and faulting and may form primarily by compression, deformation, and uplift of the lithosphere. Within a given lowland, it is common for deformation belts to lie roughly parallel to one another, spaced typically several hundred kilometres apart.

Rifts (rift valley)
 Rifts (see rift valley) are among the most spectacular tectonic features on Venus. The best-developed rifts are found atop broad, raised areas such as Beta Regio, sometimes radiating outward from their centres like the spokes of a giant wheel. Beta and several other similar regions on Venus appear to be places where large areas of the lithosphere have been forced upward from below, splitting the surface to form great rift valleys. The rifts are composed of innumerable faults, and their floors typically lie 1–2 km (0.6–1.2 miles) below the surrounding terrain. In many ways the rifts on Venus are similar to great rifts elsewhere, such as the East African Rift (East African Rift System) on Earth or Valles Marineris on Mars; volcanic eruptions, for example, appear to have been associated with all these features. The Venusian rifts differ from Earth and Martian ones, however, in that little erosion has taken place within them owing to the lack of water.

 Coronae (Latin: “garlands” or “crowns”) are landforms that apparently owe their origin to the effects of hot, buoyant blobs of material, known from terrestrial geology as diapirs (diapir), that originate deep beneath the surface of Venus. Coronae evolve through several stages. As diapirs first rise through the planet's interior and approach the surface, they can lift the rocks above them, fracturing the surface in a radial pattern. This results in a distinctive starburst of faults and fractures, often lying atop a broad, gently sloping topographic rise. (Such features are sometimes called novae, a name given to them when their evolutionary relationship to coronae was less certain.) Once a diapir has neared the surface and cooled, it loses its buoyancy. The initially raised crust then can sag under its own weight, developing concentric faults as it does so. The result is a circular-to-oval pattern of faults, fractures, and ridges. Volcanism can occur through all stages of corona formation. During the late stages it tends to obscure the radial faulting that is characteristic of the early stages.

 Coronae are typically a few hundred kilometres in diameter. Although they may have a raised outer rim, many coronae sag noticeably in their interiors and also outside their rims. Hundreds of coronae are found on Venus, observed at all stages of development. The radially fractured domes of the early stages are comparatively uncommon, while the concentric scars characteristic of mature coronae are among the most numerous large tectonic features on the planet.

Tesserae
 Tesserae (Latin: “mosaic tiles”) are the most geologically complex regions seen on Venus. Several large elevated regions, such as Alpha Regio, are composed largely of tessera terrain. Such terrain appears extraordinarily rugged and highly deformed in radar images, and in some instances it displays several different trends of parallel ridges and troughs that cut across one another at a wide range of angles. The deformation in tessera terrain can be so complex that sometimes it is difficult to determine what kinds of stresses in the lithosphere were responsible for forming it. In fact, probably no single process can explain all tessera formation. Tesserae typically appear very bright in radar images, which suggests an extremely rough and blocky surface at scales of metres. Some tesserae may be old terrain that has been subjected to more episodes of mountain building and faulting than have the materials around it, each one superimposed on its predecessor to produce the complex pattern observed.

Volcanic features
 Along with intense tectonic activity, Venus has undergone much volcanism. The largest volcanic outpourings are the huge lava fields that cover most of the rolling plains. These are similar in many respects to fields of overlapping lava flows seen on other planets, including Earth, but they are far more extensive. Individual flows are for the most part long and thin, which indicates that the erupting lavas were very fluid and hence were able to flow long distances over gentle slopes. Lavas on Earth and the Moon that flow this readily typically consist of basalts (basalt), and so it is probable that basalts are common on the plains of Venus as well.

      Of the many types of lava-flow features seen on the Venusian plains, none are more remarkable than the long, sinuous canali. These meandering channels usually have remarkably constant widths, which can be as much as 3 km (2 miles). They commonly extend as far as 500 km (300 miles) across the surface; one is 6,800 km (4,200 miles) long. Canali probably were carved by very low-viscosity lavas that erupted at sustained high rates of discharge. In a few instances segments of canali appear to proceed uphill, which suggests that crustal deformation took place after the channels were carved and reversed the gentle downward surface slopes to upward ones. Other channel-like volcanic features on Venus include sinuous rilles that may be collapsed lava tubes (see lava), and large, complex compound valleys that apparently result from particularly massive outpourings of lava.

