Mercury (planet)

From Wikipedia, the free encyclopedia
Jump to: navigation, search
Mercury
Mercury in color - Prockter07 centered.jpg
Mercury in color, taken by MESSENGER
Designations
Pronunciation Listeni/ˈmɜrkjəri/
Adjectives Mercurian,[1] Hermian[2]
Orbital characteristics[5]
Epoch J2000
Aphelion
  • 69,816,900 km
  • 0.466 697 AU
Perihelion
  • 46,001,200 km
  • 0.307 499 AU
  • 57,909,050 km
  • 0.387 098 AU
Eccentricity 0.205 630[3]
115.88 d[3]
47.362 km/s[3]
174.796°
Inclination
48.331°
29.124°
Satellites None
Physical characteristics
Mean radius
  • 2439.7±1.0 km[6][7]
  • 0.3829 Earths
Flattening 0[7]
  • 7.48×107 km2[6]
  • 0.147 Earths
Volume
  • 6.083×1010 km3[6]
  • 0.056 Earths
Mass
  • 3.3022×1023 kg[6]
  • 0.055 Earths
Mean density
5.427 g/cm3[6]
  • 3.7 m/s2
  • 0.38 g[6]
0.346±0.014
4.25 km/s[6]
  • 58.646 d
  • 1407.5 h[6]
Equatorial rotation velocity
10.892 km/h (3.026 m/s)
2.11′ ± 0.1′[8]
North pole right ascension
  • 18h 44m 2s
  • 281.01°[3]
North pole declination
61.45°[3]
Albedo
Surface temp. min mean max
0°N, 0°W [12] 100 K 340 K 700 K
85°N, 0°W[12] 80 K 200 K 380 K
−2.6[10] to 5.7[3][11]
4.5–13″[3]
Atmosphere[3]
Surface pressure
trace
Composition

Mercury is the smallest and closest to the Sun of the eight planets in the Solar System,[a] with an orbital period of about 88 Earth days. Seen from Earth, it appears to move around its orbit in about 116 days, which is much faster than any other planet. It has no known natural satellites.[b] The planet is named after the Roman deity Mercury, the messenger to the gods.

Because it has almost no atmosphere to retain heat, Mercury's surface experiences the greatest temperature variation of all the planets, ranging from 100 K (−173 °C; −280 °F) at night to 700 K (427 °C; 800 °F) during the day at some equatorial regions. The poles are constantly below 180 K (−93 °C; −136 °F). Mercury's axis has the smallest tilt of any of the Solar System's planets (about 130 of a degree), but it has the largest orbital eccentricity.[a] As such it does not experience seasons in the same way as most other planets such as Earth. At aphelion, Mercury is about 1.5 times as far from the Sun as it is at perihelion. Mercury's surface is heavily cratered and similar in appearance to the Moon, indicating that it has been geologically inactive for billions of years.

Mercury is gravitationally locked and rotates in a way that is unique in the Solar System. As seen relative to the fixed stars, it rotates exactly three times for every two revolutions[c] it makes around its orbit.[13] As seen from the Sun, in a frame of reference that rotates with the orbital motion, it appears to rotate only once every two Mercurian years. An observer on Mercury would therefore see only one day every two years.

Because Mercury moves in an orbit around the Sun that lies within Earth's orbit (as does Venus), it can appear in Earth's sky in the morning or the evening, but not in the middle of the night. Also, like Venus and the Moon, it displays a complete range of phases as it moves around its orbit relative to Earth. Although Mercury can appear as a bright object when viewed from Earth, its proximity to the Sun makes it more difficult to see than Venus. Two spacecraft have visited Mercury: Mariner 10 flew by in the 1970s and MESSENGER, launched in 2004, remains in orbit.

Internal structure

Internal structure of Mercury:
1. Crust: 100–300 km thick
2. Mantle: 600 km thick
3. Core: 1,800 km radius

Mercury is one of four terrestrial planets in the Solar System, and is a rocky body like Earth. It is the smallest planet in the Solar System, with an equatorial radius of 2,439.7 kilometres (1,516.0 mi).[3] Mercury is also smaller—albeit more massive—than the largest natural satellites in the Solar System, Ganymede and Titan. Mercury consists of approximately 70% metallic and 30% silicate material.[14] Mercury's density is the second highest in the Solar System at 5.427 g/cm3, only slightly less than Earth's density of 5.515 g/cm3.[3] If the effect of gravitational compression were to be factored out, the materials of which Mercury is made would be denser, with an uncompressed density of 5.3 g/cm3 versus Earth's 4.4 g/cm3.[15]

Mercury's density can be used to infer details of its inner structure. Although Earth's high density results appreciably from gravitational compression, particularly at the core, Mercury is much smaller and its inner regions are not as compressed. Therefore, for it to have such a high density, its core must be large and rich in iron.[16]

Geologists estimate that Mercury's core occupies about 42% of its volume; for Earth this proportion is 17%. Research published in 2007 suggests that Mercury has a molten core.[17][18] Surrounding the core is a 500–700 km mantle consisting of silicates.[19][20] Based on data from the Mariner 10 mission and Earth-based observation, Mercury's crust is believed to be 100–300 km thick.[21] One distinctive feature of Mercury's surface is the presence of numerous narrow ridges, extending up to several hundred kilometers in length. It is believed that these were formed as Mercury's core and mantle cooled and contracted at a time when the crust had already solidified.[22]

Mercury's core has a higher iron content than that of any other major planet in the Solar System, and several theories have been proposed to explain this. The most widely accepted theory is that Mercury originally had a metal-silicate ratio similar to common chondrite meteorites, thought to be typical of the Solar System's rocky matter, and a mass approximately 2.25 times its current mass.[23] Early in the Solar System's history, Mercury may have been struck by a planetesimal of approximately 1/6 that mass and several thousand kilometers across.[23] The impact would have stripped away much of the original crust and mantle, leaving the core behind as a relatively major component.[23] A similar process, known as the giant impact hypothesis, has been proposed to explain the formation of the Moon.[23]

Alternatively, Mercury may have formed from the solar nebula before the Sun's energy output had stabilized. It would initially have had twice its present mass, but as the protosun contracted, temperatures near Mercury could have been between 2,500 and 3,500 K and possibly even as high as 10,000 K.[24] Much of Mercury's surface rock could have been vaporized at such temperatures, forming an atmosphere of "rock vapor" that could have been carried away by the solar wind.[24]

A third hypothesis proposes that the solar nebula caused drag on the particles from which Mercury was accreting, which meant that lighter particles were lost from the accreting material and not gathered by Mercury.[25] Each hypothesis predicts a different surface composition, and two space missions, MESSENGER and BepiColombo, both will make observations to test them.[26][27] MESSENGER has found higher-than-expected potassium and sulfur levels on the surface, suggesting that the giant impact hypothesis and vaporization of the crust and mantle did not occur because potassium and sulfur would have been driven off by the extreme heat of these events. The findings would seem to favor the third hypothesis; however, further analysis of the data is needed.[28]

Surface geology

Main article: Geology of Mercury
Mercury's surface

Mercury's surface is similar in appearance to that of the Moon, showing extensive mare-like plains and heavy cratering, indicating that it has been geologically inactive for billions of years. Because our knowledge of Mercury's geology has been based only on the 1975 Mariner flyby and terrestrial observations, it is the least understood of the terrestrial planets.[18] As data from the recent MESSENGER flyby is processed, this knowledge will increase. For example, an unusual crater with radiating troughs has been discovered that scientists called "the spider".[29] It later received the name Apollodorus.[30]

Albedo features are areas of markedly different reflectivity, as seen by telescopic observation. Mercury possesses dorsa (also called "wrinkle-ridges"), Moon-like highlands, montes (mountains), planitiae (plains), rupes (escarpments), and valles (valleys).[31][32]

Names for features on Mercury come from a variety of sources. Names coming from people are limited to the deceased. Craters are named for artists, musicians, painters, and authors who have made outstanding or fundamental contributions to their field. Ridges, or dorsa, are named for scientists who have contributed to the study of Mercury. Depressions or fossae are named for works of architecture. Montes are named for the word "hot" in a variety of languages. Plains or planitiae are named for Mercury in various languages. Escarpments or rupēs are named for ships of scientific expeditions. Valleys or valles are named for radio telescope facilities.[33]

Mercury was heavily bombarded by comets and asteroids during and shortly following its formation 4.6 billion years ago, as well as during a possibly separate subsequent episode called the late heavy bombardment that came to an end 3.8 billion years ago.[34] During this period of intense crater formation, the planet received impacts over its entire surface,[32] facilitated by the lack of any atmosphere to slow impactors down.[35] During this time the planet was volcanically active; basins such as the Caloris Basin were filled by magma, producing smooth plains similar to the maria found on the Moon.[36][37]

Data from the October 2008 flyby of MESSENGER gave researchers a greater appreciation for the jumbled nature of Mercury's surface. Mercury's surface is more heterogeneous than either Mars's or the Moon's, both of which contain significant stretches of similar geology, such as maria and plateaus.[38]

Impact basins and craters

The so-called "Weird Terrain" was formed at the point antipodal to the Caloris Basin impact

Craters on Mercury range in diameter from small bowl-shaped cavities to multi-ringed impact basins hundreds of kilometers across. They appear in all states of degradation, from relatively fresh rayed craters to highly degraded crater remnants. Mercurian craters differ subtly from lunar craters in that the area blanketed by their ejecta is much smaller, a consequence of Mercury's stronger surface gravity.[39] According to IAU rules, each new crater must be named after an artist that was famous for more than fifty years, and dead for more than three years, before the date the crater is named.[40]

The largest known crater is Caloris Basin, with a diameter of 1,550 km.[41] The impact that created the Caloris Basin was so powerful that it caused lava eruptions and left a concentric ring over 2 km tall surrounding the impact crater. At the antipode of the Caloris Basin is a large region of unusual, hilly terrain known as the "Weird Terrain". One hypothesis for its origin is that shock waves generated during the Caloris impact traveled around the planet, converging at the basin's antipode (180 degrees away). The resulting high stresses fractured the surface.[42] Alternatively, it has been suggested that this terrain formed as a result of the convergence of ejecta at this basin's antipode.[43]

