Atmosphere of Uranus

Summary

The atmosphere of Uranus is composed primarily of hydrogen and helium. At depth it is significantly enriched in volatiles (dubbed "ices") such as water, ammonia and methane. The opposite is true for the upper atmosphere, which contains very few gases heavier than hydrogen and helium due to its low temperature. Uranus's atmosphere is the coldest of all the planets, with its temperature reaching as low as 49 K.[1]

A whitish blue spherical planet against the black background of space
True-color image of Uranus by Voyager 2

The Uranian atmosphere can be divided into five main layers: the troposphere, between altitudes of −300[a] and 50 km and pressures from 100 to 0.1 bar; the stratosphere, spanning altitudes between 50 and 4000 km and pressures of between 0.1 and 10−10 bar; and the hot thermosphere (and exosphere) extending from an altitude of 4,056 km to several Uranian radii from the nominal surface at 1 bar pressure.[2] Unlike Earth's, Uranus's atmosphere has no mesosphere.

The troposphere hosts four cloud layers: methane clouds at about 1.2 bar, hydrogen sulfide and ammonia clouds at 3–10 bar, ammonium hydrosulfide clouds at 20–40 bar, and finally water clouds below 50 bar. Only the upper two cloud layers have been observed directly—the deeper clouds remain speculative. Above the clouds lie several tenuous layers of photochemical haze. Discrete bright tropospheric clouds are rare on Uranus, probably due to sluggish convection in the planet's interior. Nevertheless, observations of such clouds were used to measure the planet's zonal winds, which are remarkably fast with speeds up to 240 m/s.

Little is known about the Uranian atmosphere as to date only one spacecraft, Voyager 2, which passed by the planet in 1986, obtained some valuable compositional data. The Uranus Orbiter and Probe is scheduled to launch in 2031, arriving at Uranus in 2044. Its primary science objectives include a detailed study of Uranus' atmosphere.

Observation and exploration edit

 
Uranus's atmosphere taken during the Outer Planet Atmosphere Legacy (OPAL) program.

Although there is no well-defined solid surface within Uranus's interior, the outermost part of Uranus's gaseous envelope (the region accessible to remote sensing) is called its atmosphere.[2] Remote sensing capability extends down to roughly 300 km below the 1 bar level, with a corresponding pressure of around 100 bar and temperature of 320 K.[3]

The observational history of the Uranian atmosphere is long and full of error and frustration. Uranus is a relatively faint object, and its visible angular diameter is smaller than 5″.[4] The first spectra of Uranus were observed through a prism in 1869 and 1871 by Angelo Secchi and William Huggins, who found a number of broad dark bands, which they were unable to identify.[4] They also failed to detect any solar Fraunhofer lines—the fact later interpreted by Norman Lockyer as indicating that Uranus emitted its own light as opposed to reflecting light from the Sun.[4][5] In 1889 however, astronomers observed solar Fraunhofer lines in photographic ultraviolet spectra of the planet, proving once and for all that Uranus was shining by reflected light.[6] The nature of the broad dark bands in its visible spectrum remained unknown until the fourth decade of the twentieth century.[4]

Although Uranus is presently largely blank in appearance, it has been historically shown to have occasional features, such as in March and April 1884, when astronomers Henri Joseph Perrotin, Norman Lockyer, and Charles Trépied observed a bright, elongated spot (presumably a storm) circling the equator of the planet.[7]

 
Planet Uranus - North Pole - Cyclone (VLA; October 2021)

The key to deciphering Uranus's spectrum was found in the 1930s by Rupert Wildt and Vesto Slipher,[8] who found that the dark bands at 543, 619, 925, 865 and 890 nm belonged to gaseous methane.[4] They had never been observed before because they were very weak and required a long path length to be detected.[8] This meant that the atmosphere of Uranus was transparent to a much greater depth compared to those of other giant planets.[4] In 1950, Gerard Kuiper noticed another diffuse dark band in the spectrum of Uranus at 827 nm, which he failed to identify.[9] In 1952 Gerhard Herzberg, a future Nobel Prize winner, showed that this band was caused by the weak quadrupole absorption of molecular hydrogen, which thus became the second compound detected on Uranus.[10] Until 1986 only two gases, methane and hydrogen, were known in the Uranian atmosphere.[4] The far-infrared spectroscopic observation beginning from 1967 consistently showed the atmosphere of Uranus was in approximate thermal balance with incoming solar radiation (in other words, it radiated as much heat as it received from the Sun), and no internal heat source was required to explain observed temperatures.[11] No discrete features had been observed on Uranus prior to the Voyager 2 visit in 1986.[12]