  In many locations on Venus, volcanic eruptions have built edifices similar to the great volcanoes of Hawaii on Earth or those associated with the Tharsis region on Mars. Sif Mons is an example of such a volcano; there are more than 100 others distributed widely over the planet. Known as shield volcanoes, they reach heights of several kilometres above the surrounding plains and can be hundreds of kilometres across at their base. They are made up of many individual lava flows piled on one another in a radial pattern. They develop when a source of lava below the surface remains fixed and active at one location long enough to allow the volcanic materials it extrudes to accumulate above it in large quantities. Like those found on the rolling plains, the flows constituting the shield volcanoes are generally very long and thin and are probably composed of basalt.

 When a subsurface source of lava is drained of its contents, the ground above it may collapse, forming a depression called a caldera. Many volcanic calderas are observed on Venus, both atop shield volcanoes and on the widespread lava plains. They are often roughly circular in shape and overall are similar to calderas observed on Earth and Mars. The summit region of Sif Mons, for example, exhibits a caldera-like feature 40–50 km (25–30 miles) in diameter.

  Along with the extensive lava plains and the massive shield volcanoes are many smaller volcanic landforms. Enormous numbers of small volcanic cones are distributed throughout the plains. Particularly unusual in appearance are so-called pancake domes, which are typically a few tens of kilometres in diameter and about 1 km (0.6 mile) high and are remarkably circular in shape. Flat-topped and steep-sided, they appear to have formed when a mass of thick lava was extruded from a central vent and spread outward for a short distance in all directions before solidifying. The lavas that formed such domes clearly were much more viscous than most lavas on Venus. Their composition is unknown, but—given the knowledge of lavas on Earth—they are likely to be much richer in silica (see silica mineral) than the basalts thought to predominate elsewhere on the planet.

      Volcanic edifices are not uniformly distributed on Venus. Although they are common everywhere, they are particularly concentrated in the Beta-Atla-Themis region, between longitudes 180° and 300° E. This concentration may be the consequence of a broad active upwelling of the Venusian mantle in this area, which has led to enhanced heat flow and formation of magma reservoirs.

Impact craters (meteorite crater)
  The Venusian surface has been altered by objects from outside the planet as well as by forces from within. Impact craters dot the landscape, created by meteorites that passed through the atmosphere and struck the surface. Nearly all solid bodies in the solar system bear the scars of meteoritic impacts, with small craters typically being more common than large ones. This general tendency is encountered on Venus as well—craters a few hundred kilometres across are present but rare, while craters tens of kilometres in diameter and smaller are common. Venus has an interesting limitation, however, in that craters smaller than about 1.5–2 km (1–1.2 miles) in diameter are not found. Their absence is attributable to the planet's dense atmosphere, which causes intense frictional heating and strong aerodynamic forces as meteorites plunge through it at high velocities. The larger meteorites reach the surface intact, but the smaller ones are slowed and fragmented in the atmosphere. In fact, craters several kilometres in size—i.e., near the minimum size observed—tend not to be circular. Instead they have complex shapes, often with several irregular pits rather than a single central depression, which suggests that the impacting body broke up into a number of fragments that struck the surface individually. Radar images also show diffuse dark and bright “smudges” that may be have resulted from the explosions of small meteorites above the surface.