Overall, about 15 impact basins have been identified on the imaged part of Mercury. A notable basin is the 400 km wide, multi-ring Tolstoj Basin that has an ejecta blanket extending up to 500 km from its rim and a floor that has been filled by smooth plains materials. Beethoven Basin has a similar-sized ejecta blanket and a 625 km diameter rim.[39] Like the Moon, the surface of Mercury has likely incurred the effects of space weathering processes, including Solar wind and micrometeorite impacts.[44]

Plains

Degas crater

There are two geologically distinct plains regions on Mercury.[39][45] Gently rolling, hilly plains in the regions between craters are Mercury's oldest visible surfaces,[39] predating the heavily cratered terrain. These inter-crater plains appear to have obliterated many earlier craters, and show a general paucity of smaller craters below about 30 km in diameter.[45] It is not clear whether they are of volcanic or impact origin.[45] The inter-crater plains are distributed roughly uniformly over the entire surface of the planet.[citation needed]

Smooth plains are widespread flat areas that fill depressions of various sizes and bear a strong resemblance to the lunar maria. Notably, they fill a wide ring surrounding the Caloris Basin. Unlike lunar maria, the smooth plains of Mercury have the same albedo as the older inter-crater plains. Despite a lack of unequivocally volcanic characteristics, the localisation and rounded, lobate shape of these plains strongly support volcanic origins.[39] All the Mercurian smooth plains formed significantly later than the Caloris basin, as evidenced by appreciably smaller crater densities than on the Caloris ejecta blanket.[39] The floor of the Caloris Basin is filled by a geologically distinct flat plain, broken up by ridges and fractures in a roughly polygonal pattern. It is not clear whether they are volcanic lavas induced by the impact, or a large sheet of impact melt.[39]

One unusual feature of the planet's surface is the numerous compression folds, or rupes, that crisscross the plains. As the planet's interior cooled, it may have contracted and its surface began to deform, creating these features. The folds can be seen on top of other features, such as craters and smoother plains, indicating that the folds are more recent.[46] Mercury's surface is flexed by significant tidal bulges raised by the Sun—the Sun's tides on Mercury are about 17 times stronger than the Moon's on Earth.[47]

Surface conditions and "atmosphere" (exosphere)

Main article: Atmosphere of Mercury
Composite image of Mercury taken by MESSENGER
Radar image of Mercury's north pole
Composite of the north pole of Mercury, where NASA confirmed the discovery of a large volume of water ice, in permanently dark craters that exist there.[48]

The surface temperature of Mercury ranges from 100 K to 700 K[49] at the most extreme places: 0°N, 0°W, or 180°W. It never rises above 180 K at the poles,[12] due to the absence of an atmosphere and a steep temperature gradient between the equator and the poles. The subsolar point reaches about 700 K during perihelion (0°W or 180°W), but only 550 K at aphelion (90° or 270°W).[50] On the dark side of the planet, temperatures average 110 K.[12][51] The intensity of sunlight on Mercury's surface ranges between 4.59 and 10.61 times the solar constant (1,370 W·m−2).[52]

Although the daylight temperature at the surface of Mercury is generally extremely high, observations strongly suggest that ice (frozen water) exists on Mercury. The floors of deep craters at the poles are never exposed to direct sunlight, and temperatures there remain below 102 K; far lower than the global average.[53] Water ice strongly reflects radar, and observations by the 70 m Goldstone telescope and the VLA in the early 1990s revealed that there are patches of high radar reflection near the poles.[54] Although ice was not the only possible cause of these reflective regions, astronomers believed it was the most likely.[55]

The icy regions are believed to contain about 1014–1015 kg of ice,[56] and may be covered by a layer of regolith that inhibits sublimation.[57] By comparison, the Antarctic ice sheet on Earth has a mass of about 4×1018 kg, and Mars's south polar cap contains about 1016 kg of water.[56] The origin of the ice on Mercury is not yet known, but the two most likely sources are from outgassing of water from the planet's interior or deposition by impacts of comets.[56]

Mercury is too small and hot for its gravity to retain any significant atmosphere over long periods of time; it does have a "tenuous surface-bounded exosphere"[58] containing hydrogen, helium, oxygen, sodium, calcium, potassium and others. This exosphere is not stable—atoms are continuously lost and replenished from a variety of sources. Hydrogen and helium atoms probably come from the solar wind, diffusing into Mercury's magnetosphere before later escaping back into space. Radioactive decay of elements within Mercury's crust is another source of helium, as well as sodium and potassium. MESSENGER found high proportions of calcium, helium, hydroxide, magnesium, oxygen, potassium, silicon and sodium. Water vapor is present, released by a combination of processes such as: comets striking its surface, sputtering creating water out of hydrogen from the solar wind and oxygen from rock, and sublimation from reservoirs of water ice in the permanently shadowed polar craters. The detection of high amounts of water-related ions like O+, OH, and H2O+ was a surprise.[59][60] Because of the quantities of these ions that were detected in Mercury's space environment, scientists surmise that these molecules were blasted from the surface or exosphere by the solar wind.[61][62]

Sodium, potassium and calcium were discovered in the atmosphere during the 1980–1990s, and are believed to result primarily from the vaporization of surface rock struck by micrometeorite impacts.[63] In 2008 magnesium was discovered by MESSENGER probe.[64] Studies indicate that, at times, sodium emissions are localized at points that correspond to the planet's magnetic poles. This would indicate an interaction between the magnetosphere and the planet's surface.[65]

On November 29, 2012, NASA confirmed that images from MESSENGER had detected that craters at the north pole contained water ice. Sean C. Solomon was quoted in the New York Times as estimating the volume of the ice as large enough to "encase Washington, D.C., in a frozen block two and a half miles deep".[48][d]

Magnetic field and magnetosphere

Graph showing relative strength of Mercury's magnetic field

Despite its small size and slow 59-day-long rotation, Mercury has a significant, and apparently global, magnetic field. According to measurements taken by Mariner 10, it is about 1.1% as strong as Earth's. The magnetic field strength at the Mercurian equator is about 300 nT.[66][67] Like that of Earth, Mercury's magnetic field is dipolar.[65] Unlike Earth, Mercury's poles are nearly aligned with the planet's spin axis.[68] Measurements from both the Mariner 10 and MESSENGER space probes have indicated that the strength and shape of the magnetic field are stable.[68]

It is likely that this magnetic field is generated by way of a dynamo effect, in a manner similar to the magnetic field of Earth.[69][70] This dynamo effect would result from the circulation of the planet's iron-rich liquid core. Particularly strong tidal effects caused by the planet's high orbital eccentricity would serve to keep the core in the liquid state necessary for this dynamo effect.[71]

Mercury's magnetic field is strong enough to deflect the solar wind around the planet, creating a magnetosphere. The planet's magnetosphere, though small enough to fit within Earth,[65] is strong enough to trap solar wind plasma. This contributes to the space weathering of the planet's surface.[68] Observations taken by the Mariner 10 spacecraft detected this low energy plasma in the magnetosphere of the planet's nightside. Bursts of energetic particles were detected in the planet's magnetotail, which indicates a dynamic quality to the planet's magnetosphere.[65]

During its second flyby of the planet on October 6, 2008, MESSENGER discovered that Mercury's magnetic field can be extremely "leaky". The spacecraft encountered magnetic "tornadoes" – twisted bundles of magnetic fields connecting the planetary magnetic field to interplanetary space – that were up to 800 km wide or a third of the radius of the planet. These "tornadoes" form when magnetic fields carried by the solar wind connect to Mercury's magnetic field. As the solar wind blows past Mercury's field, these joined magnetic fields are carried with it and twist up into vortex-like structures. These twisted magnetic flux tubes, technically known as flux transfer events, form open windows in the planet's magnetic shield through which the solar wind may enter and directly impact Mercury's surface.[72]

The process of linking interplanetary and planetary magnetic fields, called magnetic reconnection, is common throughout the cosmos. It occurs in Earth's magnetic field, where it generates magnetic tornadoes as well. The MESSENGER observations show the reconnection rate is ten times higher at Mercury. Mercury's proximity to the Sun only accounts for about a third of the reconnection rate observed by MESSENGER.[72]

Orbit, rotation, and longitude

Orbit of Mercury (yellow). Dates refer to 2006.
Animation of Mercury's and Earth's revolution around the Sun

Mercury has the most eccentric orbit of all the planets; its eccentricity is 0.21 with its distance from the Sun ranging from 46,000,000 to 70,000,000 km (29,000,000 to 43,000,000 mi). It takes 87.969 Earth days to complete an orbit. The diagram on the right illustrates the effects of the eccentricity, showing Mercury's orbit overlaid with a circular orbit having the same semi-major axis. Mercury's higher velocity when it is near perihelion is clear from the greater distance it covers in each 5-day interval. In the diagram the varying distance of Mercury to the Sun is represented by the size of the planet, which is inversely proportional to Mercury's distance from the Sun. This varying distance to the Sun, combined with a 3:2 spin–orbit resonance of the planet's rotation around its axis, result in complex variations of the surface temperature.[14] This resonance makes a single day on Mercury last exactly two Mercury years, or about 176 Earth days.[73]

Mercury's orbit is inclined by 7 degrees to the plane of Earth's orbit (the ecliptic), as shown in the diagram on the right. As a result, transits of Mercury across the face of the Sun can only occur when the planet is crossing the plane of the ecliptic at the time it lies between Earth and the Sun. This occurs about every seven years on average.[74]

Mercury's axial tilt is almost zero,[75] with the best measured value as low as 0.027 degrees.[8] This is significantly smaller than that of Jupiter, which has the second smallest axial tilt of all planets at 3.1 degrees. This means that to an observer at Mercury's poles, the center of the Sun never rises more than 2.1 arcminutes above the horizon.[8]

At certain points on Mercury's surface, an observer would be able to see the Sun rise about halfway, then reverse and set before rising again, all within the same Mercurian day. This is because approximately four Earth days before perihelion, Mercury's angular orbital velocity equals its angular rotational velocity so that the Sun's apparent motion ceases; closer to perihelion, Mercury's angular orbital velocity then exceeds the angular rotational velocity. Thus, to a hypothetical observer on Mercury, the Sun appears to move in a retrograde direction. Four Earth days after perihelion, the Sun's normal apparent motion resumes.[14]

For the same reason, there are two points on Mercury's equator, 180 degrees apart in longitude, at either of which, around perihelion in alternate Mercurian years (once a Mercurian day), the Sun passes overhead, then reverses its apparent motion and passes overhead again, then reverses a second time and passes overhead a third time, taking a total of about 16 Earth-days for this entire process. In the other alternate Mercurian years, the same thing happens at the other of these two points. The amplitude of the retrograde motion is small, so the overall effect is that, for two or three weeks, the Sun is almost stationary overhead, and is at its most brilliant because Mercury is at perihelion, its closest to the Sun. This prolonged exposure to the Sun at its brightest makes these two points the hottest places on Mercury. Conversely, there are two other points on the equator, 90 degrees of longitude apart from the first ones, where the Sun passes overhead only when the planet is at aphelion in alternate years, when the apparent motion of the Sun in the Mercurian sky is relatively rapid. These points, which are the ones on the equator where the apparent retrograde motion of the Sun happens when it is crossing the horizon as described in the preceding paragraph, receive much less solar heat than the first ones described above.