In January 1986, the Voyager 2 spacecraft flew by Uranus at a minimal distance of 107,100 km[13] providing the first close-up images and spectra of its atmosphere. They generally confirmed that the atmosphere was made of mainly hydrogen and helium with around 2% methane.[14] The atmosphere appeared highly transparent and lacking thick stratospheric and tropospheric hazes. Only a limited number of discrete clouds were observed.[15]

In the 1990s and 2000s, observations by the Hubble Space Telescope and by ground-based telescopes equipped with adaptive optics systems (the Keck telescope and NASA Infrared Telescope Facility, for instance) made it possible for the first time to observe discrete cloud features from Earth.[16] Tracking them has allowed astronomers to re-measure wind speeds on Uranus, known before only from the Voyager 2 observations, and to study the dynamics of the Uranian atmosphere.[17]

Composition edit

The composition of the Uranian atmosphere is different from that of Uranus as a whole, consisting mainly of molecular hydrogen and helium.[18] The helium molar fraction, i.e. the number of helium atoms per molecule of hydrogen/helium, was determined from the analysis of Voyager 2 far infrared and radio occultation observations.[19] The currently accepted value is 0.152±0.033 in the upper troposphere, which corresponds to a mass fraction 0.262±0.048.[18][20] This value is very close to the protosolar helium mass fraction of 0.2741±0.0120,[21] indicating that helium has not settled towards the centre of the planet as it has in the gas giants.[22]

The third most abundant constituent of the Uranian atmosphere is methane (CH4),[23] the presence of which has been known for some time as a result of the ground-based spectroscopic observations.[18] Methane possesses prominent absorption bands in the visible and near-infrared, making Uranus aquamarine or cyan in colour.[24] Below the methane cloud deck at 1.3 bar methane molecules account for about 2.3%[25] of the atmosphere by molar fraction; about 10 to 30 times that found in the Sun.[18][19] The mixing ratio is much lower in the upper atmosphere due to the extremely low temperature at the tropopause, which lowers the saturation level and causes excess methane to freeze out.[26] Methane appears to be undersaturated in the upper troposphere above the clouds having a partial pressure of only 30% of the saturated vapor pressure there.[25] The concentration of less volatile compounds such as ammonia, water and hydrogen sulfide in the deep atmosphere is poorly known.[18] However, as with methane, their abundances are probably greater than solar values by a factor of at least 20 to 30,[27] and possibly by a factor of a few hundred.[28]

Knowledge of the isotopic composition of Uranus's atmosphere is very limited.[29] To date the only known isotope abundance ratio is that of deuterium to light hydrogen: 5.5+3.5
−1.5
×10−5
, which was measured by the Infrared Space Observatory (ISO) in the 1990s. It appears to be higher than the protosolar value of (2.25±0.35)×10−5 measured in Jupiter.[30] The deuterium is found almost exclusively in hydrogen deuteride molecules which it forms with normal hydrogen atoms.[31]

Infrared spectroscopy, including measurements with Spitzer Space Telescope (SST),[32] and UV occultation observations,[33] found trace amounts of complex hydrocarbons in the stratosphere of Uranus, which are thought to be produced from methane by photolysis induced by solar UV radiation.[34] They include ethane (C2H6), acetylene (C2H2),[33][35] methylacetylene (CH3C2H), diacetylene (C2HC2H).[36] Infrared spectroscopy also uncovered traces of water vapour,[37] carbon monoxide[38] and carbon dioxide in the stratosphere, which are likely to originate from an external source such as infalling dust and comets.[36]

Structure edit

 
Temperature profile of the Uranian troposphere and lower stratosphere. Cloud and haze layers are also indicated.

The Uranian atmosphere can be divided into three main layers: the troposphere, between altitudes of −300[a] and 50 km and pressures from 100 to 0.1 bar; the stratosphere, spanning altitudes between 50 and 4000 km and pressures between 0.1 and 10−10 bar; and the thermosphere/exosphere extending from 4000 km to as high as a few Uranus radii from the surface. There is no mesosphere.[2][39]

Troposphere edit

The troposphere is the lowest and densest part of the atmosphere and is characterised by a decrease in temperature with altitude.[2] The temperature falls from about 320 K at the base of the troposphere at −300 km to about 53 K at 50 km.[3][19] The temperature at the cold upper boundary of the troposphere (the tropopause) actually varies in the range between 49 and 57 K depending on planetary latitude, with the lowest temperature reached near 25° southern latitude.[40][41] The troposphere holds almost all of the mass of the atmosphere, and the tropopause region is also responsible for the vast majority of the planet's thermal far infrared emissions, thus determining its effective temperature of 59.1±0.3 K.[41][42]