      The large craters that are seen on Venus are different in a number of respects from those observed on other planets. Most impact craters, on Venus and elsewhere, show ejecta around them. Venusian ejecta is unusual, however, in that its outer border commonly shows a lobed or flower-petal pattern, which suggests that much of it poured outward in a ground-hugging flow rather than arcing high above the ground ballistically and falling back to the surface. This behaviour was probably produced by dense atmospheric gases that became entrained in the flow and resulted in a turbulent cloud of gas and ejecta. Another peculiarity of large Venusian craters is the sinuous flows that have emerged from the ejecta, spreading outward from it just as lava flows would. These flows are apparently composed of rock that was melted by the high pressures and temperatures reached during the impact. The prevalence of these flow features on Venus must be due in large part to the planet's high surface temperature—rocks are closer to their melting temperature when craters form, which allows more melt to be produced than on other planets. For the same reason, the molten rock will remain fluid longer, which allows it to flow for significant distances.

 Perhaps the strangest property of Venusian craters is one associated with some of the youngest. In addition to the normal ejecta, these craters are partially surrounded by huge parabola-shaped regions of dark material, a feature not found elsewhere in the solar system. In every case, the parabola opens to the west, and the crater is nestled within it, toward its eastern extremity. In radar images the dark materials tend to be smooth at small scales; it is likely that these parabolas are composed of deposits of fine-grained ejecta that was thrown upward during the impact event. Apparently the material rose above the Venusian atmosphere, fell back, and was picked up by the high-speed westward-blowing winds that encircle the planet. It was then carried far downwind from the impact site, eventually descending to the surface to form a parabola-shaped pattern.

      For planets and moons that have impact craters, crater populations are an important source of information about the ages of the surfaces on which they lie. The concept is simple in principle—on a given body older surfaces have more craters than do younger ones. Determining an absolute age in years is difficult, however, and requires knowledge about the rate of crater formation, which usually must be inferred indirectly. The absolute ages of materials on the surface of Venus are not known, but the overall density of craters on Venus is lower than on many other bodies in the solar system. Estimates vary, but the average age of materials on Venus is almost certainly less than one billion years and may in fact be substantially less.

      The spatial distribution of craters on Venus is essentially random. If craters were clustered in distinct regions, scientists could infer that a wide range of surface ages was represented over the planet. With a near-random global crater distribution, however, they are led instead to the conclusion that essentially the entire planet has been geologically resurfaced in the last billion years or less and that much of the resurfacing took place in a comparatively brief time.

Interior structure and geologic evolution
      Much less is known about the interior of Venus than about its surface and atmosphere. Nevertheless, because the planet is much like Earth in overall size and density and because it presumably accreted from similar materials (see solar system: Origin of the solar system (solar system)), scientists expect that it evolved at least a crudely similar internal state. Therefore, it probably has a core of metal, a mantle of dense rock, and a crust of less-dense rock. The core, like that of Earth, is probably composed primarily of iron and nickel, although Venus's somewhat lower density may indicate that its core also contains some other, less-dense material such as sulfur. Because no intrinsic magnetic field has been detected for Venus, there is no direct evidence for a metallic core, as there is for Earth. Calculations of Venus's internal structure suggest that the outer boundary of the core lies a little more than 3,000 km (1,860 miles) from the centre of the planet.

      Above the core and below the crust lies Venus's mantle, making up the bulk of the planet's volume. Despite the high surface temperatures, temperatures within the mantle are likely similar to those in Earth's mantle. Even though a planetary mantle is composed of solid rock, the material there can slowly creep or flow, just as glacial ice does, allowing sweeping convective motions to take place. convection is a great equalizer of the temperatures of planetary interiors. Similar to heat production within Earth, heat within Venus is thought to be generated by the decay of natural radioactive materials. This heat is transported to the surface by convection. If temperatures deep within Venus were substantially higher than those within Earth, the viscosity of the rocks in the mantle would drop sharply, speeding convection and removing the heat more rapidly. Therefore, the deep interiors of Venus and Earth are not expected to differ dramatically in temperature.

      As noted above, the composition of the Venusian crust is believed to be dominated by basalt. Gravity data suggest that the thickness of the crust is fairly uniform over much of the planet, with typical values of perhaps 20–50 km (12–30 miles). Possible exceptions are the tessera highlands, where the crust may be significantly thicker.