Mercury attains inferior conjunction (nearest approach to Earth) every 116 Earth days on average,[3] but this interval can range from 105 days to 129 days due to the planet's eccentric orbit. Mercury can come as near as 82.2 Gm to Earth, and that is slowly declining: The next approach to within 82.1 Gm is in 2679, and to within 82 Gm in 4487, but it will not be closer to Earth than 80 Gm until AD 28,622.[76] Its period of retrograde motion as seen from Earth can vary from 8 to 15 days on either side of inferior conjunction. This large range arises from the planet's high orbital eccentricity.[14]

Longitude convention

The longitude convention for Mercury puts the zero of longitude at one of the two hottest points on the surface, as described above. However, when this area was first visited, by Mariner 10, this zero meridian was in darkness, so it was impossible to select a feature on the surface to define the exact position of the meridian. Therefore, a small crater further west was chosen, called Hun Kal, which provides the exact reference point for measuring longitude. The center of Hun Kal defines the 20° West meridian. A 1970 International Astronomical Union resolution suggests that longitudes be measured positively in the westerly direction on Mercury.[77] The two hottest places on the equator are therefore at longitudes 0°W and 180°W, and the coolest points on the equator are at longitudes 90°W and 270°W. However the MESSENGER project uses an east-positive convention.[78]

Spin–orbit resonance

After one orbit, Mercury has rotated 1.5 times, so after two complete orbits the same hemisphere is again illuminated.

For many years it was thought that Mercury was synchronously tidally locked with the Sun, rotating once for each orbit and always keeping the same face directed towards the Sun, in the same way that the same side of the Moon always faces Earth. Radar observations in 1965 proved that the planet has a 3:2 spin–orbit resonance, rotating three times for every two revolutions around the Sun; the eccentricity of Mercury's orbit makes this resonance stable—at perihelion, when the solar tide is strongest, the Sun is nearly still in Mercury's sky.[79]

The original reason astronomers thought it was synchronously locked was that, whenever Mercury was best placed for observation, it was always nearly at the same point in its 3:2 resonance, hence showing the same face. This is because, coincidentally, Mercury's rotation period is almost exactly half of its synodic period with respect to Earth. Due to Mercury's 3:2 spin–orbit resonance, a solar day (the length between two meridian transits of the Sun) lasts about 176 Earth days.[14] A sidereal day (the period of rotation) lasts about 58.7 Earth days.[14]

Simulations indicate that the orbital eccentricity of Mercury varies chaotically from nearly zero (circular) to more than 0.45 over millions of years due to perturbations from the other planets.[14][80] This is thought to explain Mercury's 3:2 spin–orbit resonance (rather than the more usual 1:1), because this state is more likely to arise during a period of high eccentricity.[81] Numerical simulations show that a future secular orbital resonant perihelion interaction with Jupiter may cause the eccentricity of Mercury's orbit to increase to the point where there is a 1% chance that the planet may collide with Venus within the next five billion years.[82][83]

Advance of perihelion

In 1859, the French mathematician and astronomer Urbain Le Verrier reported that the slow precession of Mercury's orbit around the Sun could not be completely explained by Newtonian mechanics and perturbations by the known planets. He suggested, among possible explanations, that another planet (or perhaps instead a series of smaller 'corpuscules') might exist in an orbit even closer to the Sun than that of Mercury, to account for this perturbation.[84] (Other explanations considered included a slight oblateness of the Sun.) The success of the search for Neptune based on its perturbations of the orbit of Uranus led astronomers to place faith in this possible explanation, and the hypothetical planet was named Vulcan, but no such planet was ever found.[85]

The perihelion precession of Mercury is 5600 arcseconds (1.5556°) per century relative to Earth, or 574.10±0.65 arcseconds per century[86] relative to the inertial ICFR. Newtonian mechanics, taking into account all the effects from the other planets, predicts a precession of 5557 arcseconds (1.5436°) per century.[86] In the early 20th century, Albert Einstein's General Theory of Relativity provided the explanation for the observed precession. The effect is small: the Mercurian relativistic perihelion advance excess is just 42.98 arcseconds per century; therefore, it requires a little over twelve million orbits for a full excess turn. Similar, but much smaller, effects operate for other planets: 8.62 arcseconds per century for Venus, 3.84 for Earth, 1.35 for Mars, and 10.05 for 1566 Icarus.[87][88]

Observation

Mosaic image by Mariner, 1974

Mercury's apparent magnitude varies between −2.6[10] (brighter than the brightest star Sirius) and about +5.7 (approximating the theoretical limit of naked-eye visibility). The extremes occur when Mercury is close to the Sun in the sky.[10][11] Observation of Mercury is complicated by its proximity to the Sun, as it is lost in the Sun's glare for much of the time. Mercury can be observed for only a brief period during either morning or evening twilight.[89]

Mercury can, like several other planets and the brightest stars, be seen during a total solar eclipse.[90]

Like the Moon and Venus, Mercury exhibits phases as seen from Earth. It is "new" at inferior conjunction and "full" at superior conjunction. The planet is rendered invisible from Earth on both of these occasions because of its being obscured by the Sun's disk.[89]

Mercury is technically brightest as seen from Earth when it is at a full phase. Although the planet is farthest away from Earth when it is full the greater illuminated area that is visible and the opposition brightness surge more than compensates for the distance.[10] The opposite is true for Venus, which appears brightest when it is a crescent, because it is much closer to Earth than when gibbous.[10][91]

Nonetheless, the brightest (full phase) appearance of Mercury is an essentially impossible time for practical observation, because of the extreme proximity of the Sun. Mercury is best observed at the first and last quarter, although they are phases of lesser brightness. The first and last quarter phases occur at greatest elongation east and west, respectively. At both of these times Mercury's separation from the Sun ranges anywhere from 17.9° at perihelion to 27.8° at aphelion.[92][93] At greatest elongation west, Mercury rises at its earliest before the Sun, and at greatest elongation east, it sets at its latest after the Sun.[94]

At tropical and subtropical latitudes, Mercury is more easily seen than at higher latitudes. In low latitudes and at the right times of year, the ecliptic intersects the horizon at a steep angle. When Mercury is vertically above the Sun in the sky and is at maximum elongation from the Sun (28 degrees), and when the Sun is 18 degrees below the horizon, so the sky is just completely dark,[e] Mercury is 10 degrees above the horizon. This is the greatest angle of elevation at which Mercury can be seen in a completely dark sky.

At temperate latitudes, Mercury is more often easily visible from Earth's Southern Hemisphere than from its Northern Hemisphere. This is because Mercury's maximum possible elongations west of the Sun always occur when it is early autumn in the Southern Hemisphere, whereas its maximum possible eastern elongations happen during late winter in the Southern Hemisphere.[94] In both of these cases, the angle Mercury strikes with the ecliptic is maximized, allowing it to rise several hours before the Sun in the former instance and not set until several hours after sundown in the latter in countries located at southern temperate zone latitudes, such as Argentina and South Africa.[94] By contrast, at the major population centers of the northern temperate latitudes, Mercury is never above the horizon of a more-or-less fully dark night sky.[citation needed]

Another possibility is looking up the planet during daylight with a nice blue sky and when close to greatest elongation or when gibbous (but not too gibbous as it is too close to the Sun !) the planet can be found easily even with a telescope as small as 8 cm aperture. But one has to know the exact coordinates and avoid pointing the telescope to the Sun. But modern pushto / goto telescopes can make this easy. This bypasses the limitation of twilight observing when the ecliptic has a small angle with the horizon (e.g. in autumn evenings).

Ground-based telescope observations of Mercury reveal only an illuminated partial disk with limited detail. The first of two spacecraft to visit the planet was Mariner 10, which mapped about 45% of its surface from 1974 to 1975. The second is the MESSENGER spacecraft, which after three Mercury flybys between 2008 and 2009, attained orbit around Mercury on March 17, 2011,[95] to study and map the rest of the planet.[96]

The Hubble Space Telescope cannot observe Mercury at all, due to safety procedures that prevent its pointing too close to the Sun.[97]

Naked-eye viewing

At the right times, and from the right latitudes on the Earth, Mercury is easy to see,[98] although many casual observers search for it without success. Mercury is seen most easily when it is close to its greatest elongation, which means that its angular separation from the Sun is greatest. Mercury can be near greatest western elongation, which means it is west of the Sun in the sky, so it is visible soon before sunrise, or greatest eastern elongation, which means it is visible soon after sunset. However, the exact dates of the greatest elongations are not the best ones on which to try to see Mercury. The phase of the planet greatly affects its apparent brightness. At greatest elongation, it is approximately at half phase. It is brighter when it is gibbous, which means that the best times to see Mercury are a few days before greatest eastern elongation, in the evening, or a few days after greatest western elongation, in the morning. The apparent inclination of the ecliptic to the horizon is also important. When the inclination is large, as occurs near the spring equinox in the evening, and near the autumnal equinox in the morning (this is true for observers in both hemispheres), Mercury is higher in the sky when the Sun is just below the horizon, which makes it easier to see than at other times. The inclination of the ecliptic is also greater for observers at low latitudes than high ones. It is helpful if Mercury is close to aphelion at the time of observation, because this makes it further from the Sun than at other times. However, it also makes the planet less brightly illuminated, so the visibility advantage is not great. At present, Mercury is fairly close to aphelion when viewed at greatest western elongation at the March equinox, or at greatest eastern elongation at the September equinox. (Over long periods of time, this changes as Mercury's orbit shifts.)[citation needed]