The troposphere is believed to possess a highly complex cloud structure; water clouds are hypothesised to lie in the pressure range of 50 to 300 bar, ammonium hydrosulfide clouds in the range of 20 and 40 bar, ammonia or hydrogen sulfide clouds at between 3 and 10 bar and finally thin methane clouds at 1 to 2 bar.[3][24][27] Although Voyager 2 directly detected methane clouds,[25] all other cloud layers remain speculative. The existence of a hydrogen sulfide cloud layer is only possible if the ratio of sulfur and nitrogen abundances (S/N ratio) is significantly greater than its solar value of 0.16.[24] Otherwise all hydrogen sulfide would react with ammonia, producing ammonium hydrosulfide, and the ammonia clouds would appear instead in the pressure range 3–10 bar.[28] The elevated S/N ratio implies depletion of ammonia in the pressure range 20–40 bar, where the ammonium hydrosulfide clouds form. These can result from the dissolution of ammonia in water droplets within water clouds or in the deep water-ammonia ionic ocean.[27][28]

The exact location of the upper two cloud layers is somewhat controversial. Methane clouds were directly detected by Voyager 2 at 1.2–1.3 bar by radio occultation.[25] This result was later confirmed by an analysis of the Voyager 2 limb images.[24] The top of the deeper ammonia/hydrogen sulfide clouds were determined to be at 3 bar based on the spectroscopic data in the visible and near-infra spectral ranges (0.5–1 μm).[43] However a recent analysis of the spectroscopic data in the wavelength range 1–2.3 μm placed the methane cloudtops at 2 bar, and the top of the lower clouds at 6 bar.[44] This contradiction may be resolved when new data on methane absorption in Uranus's atmosphere are available.[b] The optical depth of the two upper cloud layers varies with latitude: both become thinner at the poles as compared to the equator, though in 2007 the methane cloud layer's optical depth had a local maximum at 45°S, where the southern polar collar is located (see below).[47]

The troposphere is very dynamic, exhibiting strong zonal winds, bright methane clouds,[48] dark spots[49] and noticeable seasonal changes. (see below)[50]

 
Temperature profiles in the stratosphere and thermosphere of Uranus. The shaded area is where hydrocarbons are concentrated.

Stratosphere edit

The stratosphere is the middle layer of the Uranian atmosphere, in which temperature generally increases with altitude from 53 K in the tropopause to between 800 and 850 K at the base thermosphere.[51] The heating of the stratosphere is caused by the downward heat conduction from the hot thermosphere[52][53] as well as by absorption of solar UV and IR radiation by methane and the complex hydrocarbons formed as a result of methane photolysis.[34][52] The methane enters the stratosphere through the cold tropopause, where its mixing ratio relative to molecular hydrogen is about 3 × 10–5, three times below saturation.[26] It decreases further to about 10−7 at the altitude corresponding to pressure of 0.1 mbar.[54]

Hydrocarbons heavier than methane are present in a relatively narrow layer between 160 and 320 km in altitude, corresponding to the pressure range from 10 to 0.1 mbar and temperatures from 100 to 130 K.[26][36] The most abundant stratospheric hydrocarbons after methane are acetylene and ethane, with mixing ratios of around 10−7.[54] Heavier hydrocarbons like methylacetylene and diacetylene have mixing ratios of about 10−10—three orders of magnitude lower.[36] The temperature and hydrocarbon mixing ratios in the stratosphere vary with time and latitude.[55][c] Complex hydrocarbons are responsible for the cooling of the stratosphere, especially acetylene, having a strong emission line at the wavelength of 13.7 μm.[52]

In addition to hydrocarbons, the stratosphere contains carbon monoxide, as well as traces of water vapor and carbon dioxide. The mixing ratio of carbon monoxide—3 × 10−8—is very similar to that of the hydrocarbons,[38] while the mixing ratios of carbon dioxide and water are about 10−11 and 8×10−9, respectively.[36][58] These three compounds are distributed relatively homogeneously in the stratosphere and are not confined to a narrow layer like hydrocarbons.[36][38]