      Convective motions in a planet's mantle can cause materials near the surface to experience stress, and motions in the Venusian mantle may be largely responsible for the tectonic deformation observed in radar images. On Venus the gravity field is found to correlate more strongly with topography over broad regional scales than it is on Earth—i.e., large regions where the topography is higher than the mean elevation on Venus also tend to be regions where the measured gravity is higher than average. This implies that much of the increased mass associated with the elevated topography is not offset by a compensating deficit of mass in the underlying crust that supports it (so-called low-density roots), as it is on Earth (see isostasy). Instead, some of the broad-scale relief on Venus may owe its origin directly to present-day convective motions in the mantle. Raised topography, such as Beta Regio, could lie above regions of mantle upwelling, whereas lowered topography, such as Lavinia Planitia, could lie above regions of mantle downwelling.

      Despite the many overall similarities between Venus and Earth, the geologic evolution of the two planets has been strikingly different. Evidence suggests that the process of plate tectonics does not now operate on Venus. Although deformation of the lithosphere does indeed seem to be driven by mantle motions, lithospheric plates do not move mainly horizontally relative to each other, as they do on Earth. Instead, motions are mostly vertical, with the lithosphere warping up and down in response to the underlying convective motions. Volcanism, coronae, and rifts tend to be concentrated in regions of upwelling, while plains deformation belts are concentrated in regions of downwelling. The formation of rugged uplands such as Aphrodite (Aphrodite Terra) and Ishtar (Ishtar Terra) is not as well understood, but the mechanism probably involves some kind of local crustal thickening in response to mantle motions.

      The lack of plate tectonics on Venus may be due in part to the planet's high surface temperature, which makes the upper rigid layer of the planet—the lithosphere—more buoyant and hence more resistant to subduction than Earth's lithosphere, other factors being equal. Interestingly, there is evidence that the Venusian lithosphere may be thicker than Earth's and that it has thickened with time. A gradual, long-term thickening of Venus's lithosphere in fact could be related to the curious conclusion drawn from Venus's cratering record (see above the section Impact craters (Venus))—that most of the planet underwent a brief but intense period of geologic resurfacing less than a billion years ago. One possible explanation is that Venus may experience episodic global overturns of its mantle, in which an initially thin lithosphere slowly thickens until it founders on a near-global scale, triggering a brief, massive geologic resurfacing event. How many times this may have occurred during the planet's history and when it may happen again are unknown.

Observations from Earth
      Since Galileo's discovery of Venus's phases, the planet has been studied in detail, using Earth-based telescopes (telescope), radar, and other instruments. Over the centuries telescopic observers, including Gian Domenico Cassini (Cassini, Gian Domenico) of France and William Herschel (Herschel, Sir William) of England, have reported a variety of faint markings on its disk. Some of these markings may have corresponded to the cloud features observed in modern times in ultraviolet light, while others may have been illusory.

 Important early telescopic observations of Venus were conducted in the 1700s during the planet's solar transits (transit) (see eclipse: Transits of Mercury and Venus (eclipse)). In a solar transit an object passes directly between the Sun and Earth and is silhouetted briefly against the Sun's disk. Transits of Venus are rare events, occurring in pairs eight years apart with more than a century between pairs. They were extremely important events to 18th-century astronomy, since they provided the most accurate method known at that time for determining the distance between Earth and the Sun. (This distance, known as the astronomical unit, is one of the fundamental units of astronomy.) Observations of the 1761 transit were only partially successful but did result in the first suggestion, by the Russian scientist Mikhail V. Lomonosov (Lomonosov, Mikhail Vasilyevich), that Venus has an atmosphere. The second transit of the pair, in 1769, was observed with somewhat greater success. Transits must be viewed from many points on Earth to yield accurate distances, and the transits of 1761 and, particularly, 1769 prompted the launching of many scientific expeditions to remote parts of the globe. Among these was the first of the three voyages of exploration by the British naval officer James Cook (Cook, James), who, with scientists from the Royal Society, observed the 1769 transit from Tahiti. The transit observations of the 1700s not only gave an improved value for the astronomical unit but also provided the impetus for many unrelated but important discoveries concerning Earth's geography. By the time the subsequent pair of transits occurred, in 1874 and 1882, the nascent field of celestial photography had advanced enough to allow scientists to record on glass plates what they saw through their telescopes. No transits took place in the 20th century; the first of the next pair was widely observed and imaged in 2004.