Putting all these factors together, the best time for an observer in the Southern Hemisphere to see Mercury is in the morning, near the March equinox, a few days after Mercury is at greatest western elongation, or in the evening, near the September equinox, a few days before greatest eastern elongation. An observer in the Northern Hemisphere cannot optimize all the factors simultaneously. Usually, the best chances of seeing the planet are in the evening, near the March equinox, a few days before greatest eastern elongation, or in the morning, near the September equinox, a few days after greatest western elongation. The inclination of the ecliptic is then large, but Mercury is not close to aphelion.[citation needed]

Mercury's period of revolution around the Sun is 88 days. It therefore makes about 4.15 revolutions around the Sun in one Earth-year. In successive years the position of Mercury on its orbit therefore shifts by 0.15 revolutions when seen on specific dates, such as the equinoxes. Therefore, if, for example, greatest eastern elongation happens on the March equinox of some year, about three years later greatest western elongation will happen near the March equinox, because the position of Mercury on its orbit at the equinox will have changed by about half (.45) a revolution. Thus, if the timings of elongations and equinoxes are unfavourable for observing Mercury in some year, they will be fairly favourable within about three years later.[citation needed] Furthermore, because the shift of 0.15 revolutions in a year makes up a seven-year cycle (0.15 × 7 ≈ 1.0), in the seventh year Mercury will follows almost exactly (earlier by 7 days) the sequence of phenomena it showed seven years before.[92]

Observation history

Ancient astronomers

The earliest known recorded observations of Mercury are from the Mul.Apin tablets. These observations were most likely made by an Assyrian astronomer around the 14th century BC.[99] The cuneiform name used to designate Mercury on the Mul.Apin tablets is transcribed as Udu.Idim.Gu\u4.Ud ("the jumping planet").[f][100] Babylonian records of Mercury date back to the 1st millennium BC. The Babylonians called the planet Nabu after the messenger to the gods in their mythology.[101]

The ancient Greeks of Hesiod's time knew the planet as Στίλβων (Stilbon), meaning "the gleaming", and Ἑρμάων (Hermaon).[102] Later Greeks called the planet Apollo when it was visible in the morning sky, and Hermes when visible in the evening. Around the 4th century BC, Greek astronomers came to understand that the two names referred to the same body, Hermes (Ἑρμής: Hermēs),[citation needed] a planetary name that is retained in modern Greek (Ερμής: Ermis).[103] The Romans named the planet after the swift-footed Roman messenger god, Mercury (Latin Mercurius), which they equated with the Greek Hermes, because it moves across the sky faster than any other planet.[104][105] The astronomical symbol for Mercury is a stylized version of Hermes' caduceus.[106]

The Roman-Egyptian astronomer Ptolemy wrote about the possibility of planetary transits across the face of the Sun in his work Planetary Hypotheses. He suggested that no transits had been observed either because planets such as Mercury were too small to see, or because the transits were too infrequent.[107]

Ibn al-Shatir's model for the appearances of Mercury, showing the multiplication of epicycles using the Tusi-couple, thus eliminating the Ptolemaic eccentrics and equant.

In ancient China, Mercury was known as Chen Xing (辰星), the Hour Star. It was associated with the direction north and the phase of water in the Wu Xing.[108] Modern Chinese, Korean, Japanese and Vietnamese cultures refer to the planet literally as the "water star" (水星), based on the Five elements.[109] Hindu mythology used the name Budha for Mercury, and this god was thought to preside over Wednesday.[110] The god Odin (or Woden) of Germanic paganism was associated with the planet Mercury and Wednesday.[111] The Maya may have represented Mercury as an owl (or possibly four owls; two for the morning aspect and two for the evening) that served as a messenger to the underworld.[112]

The ancient association of Mercury with Wednesday is still visible in the names of Wednesday in various modern languages of Latin descent, e.g. mercredi in French, miércoles in Spanish, or miercuri in Romanian. The names of the days of the week were, in late classical times, all related to the names of the seven bodies that were then considered to be planets.[g][citation needed]

In ancient Indian astronomy, the Surya Siddhanta, an Indian astronomical text of the 5th century, estimates the diameter of Mercury as 4,841 kilometres (3,008 mi), an error of less than 1% from the accepted diameter of 4,880 kilometres (3,032 mi). This estimate was based upon an inaccurate guess of the planet's angular diameter as 3.0 arcminutes (50 millidegrees).[citation needed]

In medieval Islamic astronomy, the Andalusian astronomer Abū Ishāq Ibrāhīm al-Zarqālī in the 11th century described the deferent of Mercury's geocentric orbit as being oval, like an egg or a pignon, although this insight did not influence his astronomical theory or his astronomical calculations.[113][114] In the 12th century, Ibn Bajjah observed "two planets as black spots on the face of the Sun", which was later suggested as the transit of Mercury and/or Venus by the Maragha astronomer Qotb al-Din Shirazi in the 13th century.[115] (Note that most such medieval reports of transits were later taken as observations of sunspots.[116])

In India, the Kerala school astronomer Nilakantha Somayaji in the 15th century developed a partially heliocentric planetary model in which Mercury orbits the Sun, which in turn orbits Earth, similar to the Tychonic system later proposed by Tycho Brahe in the late 16th century.[117]

Ground-based telescopic research

Transit of Mercury. Mercury is the small dot in the lower center, in front of the Sun. The dark area on the left of the solar disk is a sunspot.
Elongation is the angle between the Sun and the planet, with Earth as the reference point. Mercury appears close to the Sun.

The first telescopic observations of Mercury were made by Galileo in the early 17th century. Although he observed phases when he looked at Venus, his telescope was not powerful enough to see the phases of Mercury. In 1631, Pierre Gassendi made the first telescopic observations of the transit of a planet across the Sun when he saw a transit of Mercury predicted by Johannes Kepler. In 1639, Giovanni Zupi used a telescope to discover that the planet had orbital phases similar to Venus and the Moon. The observation demonstrated conclusively that Mercury orbited around the Sun.[14]

A rare event in astronomy is the passage of one planet in front of another (occultation), as seen from Earth. Mercury and Venus occult each other every few centuries, and the event of May 28, 1737 is the only one historically observed, having been seen by John Bevis at the Royal Greenwich Observatory.[118] The next occultation of Mercury by Venus will be on December 3, 2133.[119]

The difficulties inherent in observing Mercury mean that it has been far less studied than the other planets. In 1800, Johann Schröter made observations of surface features, claiming to have observed 20 km high mountains. Friedrich Bessel used Schröter's drawings to erroneously estimate the rotation period as 24 hours and an axial tilt of 70°.[120] In the 1880s, Giovanni Schiaparelli mapped the planet more accurately, and suggested that Mercury's rotational period was 88 days, the same as its orbital period due to tidal locking.[121] This phenomenon is known as synchronous rotation. The effort to map the surface of Mercury was continued by Eugenios Antoniadi, who published a book in 1934 that included both maps and his own observations.[65] Many of the planet's surface features, particularly the albedo features, take their names from Antoniadi's map.[122]

In June 1962, Soviet scientists at the Institute of Radio-engineering and Electronics of the USSR Academy of Sciences led by Vladimir Kotelnikov became first to bounce radar signal off Mercury and receive it, starting radar observations of the planet.[123][124][125] Three years later radar observations by Americans Gordon Pettengill and R. Dyce using 300-meter Arecibo Observatory radio telescope in Puerto Rico showed conclusively that the planet's rotational period was about 59 days.[126][127] The theory that Mercury's rotation was synchronous had become widely held, and it was a surprise to astronomers when these radio observations were announced. If Mercury were tidally locked, its dark face would be extremely cold, but measurements of radio emission revealed that it was much hotter than expected. Astronomers were reluctant to drop the synchronous rotation theory and proposed alternative mechanisms such as powerful heat-distributing winds to explain the observations.[128]

Italian astronomer Giuseppe Colombo noted that the rotation value was about two-thirds of Mercury's orbital period, and proposed that the planet's orbital and rotational periods were locked into a 3:2 rather than a 1:1 resonance.[129] Data from Mariner 10 subsequently confirmed this view.[130] This means that Schiaparelli's and Antoniadi's maps were not "wrong". Instead, the astronomers saw the same features during every second orbit and recorded them, but disregarded those seen in the meantime, when Mercury's other face was toward the Sun, because the orbital geometry meant that these observations were made under poor viewing conditions.[120]

Ground-based optical observations did not shed much further light on the innermost planet, but radio astronomers using interferometry at microwave wavelengths, a technique that enables removal of the solar radiation, were able to discern physical and chemical characteristics of the subsurface layers to a depth of several meters.[131][132] Not until the first space probe flew past Mercury did many of its most fundamental morphological properties become known. Moreover, recent technological advances have led to improved ground-based observations. In 2000, high-resolution lucky imaging observations were conducted by the Mount Wilson Observatory 1.5 meter Hale telescope. They provided the first views that resolved surface features on the parts of Mercury that were not imaged in the Mariner mission.[133] Most of the planet has been mapped by the Arecibo radar telescope, with 5 km resolution, including polar deposits in shadowed craters of what may be water ice.[134]

Research with space probes

MESSENGER being prepared for launch

Reaching Mercury from Earth poses significant technical challenges, because the planet orbits so much closer to the Sun than Earth. A Mercury-bound spacecraft launched from Earth must travel over 91 million kilometers into the Sun's gravitational potential well. Mercury has an orbital speed of 48 km/s, whereas Earth's orbital speed is 30 km/s. Thus the spacecraft must make a large change in velocity (delta-v) to enter a Hohmann transfer orbit that passes near Mercury, as compared to the delta-v required for other planetary missions.[135]

The potential energy liberated by moving down the Sun's potential well becomes kinetic energy; requiring another large delta-v change to do anything other than rapidly pass by Mercury. To land safely or enter a stable orbit the spacecraft would rely entirely on rocket motors. Aerobraking is ruled out because the planet has little atmosphere. A trip to Mercury requires more rocket fuel than that required to escape the Solar System completely. As a result, only two space probes have visited the planet so far.[136] A proposed alternative approach would use a solar sail to attain a Mercury-synchronous orbit around the Sun.[137]

Mariner 10

Main article: Mariner 10
Mariner 10, the first probe to visit the innermost planet

The first spacecraft to visit Mercury was NASA's Mariner 10 (1974–1975).[104] The spacecraft used the gravity of Venus to adjust its orbital velocity so that it could approach Mercury, making it both the first spacecraft to use this gravitational "slingshot" effect and the first NASA mission to visit multiple planets.[135] Mariner 10 provided the first close-up images of Mercury's surface, which immediately showed its heavily cratered nature, and revealed many other types of geological features, such as the giant scarps that were later ascribed to the effect of the planet shrinking slightly as its iron core cools.[138] Unfortunately, due to the length of Mariner 10's orbital period, the same face of the planet was lit at each of Mariner 10's close approaches. This made observation of both sides of the planet impossible,[139] and resulted in the mapping of less than 45% of the planet's surface.[140]

On March 27, 1974, two days before its first flyby of Mercury, Mariner 10's instruments began registering large amounts of unexpected ultraviolet radiation near Mercury. This led to the tentative identification of Mercury's moon. Shortly afterward, the source of the excess UV was identified as the star 31 Crateris, and Mercury's moon passed into astronomy's history books as a footnote.[citation needed]

The spacecraft made three close approaches to Mercury, the closest of which took it to within 327 km of the surface.[141] At the first close approach, instruments detected a magnetic field, to the great surprise of planetary geologists—Mercury's rotation was expected to be much too slow to generate a significant dynamo effect. The second close approach was primarily used for imaging, but at the third approach, extensive magnetic data were obtained. The data revealed that the planet's magnetic field is much like Earth's, which deflects the solar wind around the planet. The origin of Mercury's magnetic field is still the subject of several competing theories.[142]

On March 24, 1975, just eight days after its final close approach, Mariner 10 ran out of fuel. Because its orbit could no longer be accurately controlled, mission controllers instructed the probe to shut down.[143] Mariner 10 is thought to be still orbiting the Sun, passing close to Mercury every few months.[144]

MESSENGER

Main article: MESSENGER

A second NASA mission to Mercury, named MESSENGER (MErcury Surface, Space ENvironment, GEochemistry, and Ranging), was launched on August 3, 2004, from the Cape Canaveral Air Force Station aboard a Boeing Delta 2 rocket. It made a fly-by of Earth in August 2005, and of Venus in October 2006 and June 2007 to place it onto the correct trajectory to reach an orbit around Mercury.[145] A first fly-by of Mercury occurred on January 14, 2008, a second on October 6, 2008,[146] and a third on September 29, 2009.[147] Most of the hemisphere not imaged by Mariner 10 has been mapped during these fly-bys. The probe successfully entered an elliptical orbit around the planet on March 18, 2011. The first orbital image of Mercury was obtained on March 29, 2011. The probe finished a one-year mapping mission,[146] and then entered a one-year extended mission into 2013. In addition to continued observations and mapping of Mercury, MESSENGER observed the 2012 solar maximum.[148]

Mercury transiting the Sun as viewed by the Curiosity rover on Mars (June 3, 2014).[149]

The mission is designed to clear up six key issues: Mercury's high density, its geological history, the nature of its magnetic field, the structure of its core, whether it has ice at its poles, and where its tenuous atmosphere comes from. To this end, the probe is carrying imaging devices that will gather much higher resolution images of much more of the planet than Mariner 10, assorted spectrometers to determine abundances of elements in the crust, and magnetometers and devices to measure velocities of charged particles. Detailed measurements of tiny changes in the probe's velocity as it orbits will be used to infer details of the planet's interior structure.[26]

Curiosity rover

On June 3, 2014, the Curiosity rover on the planet Mars observed the planet Mercury transiting the Sun, marking the first time a planetary transit has been observed from a celestial body besides Earth.[149]

BepiColombo

Main article: BepiColombo

The European Space Agency is planning a joint mission with Japan called BepiColombo, which will orbit Mercury with two probes: one to map the planet and the other to study its magnetosphere.[150] Once launched in 2015, the spacecraft bus is expected to reach Mercury in 2019.[151] The bus will release a magnetometer probe into an elliptical orbit, then chemical rockets will fire to deposit the mapper probe into a circular orbit. Both probes will operate for a terrestrial year.[150] The mapper probe will carry an array of spectrometers similar to those on MESSENGER, and will study the planet at many different wavelengths including infrared, ultraviolet, X-ray and gamma ray.[152]

Comparison

Mercury, Venus, Earth, and Mars (terrestrial planets, to scale)

See also

Mercury, from Liber astronomiae, 1550

Notes

  1. ^ a b Pluto was considered a planet from its discovery in 1930 to 2006, but after that it has been classified as a dwarf planet. Pluto's orbital eccentricity is greater than that of Mercury. Pluto is also smaller than Mercury, but was assumed to be larger until 1976.
  2. ^ Mercury and Venus are the two closest planets to the Sun, and are the only planets in the Solar System without any natural satellites. The Sun's gravity would perturb the orbit of any such satellite until it crashed into the planet. These planets do have artificial satellites, but their orbits have to be adjusted occasionally by the use of rockets to compensate for the perturbation. After the rocket fuel runs out, these satellites will not survive for long.
  3. ^ In astronomy, the words "rotation" and "revolution" have different meanings. "Rotation" is the turning of a body about an axis that passes through the body, as in "Earth rotates once a day.". "Revolution" is motion around a centre that is external to the body, usually in orbit, as in "Earth takes a year for each revolution around the Sun.". The verbs "rotate" and "revolve" mean doing rotation and revolution, respectively.
  4. ^ If the area of Washington is about 177 km2 and 2.5 miles is taken to equal 4 km, Solomon's estimate would equal about 700 cubic kilometres of ice, which would have a mass of about 600 billion tons (6×1014 kg).
  5. ^ See Twilight#Astronomical twilight
  6. ^ Some sources precede the cuneiform transcription with "MUL". "MUL" is a cuneiform sign that was used in the Sumerian language to designate a star or planet, but it is not considered part of the actual name. The "4" is a reference number in the Sumero-Akkadian transliteration system to designate which of several syllables a certain cuneiform sign is most likely designating.
  7. ^ In Romance languages, i.e. those descended from Latin, the names of the weekdays, Monday to Friday, are still recognizably related to those of the "planets": the Moon (Luna), Mars, Mercury, Jupiter (Jove), and Venus. Saturday and Sunday have names that have been changed for religious reasons. In English, Saturday, Sunday, and Monday have names that are clearly related to those of Saturn, the Sun, and the Moon. The other days of the week have names of Germanic deities corresponding to the Roman gods whose names still govern Romance usage. For example, Wednesday is named after Wōden (or Wotan), a Germanic god who was roughly equivalent to the Roman god Mercury. For further information on this, see the article Names of the days of the week.