      In the modern era Venus has also been observed at wavelengths outside the visible spectrum (spectroscopy). The cloud features were discovered with certainty in 1927–28 in ultraviolet photographs. The first studies of the infrared spectrum of Venus, in 1932, showed that its atmosphere is composed primarily of carbon dioxide. Subsequent infrared observations revealed further details about the composition of both the atmosphere and the clouds. Observations in the microwave portion of the spectrum, beginning in earnest in the late 1950s and early '60s, provided the first evidence of the extremely high surface temperatures on the planet and prompted the study of the greenhouse effect as a means of producing these temperatures.

      After finding that Venus is completely enshrouded by clouds, astronomers turned to other techniques to study its surface. Foremost among these has been radar (see radio and radar astronomy). If equipped with an appropriate transmitter, a large radio telescope can be used as a radar system to bounce a radio signal off a planet and detect its return. Because radio wavelengths penetrate the thick Venusian atmosphere, the technique is an effective means of probing the planet's surface.

      Earth-based radar observations have been conducted primarily from Arecibo Observatory in the mountains of Puerto Rico, the Goldstone tracking station complex in the desert of southern California, and Haystack Observatory in Massachusetts. The first successful radar observations of Venus took place at Goldstone and Haystack in 1961 and revealed the planet's slow rotation. Subsequent observations determined the rotation properties more precisely and began to unveil some of the major features on the planet's surface. The first features to be observed were dubbed Alpha, Beta, and Maxwell (Maxwell Montes), the last after James Clerk Maxwell (Maxwell, James Clerk), the Scottish physicist who first derived some of the basic equations that describe the propagation of electromagnetic radiation. These three features are among the brightest on the planet in radar images, and their names have been preserved to the present as Alpha Regio, Beta Regio, and Maxwell Montes.

      By the mid-1980s Earth-based radar technology had advanced such that images from Arecibo were revealing surface features as small as a few kilometres in size. Nevertheless, because Venus always presents nearly the same face toward Earth when the planets are at their closest, much of the surface went virtually unobserved from Earth.

Spacecraft exploration
      The greatest advances in the study of Venus were achieved through the use of robotic spacecraft. The first spacecraft to reach the vicinity of another planet and return data was the U.S. Mariner 2 in its flyby of Venus in 1962. Since then, Venus has been the target of more than 20 spacecraft missions.

 Successful early Venus missions undertaken by the United States involved Mariner 2, Mariner 5 (1967), and Mariner 10 (1974). Each spacecraft made a single close flyby, providing successively improved scientific data in accord with concurrent advances in spacecraft and instrument technology. After visiting Venus, Mariner 10 went on to a successful series of flybys of Mercury. In 1978 the United States launched the Pioneer Venus mission, comprising two complementary spacecraft. The Orbiter went into orbit around the planet, while the Multiprobe released four entry probes—one large probe and three smaller ones—that were targeted to widely separated points in the Venusian atmosphere to collect data on atmospheric structure and composition. The Orbiter carried 17 scientific instruments, most of them focused on study of the planet's atmosphere, ionosphere, and interaction with the solar wind. Its radar altimeter provided the first high-quality map of Venus's surface topography. Pioneer Venus Orbiter was one of the longest-lived planetary spacecraft, returning data for more than 14 years.

 Venus was also a major target of the Soviet Union's planetary exploration program during the 1960s, '70s, and '80s, which achieved several spectacular successes. After an early sequence of failed missions, in 1967 Soviet scientists launched Venera 4, comprising a flyby spacecraft as well as a probe that entered the planet's atmosphere. Highlights of subsequent missions included the first successful soft landing on another planet (Venera 7 in 1970), the first images returned from the surface of another planet (Venera 9 and 10 landers in 1975), and the first spacecraft placed in orbit around Venus (Venera 9 and 10 orbiters).

      In terms of the advances they provided in the global understanding of Venus, the most important Soviet missions were Veneras 15 and 16 in 1983. The twin orbiters carried the first radar systems flown to another planet that were capable of producing high-quality images of the surface. They produced a map of the northern quarter of Venus with a resolution of 1–2 km (0.6–1.2 miles), and many types of geologic features now known to exist on the planet were either discovered or first observed in detail in the Venera 15 and 16 data. Late the following year the Soviet Union launched two more spacecraft to Venus, Vegas 1 and 2. These delivered Venera-style landers and dropped off two balloons in the Venusian atmosphere, each of which survived for about two days and transmitted data from their float altitudes in the middle cloud layer. The Vega spacecraft themselves continued past Venus to conduct successful flybys of Halley's Comet in 1986.

 In 1990, on its way to Jupiter, the U.S. Galileo spacecraft flew by Venus. Among its more notable observations were images at near-infrared wavelengths that viewed deep into the atmosphere and showed the highly variable opacity of the main cloud deck.

  The most ambitious mission yet to Venus, the U.S. Magellan spacecraft, was launched in 1989 and the next year entered orbit around the planet, where it conducted observations until late 1994. Magellan carried a radar system capable of producing images with a resolution better than 100 metres (330 feet). Because the orbit was nearly polar, the spacecraft was able to view essentially all latitudes on the planet. On each orbit the radar system obtained an image strip about 20 km (12 miles) wide and typically more than 16,000 km (almost 10,000 miles) long, extending nearly from pole to pole. The image strips were assembled into mosaics, and high-quality radar images of about 98 percent of the planet were ultimately produced. Magellan also carried a radar altimeter system that measured the planet's surface topography as well as some properties of its surface materials. After the main radar objectives of the mission were completed, the spacecraft's orbit was modified slightly so that it passed repeatedly through the upper fringes of the Venusian atmosphere. The resulting drag on the spacecraft gradually removed energy from its orbit, turning an initially elliptical orbit into a low, circular one. This procedure, known as aerobraking, has since been used on other planetary missions to conserve large amounts of fuel by reducing the use of thrusters for orbital reshaping. From its new circular orbit, the Magellan spacecraft was able to make the first detailed map of Venus's gravitational field.

      The U.S. Cassini (Cassini-Huygens)-Huygens spacecraft flew by Venus twice, in 1998 and 1999, on the way to its primary target, Saturn. During its brief passages near Venus, Cassini failed to corroborate signs of the existence of lightning in the planet's atmosphere that had been observed by previous spacecraft. This suggested to some scientists that lightning on Venus is either rare or different from the lightning that occurs on Earth.

      The European Space Agency's Venus Express, which was launched in 2005, entered into orbit around Venus the following year, becoming the first European spacecraft to visit the planet. Venus Express carried a camera, a visible-light and infrared imaging spectrometer, and other instruments to study Venus's magnetic field, plasma environment, atmosphere, and surface for a planned mission of more than two Venusian years. Among its early accomplishments was the return of the first images of cloud structures over the planet's south pole.