References

  1. ^ "mercurial". Merriam-Webster Online. Retrieved 2008-06-12. 
  2. ^ "Hermian". Wiktionary. August 2, 2010. Retrieved 2 November 2013. 
  3. ^ a b c d e f g h i j k "Mercury Fact Sheet". NASA Goddard Space Flight Center. November 30, 2007. Retrieved 2008-05-28. 
  4. ^ "The MeanPlane (Invariable plane) of the Solar System passing through the barycenter". 2009-04-03. Retrieved 2009-04-03.  (produced with Solex 10 written by Aldo Vitagliano; see also Invariable plane)
  5. ^ Yeomans, Donald K. (April 7, 2008). "HORIZONS Web-Interface for Mercury Major Body=1)". JPL Horizons On-Line Ephemeris System. Retrieved 2008-04-07.  – Select "Ephemeris Type: Orbital Elements", "Time Span: 2000-01-01 12:00 to 2000-01-02". ("Target Body: Mercury" and "Center: Sun" should be defaulted to.) Results are instantaneous osculating values at the precise J2000 epoch.
  6. ^ a b c d e f g h Munsell, Kirk; Smith, Harman; Harvey, Samantha (May 28, 2009). "Mercury: Facts & Figures". Solar System Exploration. NASA. Retrieved 2008-04-07. 
  7. ^ a b Seidelmann, P. Kenneth; Archinal, Brent A.; A'Hearn, Michael F. et al. (2007). "Report of the IAU/IAG Working Group on cartographic coordinates and rotational elements: 2006". Celestial Mechanics and Dynamical Astronomy 98 (3): 155–180. Bibcode:2007CeMDA..98..155S. doi:10.1007/s10569-007-9072-y.  edit
  8. ^ a b c Margot, J. L.; Peale, S. J.; Jurgens, R. F.; Slade, M. A.; Holin, I. V. (2007). "Large Longitude Libration of Mercury Reveals a Molten Core". Science 316 (5825): 710–714. Bibcode:2007Sci...316..710M. doi:10.1126/science.1140514. PMID 17478713. 
  9. ^ a b Mallama, A.; Wang, D.; Howard, R.A. (2002). "Photometry of Mercury from SOHO/LASCO and Earth". Icarus 155 (2): 253–264. Bibcode:2002Icar..155..253M. doi:10.1006/icar.2001.6723. 
  10. ^ a b c d e Mallama, A. (2011). "Planetary magnitudes". Sky and Telescope. 121(1): 51–56. 
  11. ^ a b Espenak, Fred (July 25, 1996). "Twelve Year Planetary Ephemeris: 1995–2006". NASA Reference Publication 1349. NASA. Retrieved 2008-05-23. 
  12. ^ a b c d Vasavada, Ashwin R.; Paige, David A.; Wood, Stephen E. (19 February 1999). "Near-Surface Temperatures on Mercury and the Moon and the Stability of Polar Ice Deposits". Icarus 141 (2): 179–193. Bibcode:1999Icar..141..179V. doi:10.1006/icar.1999.6175. Figure 3 with the "TWO model"; Figure 5 for pole. 
  13. ^ "Animated clip of orbit and rotation of Mercury". Sciencenetlinks.com. 
  14. ^ a b c d e f g h Strom, Robert G.; Sprague, Ann L. (2003). Exploring Mercury: the iron planet. Springer. ISBN 1-85233-731-1. 
  15. ^ Staff (May 8, 2003). "Mercury". US Geological Survey. Retrieved 2006-11-26. 
  16. ^ Lyttleton, R. A. (1969). "On the Internal Structures of Mercury and Venus". Astrophysics and Space Science 5 (1): 18. Bibcode:1969Ap&SS...5...18L. doi:10.1007/BF00653933. 
  17. ^ Gold, Lauren (May 3, 2007). "Mercury has molten core, Cornell researcher shows". Chronicle Online (Cornell University). Retrieved 2008-05-12. 
  18. ^ a b Finley, Dave (May 3, 2007). "Mercury's Core Molten, Radar Study Shows". National Radio Astronomy Observatory. Retrieved 2008-05-12. 
  19. ^ Spohn, Tilman; Sohl, Frank; Wieczerkowski, Karin; Conzelmann, Vera (2001). "The interior structure of Mercury: what we know, what we expect from BepiColombo". Planetary and Space Science 49 (14–15): 1561–1570. Bibcode:2001P&SS...49.1561S. doi:10.1016/S0032-0633(01)00093-9. 
  20. ^ Gallant, R. 1986. The National Geographic Picture Atlas of Our Universe. National Geographic Society, 2nd edition.
  21. ^ Anderson, J. D. et al. (July 10, 1996). "Shape and Orientation of Mercury from Radar Ranging Data". Icarus (Academic press) 124 (2): 690–697. Bibcode:1996Icar..124..690A. doi:10.1006/icar.1996.0242. 
  22. ^ Schenk, P.; Melosh, H. J. (March 1994). "Lobate Thrust Scarps and the Thickness of Mercury's Lithosphere". Abstracts of the 25th Lunar and Planetary Science Conference 1994: 1994LPI....25.1203S. Bibcode:1994LPI....25.1203S. 
  23. ^ a b c d Benz, W.; Slattery, W. L.; Cameron, A. G. W. (1988). "Collisional stripping of Mercury's mantle". Icarus 74 (3): 516–528. Bibcode:1988Icar...74..516B. doi:10.1016/0019-1035(88)90118-2. 
  24. ^ a b Cameron, A. G. W. (1985). "The partial volatilization of Mercury". Icarus 64 (2): 285–294. Bibcode:1985Icar...64..285C. doi:10.1016/0019-1035(85)90091-0. 
  25. ^ Weidenschilling, S. J. (1987). "Iron/silicate fractionation and the origin of Mercury". Icarus 35 (1): 99–111. Bibcode:1978Icar...35...99W. doi:10.1016/0019-1035(78)90064-7. 
  26. ^ a b Grayzeck, Ed. "MESSENGER Web Site". Johns Hopkins University. Retrieved 2008-04-07. 
  27. ^ "BepiColombo". ESA Science & Technology. European Space Agency. Retrieved 2008-04-07. 
  28. ^ "Messenger shines light on Mercury's formation". Chemistry World. Retrieved 1 May 2012. 
  29. ^ Staff (February 28, 2008). "Scientists see Mercury in a new light". Science Daily. Retrieved 2008-04-07. 
  30. ^ MESSENGER
  31. ^ Blue, Jennifer (April 11, 2008). "Gazetteer of Planetary Nomenclature". US Geological Survey. Retrieved 2008-04-11. 
  32. ^ a b Dunne, J. A.; Burgess, E. (1978). "Chapter Seven". The Voyage of Mariner 10 – Mission to Venus and Mercury. NASA History Office. Retrieved 2008-05-28. 
  33. ^ "Categories for Naming Features on Planets and Satellites". US Geological Survey. Retrieved 2011-08-20. 
  34. ^ Strom, Robert (1979). "Mercury: a post-Mariner assessment". Space Science Reviews 24: 3–70. Bibcode:1979SSRv...24....3S. doi:10.1007/BF00221842. 
  35. ^ Broadfoot, A. L.; S. Kumar; M. J. S. Belton; M. B. McElroy (July 12, 1974). "Mercury's Atmosphere from Mariner 10: Preliminary Results". Science 185 (4146): 166–169. Bibcode:1974Sci...185..166B. doi:10.1126/science.185.4146.166. PMID 17810510. 
  36. ^ Staff (August 5, 2003). "Mercury". U.S. Geological Survey. Retrieved 2008-04-07. 
  37. ^ Head, James W.; Solomon, Sean C. (1981). "Tectonic Evolution of the Terrestrial Planets". Science 213 (4503): 62–76. Bibcode:1981Sci...213...62H. doi:10.1126/science.213.4503.62. PMID 17741171. 
  38. ^ Morris, Jefferson (November 10, 2008). "Laser Altimetry". Aviation Week & Space Technology 169 (18): 18. "Mercury's crust is more analogous to a marbled cake than a layered cake." 
  39. ^ a b c d e f g Spudis, P. D. (2001). "The Geological History of Mercury". Workshop on Mercury: Space Environment, Surface, and Interior, Chicago: 100. Bibcode:2001mses.conf..100S. 
  40. ^ Ritzel, Rebecca (20 December 2012). "Ballet isn't rocket science, but the two aren't mutually exclusive, either". Washington Post (Washington DC, United States). Retrieved 22 December 2012. 
  41. ^ Shiga, David (January 30, 2008). "Bizarre spider scar found on Mercury's surface". NewScientist.com news service. 
  42. ^ Schultz, Peter H.; Gault, Donald E. (1975). "Seismic effects from major basin formations on the moon and Mercury". Earth, Moon, and Planets 12 (2): 159–175. Bibcode:1975Moon...12..159S. doi:10.1007/BF00577875. 
  43. ^ Wieczorek, Mark A.; Zuber, Maria T. (2001). "A Serenitatis origin for the Imbrian grooves and South Pole-Aitken thorium anomaly". Journal of Geophysical Research 106 (E11): 27853–27864. Bibcode:2001JGR...10627853W. doi:10.1029/2000JE001384. Retrieved 2008-05-12. 
  44. ^ Denevi, B. W.; Robinson; Robinson, M. S. (2008). "Albedo of Immature Mercurian Crustal Materials: Evidence for the Presence of Ferrous Iron". Lunar and Planetary Science 39: 1750. Bibcode:2008LPI....39.1750D. 
  45. ^ a b c Wagner, R. J.; Wolf, U.; Ivanov, B. A.; Neukum, G. (October 4–5, 2001). "Application of an Updated Impact Cratering Chronology Model to Mercury' s Time-Stratigraphic System". "Workshop on Mercury: Space Environment, Surface, and Interior. Proceedings of a workshop held at The Field Museum.". Chicago, IL: Lunar and Planetary Science Institute. p. 106. Bibcode:2001mses.conf..106W. 
  46. ^ Dzurisin, D. (October 10, 1978). "The tectonic and volcanic history of Mercury as inferred from studies of scarps, ridges, troughs, and other lineaments". Journal of Geophysical Research 83 (B10): 4883–4906. Bibcode:1978JGR....83.4883D. doi:10.1029/JB083iB10p04883. 
  47. ^ Van Hoolst, Tim; Jacobs, Carla (2003). "Mercury's tides and interior structure". Journal of Geophysical Research 108 (E11): 7. Bibcode:2003JGRE..108.5121V. doi:10.1029/2003JE002126. 
  48. ^ a b Chang, Kenneth (2012-11-29). "On Closest Planet to the Sun, NASA Finds Lots of Ice". New York Times. p. A3. Archived from the original on 2012-11-29. "Sean C. Solomon, the principal investigator for Messenger, said there was enough ice there to encase Washington, D.C., in a frozen block two and a half miles deep." 
  49. ^ Prockter, Louise (2005). Ice in the Solar System. Volume 26 (number 2). Johns Hopkins APL Technical Digest. Retrieved 2009-07-27. 
  50. ^ Lewis, John S. (2004). Physics and Chemistry of the Solar System (2nd ed.). Academic Press. p. 463. ISBN 0-12-446744-X. 
  51. ^ Murdock, T. L.; Ney, E. P. (1970). "Mercury: The Dark-Side Temperature". Science 170 (3957): 535–537. Bibcode:1970Sci...170..535M. doi:10.1126/science.170.3957.535. PMID 17799708. 
  52. ^ Lewis, John S. (2004). Physics and Chemistry of the Solar System. Academic Press. ISBN 978-0-12-446744-6. Retrieved 2008-06-03. 
  53. ^ Ingersoll, Andrew P.; Svitek, Tomas; Murray, Bruce C. (1992). "Stability of polar frosts in spherical bowl-shaped craters on the moon, Mercury, and Mars". Icarus 100 (1): 40–47. Bibcode:1992Icar..100...40I. doi:10.1016/0019-1035(92)90016-Z. 
  54. ^ Slade, M. A.; Butler, B. J.; Muhleman, D. O. (1992). "Mercury radar imaging – Evidence for polar ice". Science 258 (5082): 635–640. Bibcode:1992Sci...258..635S. doi:10.1126/science.258.5082.635. PMID 17748898. 
  55. ^ Williams, David R. (June 2, 2005). "Ice on Mercury". NASA Goddard Space Flight Center. Retrieved 2008-05-23. 
  56. ^ a b c Rawlins, K; Moses, J. I.; Zahnle, K.J. (1995). "Exogenic Sources of Water for Mercury's Polar Ice". Bulletin of the American Astronomical Society 27: 1117. Bibcode:1995DPS....27.2112R. 
  57. ^ Harmon, J. K.; Perillat, P. J.; Slade, M. A. (2001). "High-Resolution Radar Imaging of Mercury's North Pole". Icarus 149 (1): 1–15. Bibcode:2001Icar..149....1H. doi:10.1006/icar.2000.6544. 
  58. ^ Domingue, Deborah L., Koehn, Patrick L. et al. (2009). "Mercury's Atmosphere: A Surface-Bounded Exosphere". Space Science Reviews 131 (1–4): 161–186. Bibcode:2007SSRv..131..161D. doi:10.1007/s11214-007-9260-9. 
  59. ^ Hunten, D. M.; Shemansky, D. E.; Morgan, T. H. (1988). "The Mercury atmosphere". Mercury. University of Arizona Press. ISBN 0-8165-1085-7. Retrieved 2009-05-18. 
  60. ^ Lakdawalla, Emily (July 3, 2008). "MESSENGER Scientists 'Astonished' to Find Water in Mercury's Thin Atmosphere". Retrieved 2009-05-18. 
  61. ^ Zurbuchen, Thomas H., Raines, Jim M. et al. (2008). "MESSENGER Observations of the Composition of Mercury's Ionized Exosphere and Plasma Environment". Science 321 (5885): 90–92. Bibcode:2008Sci...321...90Z. doi:10.1126/science.1159314. PMID 18599777. 
  62. ^ "Instrument Shows What Planet Mercury Is Made Of". University of Michigan. June 30, 2008. Retrieved 2009-05-18. 
  63. ^ Killen, Rosemary; Cremonese, Gabrielle et al. (2007). "Processes that Promote and Deplete the Exosphere of Mercury". Space Science Reviews 132 (2–4): 433–509. Bibcode:2007SSRv..132..433K. doi:10.1007/s11214-007-9232-0. 
  64. ^ McClintock, William E.; Vervack Jr., Ronald J. et al. (2009). "MESSENGER Observations of Mercury's Exosphere: Detection of Magnesium and Distribution of Constituents". Science 324 (5927): 610–613. Bibcode:2009Sci...324..610M. doi:10.1126/science.1172525. PMID 19407195. 
  65. ^ a b c d e Beatty, J. Kelly; Petersen, Carolyn Collins; Chaikin, Andrew (1999). The New Solar System. Cambridge University Press. ISBN 0-521-64587-5. 
  66. ^ Seeds, Michael A. (2004). Astronomy: The Solar System and Beyond (4th ed.). Brooks Cole. ISBN 0-534-42111-3. 
  67. ^ Williams, David R. (January 6, 2005). "Planetary Fact Sheets". NASA National Space Science Data Center. Retrieved 2006-08-10. 
  68. ^ a b c Staff (January 30, 2008). "Mercury's Internal Magnetic Field". NASA. Retrieved 2008-04-07. 
  69. ^ Gold, Lauren (May 3, 2007). "Mercury has molten core, Cornell researcher shows". Cornell University. Retrieved 2008-04-07. 
  70. ^ Christensen, Ulrich R. (2006). "A deep dynamo generating Mercury's magnetic field". Nature 444 (7122): 1056–1058. Bibcode:2006Natur.444.1056C. doi:10.1038/nature05342. PMID 17183319. 
  71. ^ Spohn, T.; Sohl, F.; Wieczerkowski, K.; Conzelmann, V. (2001). "The interior structure of Mercury: what we know, what we expect from BepiColombo". Planetary and Space Science 49 (14–15): 1561–1570. Bibcode:2001P&SS...49.1561S. doi:10.1016/S0032-0633(01)00093-9. 
  72. ^ a b Steigerwald, Bill (June 2, 2009). "Magnetic Tornadoes Could Liberate Mercury's Tenuous Atmosphere". NASA Goddard Space Flight Center. Retrieved 2009-07-18. 
  73. ^ "Space Topics: Compare the Planets: Mercury, Venus, Earth, The Moon, and Mars". Planetary Society. Retrieved 2007-04-12. 
  74. ^ Espenak, Fred (April 21, 2005). "Transits of Mercury". NASA/Goddard Space Flight Center. Retrieved 2008-05-20. 
  75. ^ Biswas, Sukumar (2000). Cosmic Perspectives in Space Physics. Astrophysics and Space Science Library. Springer. p. 176. ISBN 0-7923-5813-9. 
  76. ^ Mercury Closest Approaches to Earth generated with:
    1. Solex 10 (Text Output file)
    2. Gravity Simulator charts
    3. JPL Horizons 1950–2200
    (3 sources are provided to address original research concerns and to support general long-term trends)
  77. ^ "USGS Astrogeology: Rotation and pole position for the Sun and planets (IAU WGCCRE)". Retrieved 22 October 2009. 
  78. ^ Archinal, B. A.; A’Hearn, M. F.; Bowell, E.; Conrad, A.; Consolmagno, G. J.; Courtin, R.; Fukushima, T.; Hestroffer, D.; Hilton, J. L.; Krasinsky, G. A.; Neumann, G.; Oberst, J.; Seidelmann, P. K.; Stooke, P.; Tholen, D. J.; Thomas, P. C.; Williams, I. P. (2010). "Report of the IAU Working Group on Cartographic Coordinates and Rotational Elements: 2009". Celestial Mechanics and Dynamical Astronomy 109 (2): 101–135. Bibcode:2011CeMDA.109..101A. doi:10.1007/s10569-010-9320-4. ISSN 0923-2958. 
  79. ^ Liu, Han-Shou; O'Keefe, John A. (1965). "Theory of Rotation for the Planet Mercury". Science 150 (3704): 1717. Bibcode:1965Sci...150.1717L. doi:10.1126/science.150.3704.1717. PMID 17768871. 
  80. ^ Correia, Alexandre C.M; Laskar, Jacques (2009). "Mercury's capture into the 3/2 spin–orbit resonance including the effect of core–mantle friction". Icarus 201 (1): 1. arXiv:0901.1843. Bibcode:2009Icar..201....1C. doi:10.1016/j.icarus.2008.12.034. 
  81. ^ Correia, Alexandre C. M.; Laskar, Jacques (2004). "Mercury's capture into the 3/2 spin–orbit resonance as a result of its chaotic dynamics". Nature 429 (6994): 848–850. Bibcode:2004Natur.429..848C. doi:10.1038/nature02609. PMID 15215857. 
  82. ^ Laskar, J. (2008-03-18). "Chaotic diffusion in the Solar System". Icarus 196 (1): 1–15. arXiv:0802.3371. Bibcode:2008Icar..196....1L. doi:10.1016/j.icarus.2008.02.017. 
  83. ^ Laskar, J.; Gastineau, M. (2009-06-11). "Existence of collisional trajectories of Mercury, Mars and Venus with the Earth". Nature 459 (7248): 817–819. Bibcode:2009Natur.459..817L. doi:10.1038/nature08096. PMID 19516336. 
  84. ^ U. Le Verrier (1859), (in French), "Lettre de M. Le Verrier à M. Faye sur la théorie de Mercure et sur le mouvement du périhélie de cette planète", Comptes rendus hebdomadaires des séances de l'Académie des sciences (Paris), vol. 49 (1859), pp. 379–383. (At p. 383 in the same volume Le Verrier's report is followed by another, from Faye, enthusiastically recommending to astronomers to search for a previously undetected intra-mercurial object.)
  85. ^ Baum, Richard; Sheehan, William (1997). In Search of Planet Vulcan, The Ghost in Newton's Clockwork Machine. New York: Plenum Press. ISBN 0-306-45567-6. 
  86. ^ a b Clemence, G. M. (1947). "The Relativity Effect in Planetary Motions". Reviews of Modern Physics 19 (4): 361–364. Bibcode:1947RvMP...19..361C. doi:10.1103/RevModPhys.19.361. 
  87. ^ Gilvarry, J. J. (1953). "Relativity Precession of the Asteroid Icarus". Physical Review 89 (5): 1046. Bibcode:1953PhRv...89.1046G. doi:10.1103/PhysRev.89.1046. 
  88. ^ Anonymous. "6.2 Anomalous Precession". Reflections on Relativity. MathPages. Retrieved 2008-05-22. 
  89. ^ a b Menzel, Donald H. (1964). A Field Guide to the Stars and Planets. The Peterson Field Guide Series. Boston: Houghton Mifflin Co. pp. 292–293. 
  90. ^ Tezel, Tunç (January 22, 2003). "Total Solar Eclipse of 2006 March 29". Department of Physics at Fizik Bolumu in Turkey. Retrieved 2008-05-24. 
  91. ^ Espenak, Fred (1996). "NASA Reference Publication 1349; Venus: Twelve year planetary ephemeris, 1995–2006". Twelve Year Planetary Ephemeris Directory. NASA. Archived from the original on 2000-08-17. Retrieved 2008-05-24. 
  92. ^ a b Walker, John. "Mercury Chaser's Calculator". Fourmilab Switzerland. Retrieved 2008-05-29.  (look at 1964 and 2013)
  93. ^ "Mercury Elongation and Distance". Retrieved 2008-05-30.  – Numbers generated using the Solar System Dynamics Group, Horizons On-Line Ephemeris System
  94. ^ a b c Kelly, Patrick, ed. (2007). Observer's Handbook 2007. Royal Astronomical Society of Canada. ISBN 0-9738109-3-9. 
  95. ^ Alers, Paul E. (March 17, 2011). "Celebrating Mercury Orbit". NASA Multimedia. Retrieved 2011-03-18. 
  96. ^ "NASA spacecraft now circling Mercury – a first". MSNBC. Mar 17, 2011. Retrieved 2011-03-24. 
  97. ^ Baumgardner, Jeffrey; Mendillo, Michael; Wilson, Jody K. (2000). "A Digital High-Definition Imaging System for Spectral Studies of Extended Planetary Atmospheres. I. Initial Results in White Light Showing Features on the Hemisphere of Mercury Unimaged by Mariner 10". The Astronomical Journal 119 (5): 2458–2464. Bibcode:2000AJ....119.2458B. doi:10.1086/301323. 
  98. ^ Bob King (2014). "Observing Alert: See Mercury’s Best Evening Show of the Year". 
  99. ^ Schaefer, Bradley E. (2007). "The Latitude and Epoch for the Origin of the Astronomical Lore in Mul.Apin". American Astronomical Society Meeting 210, #42.05 (American Astronomical Society) 38: 157. Bibcode:2007AAS...210.4205S. 