Steven W. Squyres

Additional Reading
Ladislav E. Roth and Stephen D. Wall (eds.), The Face of Venus: The Magellan Radar-Mapping Mission (1995), is a good post-Magellan popular-level book on the Venusian surface, with many excellent illustrations. Still informative pre-Magellan popular-level treatments include Garry E. Hunt and Patrick Moore, The Planet Venus (1982); and Eric Burgess, Venus, an Errant Twin (1985).The scientific understanding of Venus is definitively and comprehensively summarized in S.W. Bougher, D.M. Hunten, and R.J. Phillips (eds.), Venus II (1997), a collection of papers written after the Magellan and Galileo missions. Mikhail Ya. Marov and David H. Grinspoon, The Planet Venus (1998), provides an excellent post-Magellan treatment of Venusian geology and most other aspects of the planet. A pre-Magellan overview of Venus's surface is given by Alexander T. Basilevsky and James W. Head III, “The Geology of Venus,” Annual Review of Earth and Planetary Sciences, 16:295–317 (1988). An interesting collection of papers by Soviet scientists in English on Venus is V.L. Barsukov et al. (eds.), Venus Geology, Geochemistry, and Geophysics (1992). The results of major spacecraft missions to Venus are reported in several journals: on Pioneer Venus, Journal of Geophysical Research, 85:7573–8337 (1980); V.L. Barsukov et al., “The Geology and Geomorphology of the Venus Surface as Revealed by the Radar Images Obtained by Veneras 15 and 16,” Journal of Geophysical Research, pt. B, Solid Earth and Planets, 91(B4):D378–D398 (March 30, 1986); Science, 253:1457–1612 (Sept. 27, 1991), an issue devoted to the Galileo flyby of Venus; and two issues of Journal of Geophysical Research, pt. E, Planets, vol. 97, devoted to detailed Magellan mission results: no. 8 (Aug. 25, 1992); and no. 10 (Oct. 25, 1992).Steven W. Squyres

* * *


Universalium. 2010.

Игры ⚽ Поможем написать реферат
Synonyms:

Look at other dictionaries:

  • VÉNUS — est la deuxième planète du système solaire que l’on rencontre en s’éloignant du Soleil. Elle gravite sur une orbite quasi circulaire dont le rayon moyen est de 108 millions de kilomètres. De ce fait, Vénus reçoit un rayonnement solaire presque… …   Encyclopédie Universelle

  • Venus — Ve nus, n. [L. Venus, eris, the goddess of love, the planet Venus.] 1. (Class. Myth.) The goddess of beauty and love, that is, beauty or love deified. [1913 Webster] 2. (Anat.) One of the planets, the second in order from the sun, its orbit lying …   The Collaborative International Dictionary of English

  • Venus — steht für: Venus (Mythologie), die römische Göttin der Liebe, des erotischen Verlangens und der Schönheit Venus (Planet), der 2. Planet unseres Sonnensystems, benannt nach der Göttin Venus Venussymbol ♀, ein Symbol mit mehreren Bedeutungen,… …   Deutsch Wikipedia

  • Venus 5 — セーラー戦士 ヴィーナス・ファイブ (Sailor Senshi Venus Five) Жанр хентай, пародия, приключения, махо сёдзё, научная фантастика OVA …   Википедия

  • Venus —  Ne doit pas être confondu avec Vénus. Cette page d’homonymie répertorie les différents sujets et articles partageant un même nom. Venus peut désigner : Sommaire …   Wikipédia en Français

  • Venus — «Venus» Сингл Shocking Blue из альбома At Home Сторона «А» Venus Сторона «Б» Hot Sand Выпущен 1969 Записан 1969 Жанр …   Википедия

  • Venus GP — Daten Standort Space World (Japan) Typ Stahl – sitzend …   Deutsch Wikipedia

  • venus — VENUS. s. f. La Planette qui est entre Mercure & le Soleil. Venus directe, stationnaire & retrograde. la planette de Venus. Venus a son croissant & son decours aussi bien que la Lune. Venus, En termes de Chymie, signifie, Le cuivre. Venus, est… …   Dictionnaire de l'Académie française

  • Venus — (Карловы Вары,Чехия) Категория отеля: 4 звездочный отель Адрес: Sadová 815 /8, Карловы Вары, 360 …   Каталог отелей

  • venus — (De Venus, diosa mitológica de la hermosura). 1. f. Representación plástica de la diosa Venus. 2. Cada una de las estatuillas prehistóricas femeninas elaboradas en piedra, marfil o hueso. 3. Mujer muy hermosa. 4. Deleite sexual o acto carnal. 5.… …   Diccionario de la lengua española

  • Venus — (Ла Рош ан Арден,Бельгия) Категория отеля: Адрес: Chemin du Pafy, 6980 Ла Рош ан Арден, Бельг …   Каталог отелей

Share the article and excerpts

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