  100. ^ Hunger, Hermann; Pingree, David (1989). "MUL.APIN: An Astronomical Compendium in Cuneiform". Archiv für Orientforschung (Austria: Verlag Ferdinand Berger & Sohne Gesellschaft MBH) 24: 146. 
  101. ^ Staff (2008). "MESSENGER: Mercury and Ancient Cultures". NASA JPL. Retrieved 2008-04-07. 
  102. ^ Στίλβων, Ἑρμάων, Ἑρμῆς. Liddell, Henry George; Scott, Robert; A Greek–English Lexicon at the Perseus Project.
  103. ^ "Greek Names of the Planets". Retrieved 2012-07-14. "Ermis is the Greek name of the planet Mercury, which is the closest planet to the sun. It is named after the Greek God of commerce, Ermis or Hermes, who was also the messenger of the Ancient Greek gods."  See also the Greek article about the planet.
  104. ^ a b Dunne, J. A.; Burgess, E. (1978). "Chapter One". The Voyage of Mariner 10 – Mission to Venus and Mercury. NASA History Office. 
  105. ^ Antoniadi, Eugène Michel; Translated from French by Moore, Patrick (1974). The Planet Mercury. Shaldon, Devon: Keith Reid Ltd. pp. 9–11. ISBN 0-904094-02-2. 
  106. ^ Duncan, John Charles (1946). Astronomy: A Textbook. Harper & Brothers. p. 125. "The symbol for Mercury represents the Caduceus, a wand with two serpents twined around it, which was carried by the messenger of the gods." 
  107. ^ Goldstein, Bernard R. (1996). "The Pre-telescopic Treatment of the Phases and Apparent Size of Venus". Journal for the History of Astronomy 27: 1. Bibcode:1996JHA....27....1G. 
  108. ^ Kelley, David H.; Milone, E. F.; Aveni, Anthony F. (2004). Exploring Ancient Skies: An Encyclopedic Survey of Archaeoastronomy. Birkhäuser. ISBN 0-387-95310-8. 
  109. ^ China: De Groot, Jan Jakob Maria (1912). Religion in China: universism. a key to the study of Taoism and Confucianism. American lectures on the history of religions 10 (G. P. Putnam's Sons). p. 300. Retrieved 2010-01-08. 
    Japan: Crump, Thomas (1992). The Japanese numbers game: the use and understanding of numbers in modern Japan. Nissan Institute/Routledge Japanese studies series (Routledge). pp. 39–40. ISBN 0415056098. 
    Korea: Hulbert, Homer Bezaleel (1909). The passing of Korea. Doubleday, Page & company. p. 426. Retrieved 2010-01-08. 
  110. ^ Pujari, R.M.; Kolhe, Pradeep; Kumar, N. R. (2006). Pride of India: A Glimpse Into India's Scientific Heritage. Samskrita Bharati. ISBN 81-87276-27-4. 
  111. ^ Bakich, Michael E. (2000). The Cambridge Planetary Handbook. Cambridge University Press. ISBN 0-521-63280-3. 
  112. ^ Milbrath, Susan (1999). Star Gods of the Maya: Astronomy in Art, Folklore and Calendars. University of Texas Press. ISBN 0-292-75226-1. 
  113. ^ Samsó, Julio; Mielgo, Honorino; Honorino (1994). "Ibn al-Zarqālluh on Mercury". Journal for the History of Astronomy 25: 289–96 [292]. Bibcode:1994JHA....25..289S. 
  114. ^ Hartner, Willy (1955). "The Mercury Horoscope of Marcantonio Michiel of Venice". Vistas in Astronomy 1: 84–138. Bibcode:1955VA......1...84H. doi:10.1016/0083-6656(55)90016-7.  at pp. 118–122.
  115. ^ Ansari, S. M. Razaullah (2002). "History of oriental astronomy: proceedings of the joint discussion-17 at the 23rd General Assembly of the International Astronomical Union, organised by the Commission 41 (History of Astronomy), held in Kyoto, August 25–26, 1997". Springer. p. 137. ISBN 1-4020-0657-8. 
  116. ^ Goldstein, Bernard R. (1969). "Some Medieval Reports of Venus and Mercury Transits". Centaurus 14 (1): 49–59. Bibcode:1969Cent...14...49G. doi:10.1111/j.1600-0498.1969.tb00135.x. 
  117. ^ Ramasubramanian, K.; Srinivas, M. S.; Sriram, M. S. (1994). "Modification of the Earlier Indian Planetary Theory by the Kerala Astronomers (c. 1500 AD) and the Implied Heliocentric Picture of Planetary Motion". Current Science 66: 784–790. Retrieved 2010-04-23. 
  118. ^ Sinnott, RW; Meeus; Meeus, J (1986). "John Bevis and a Rare Occultation". Sky and Telescope 72: 220. Bibcode:1986S&T....72..220S. 
  119. ^ Ferris, Timothy (2003). Seeing in the Dark: How Amateur Astronomers. Simon and Schuster. ISBN 0-684-86580-7. 
  120. ^ a b Colombo, G.; Shapiro; Shapiro, I. I. (November 1965). "The Rotation of the Planet Mercury". SAO Special Report #188R 188. Bibcode:1965SAOSR.188.....C. 
  121. ^ Holden, E. S. (1890). "Announcement of the Discovery of the Rotation Period of Mercury [by Professor Schiaparelli]". Publications of the Astronomical Society of the Pacific 2 (7): 79. Bibcode:1890PASP....2...79H. doi:10.1086/120099. 
  122. ^ Merton E. Davies, et al. (1978). "Surface Mapping". Atlas of Mercury. NASA Office of Space Sciences. Retrieved 2008-05-28. 
  123. ^ Evans, J. V.; Brockelman, R. A.; Henry, J. C.; Hyde, G. M.; Kraft, L. G.; Reid, W. A.; Smith, W. W. (1965). "Radio Echo Observations of Venus and Mercury at 23 cm Wavelength". Astronomical Journal 70: 487–500. Bibcode:1965AJ.....70..486E. doi:10.1086/109772. 
  124. ^ Moore, Patrick (2000). The Data Book of Astronomy. New York: CRC Press. p. 483. ISBN 0-7503-0620-3. 
  125. ^ Butrica, Andrew J. (1996). "Chapter 5". To See the Unseen: A History of Planetary Radar Astronomy. NASA History Office, Washington D.C. ISBN 0-16-048578-9. 
  126. ^ Pettengill, G. H.; Dyce, R. B. (1965). "A Radar Determination of the Rotation of the Planet Mercury". Nature 206 (1240): 451–2. Bibcode:1965Natur.206Q1240P. doi:10.1038/2061240a0. 
  127. ^ Mercury at Eric Weisstein's 'World of Astronomy'
  128. ^ Murray, Bruce C.; Burgess, Eric (1977). Flight to Mercury. Columbia University Press. ISBN 0-231-03996-4. 
  129. ^ Colombo, G. (1965). "Rotational Period of the Planet Mercury". Nature 208 (5010): 575. Bibcode:1965Natur.208..575C. doi:10.1038/208575a0. 
  130. ^ Davies, Merton E. et al. (1976). "Mariner 10 Mission and Spacecraft". SP-423 Atlas of Mercury. NASA JPL. Retrieved 2008-04-07. 
  131. ^ Golden, Leslie M., A Microwave Interferometric Study of the Subsurface of the Planet Mercury (1977). PhD Dissertation, University of California, Berkeley
  132. ^ Mitchell, David L. and De Pater, Imke, Microwave Imaging of Mercury's Thermal Emission at Wavelengths from 0.3 to 20.5 cm (1994). Icarus, 110, 2–32
  133. ^ Dantowitz, R. F.; Teare, S. W.; Kozubal, M. J. (2000). "Ground-based High-Resolution Imaging of Mercury". Astronomical Journal 119 (4): 2455–2457. Bibcode:2000AJ....119.2455D. doi:10.1086/301328. 
  134. ^ Harmon, J. K. et al. (2007). "Mercury: Radar images of the equatorial and midlatitude zones". Icarus 187 (2): 374. Bibcode:2007Icar..187..374H. doi:10.1016/j.icarus.2006.09.026. 
  135. ^ a b Dunne, J. A. and Burgess, E. (1978). "Chapter Four". The Voyage of Mariner 10 – Mission to Venus and Mercury. NASA History Office. Retrieved 2008-05-28. 
  136. ^ "Mercury". NASA Jet Propulsion Laboratory. May 5, 2008. Retrieved 2008-05-29. 
  137. ^ Leipold, M.; Seboldt, W.; Lingner, S.; Borg, E.; Herrmann, A.; Pabsch, A.; Wagner, O.; Bruckner, J. (1996). "Mercury sun-synchronous polar orbiter with a solar sail". Acta Astronautica 39 (1): 143–151. Bibcode:1996AcAau..39..143L. doi:10.1016/S0094-5765(96)00131-2. 
  138. ^ Phillips, Tony (October 1976). "NASA 2006 Transit of Mercury". SP-423 Atlas of Mercury. NASA. Retrieved 2008-04-07. 
  139. ^ "BepiColumbo – Background Science". European Space Agency. Retrieved 2008-05-30. 
  140. ^ Tariq Malik (August 16, 2004). "MESSENGER to test theory of shrinking Mercury". USA Today. Retrieved 2008-05-23. 
  141. ^ Merton E. Davies, et al. (1978). "Mariner 10 Mission and Spacecraft". Atlas of Mercury. NASA Office of Space Sciences. Retrieved 2008-05-30. 
  142. ^ Ness, Norman F. (1978). "Mercury – Magnetic field and interior". Space Science Reviews 21 (5): 527–553. Bibcode:1978SSRv...21..527N. doi:10.1007/BF00240907. 
  143. ^ Dunne, J. A. and Burgess, E. (1978). "Chapter Eight". The Voyage of Mariner 10 – Mission to Venus and Mercury. NASA History Office. 
  144. ^ Grayzeck, Ed (April 2, 2008). "Mariner 10". NSSDC Master Catalog. NASA. Retrieved 2008-04-07. 
  145. ^ "MESSENGER Engine Burn Puts Spacecraft on Track for Venus". SpaceRef.com. 2005. Retrieved 2006-03-02. 
  146. ^ a b "Countdown to MESSENGER's Closest Approach with Mercury". Johns Hopkins University Applied Physics Laboratory. January 14, 2008. Retrieved 2008-05-30. 
  147. ^ "MESSENGER Gains Critical Gravity Assist for Mercury Orbital Observations". MESSENGER Mission News. September 30, 2009. Retrieved 2009-09-30. 
  148. ^ "NASA extends spacecraft's Mercury mission". UPI, 15 November 2011. Retrieved 2011-11-16.
  149. ^ a b Webster, Guy (June 10, 2014). "Mercury Passes in Front of the Sun, as Seen From Mars". NASA. Retrieved June 10, 2014. 
  150. ^ a b "ESA gives go-ahead to build BepiColombo". European Space Agency. February 26, 2007. Retrieved 2008-05-29. 
  151. ^ Fleming, Nic (January 18, 2008). "Star Trek-style ion engine to fuel Mercury craft". The Telegraph. Retrieved 2008-05-23. 
  152. ^ "Objectives". European Space Agency. February 21, 2006. Retrieved 2008-05-29. 

External links

Media related to Mercury (planet) at Wikimedia Commons