Ethane, acetylene and diacetylene condense in the colder lower part of stratosphere[34] forming haze layers with an optical depth of about 0.01 in visible light.[59] Condensation occurs at approximately 14, 2.5 and 0.1 mbar for ethane, acetylene and diacetylene, respectively.[60][d] The concentration of hydrocarbons in the Uranian stratosphere is significantly lower than in the stratospheres of the other giant planets—the upper atmosphere of Uranus is very clean and transparent above the haze layers.[55] This depletion is caused by weak vertical mixing, and makes Uranus's stratosphere less opaque and, as a result, colder than those of other giant planets.[55][61] The hazes, like their parent hydrocarbons, are distributed unevenly across Uranus; at the solstice of 1986, when Voyager 2 passed by the planet, they were concentrated near the sunlit pole, making it dark in ultraviolet light.[62]

Thermosphere and ionosphere edit

The outermost layer of the Uranian atmosphere, extending for thousands of kilometres, is the thermosphere/exosphere, which has a uniform temperature of around 800 to 850 K.[52][63] This is much higher than, for instance, the 420 K observed in the thermosphere of Saturn.[64] The heat sources necessary to sustain such high temperatures are not understood, since neither solar FUV/EUV radiation nor auroral activity can provide the necessary energy.[51][63] The weak cooling efficiency due to the depletion of hydrocarbons in the stratosphere may contribute to this phenomenon.[55] In addition to molecular hydrogen, the thermosphere contains a large proportion of free hydrogen atoms,[51] while helium is thought to be absent here, because it separates diffusively at lower altitudes.[65]

The thermosphere and upper part of the stratosphere contain a large concentration of ions and electrons, forming the ionosphere of Uranus.[66] Radio occultation observations by the Voyager 2 spacecraft showed that the ionosphere lies between 1,000 and 10,000 km altitude and may include several narrow and dense layers between 1,000 and 3,500 km.[66][67] The electron density in the Uranian ionosphere is on average 104 cm−3,[68] reaching to as high as 105 cm−3 in the narrow layers in the stratosphere.[67] The ionosphere is mainly sustained by solar UV radiation and its density depends on the solar activity.[68][69] The auroral activity on Uranus is not as powerful as at Jupiter and Saturn and contributes little to the ionization.[e][70] The high electron density may be in part caused by the low concentration of hydrocarbons in the stratosphere.[55]

One of the sources of information about the ionosphere and thermosphere comes from ground-based measurements of the intense mid-infrared (3–4 μm) emissions of the trihydrogen cation (H3+).[68][71] The total emitted power is 1–2 × 1011 W—an order of magnitude higher than that the near-infrared hydrogen quadrupole emissions.[f][72] Trihydrogen cation functions as one of main coolers of the ionosphere.[73]

The upper atmosphere of Uranus is the source of the far ultraviolet (90–140 nm) emissions known as dayglow or electroglow, which, like the H3+ IR radiation, emanates exclusively from the sunlit part of the planet. This phenomenon, which occurs in the thermospheres of all giant planets and was mysterious for a time after its discovery, is interpreted as a UV fluorescence of atomic and molecular hydrogen excited by solar radiation or by photoelectrons.[74]

Hydrogen corona edit

The upper part of the thermosphere, where the mean free path of the molecules exceeds the scale height,[g] is called the exosphere.[75] The lower boundary of the Uranian exosphere, the exobase, is located at a height of about 6,500 km, or 1/4 of the planetary radius, above the surface.[75] The exosphere is unusually extended, reaching as far as several Uranian radii from the planet.[76][77] It is made mainly of hydrogen atoms and is often called the hydrogen corona of Uranus.[78] The high temperature and relatively high pressure at the base of the thermosphere explain in part why Uranus's exosphere is so vast.[h][77] The number density of atomic hydrogen in the corona falls slowly with the distance from the planet, remaining as high a few hundred atoms per cm3 at a few radii from Uranus.[80] The effects of this bloated exosphere include a drag on small particles orbiting Uranus, causing a general depletion of dust in the Uranian rings. The infalling dust in turn contaminates the upper atmosphere of the planet.[78]

Dynamics edit

 
Zonal wind speeds on Uranus. Shaded areas show the southern collar and its future northern counterpart. The red curve is a symmetrical fit to the data.

Uranus has a relatively bland appearance, lacking broad colorful bands and large clouds prevalent on Jupiter and Saturn.[16][62] Discrete features were only once observed in Uranus's atmosphere before 1986.[12][7] The most conspicuous features on Uranus observed by Voyager 2 were the dark low latitude region between −40° and −20° and bright southern polar cap.[62] The northern boundary of the cap was located at about −45° of latitude. The brightest zonal band was located near the edge of the cap at −50° to −45° and was then called a polar collar.[81] The southern polar cap, which existed at the time of the solstice in 1986, faded away in 1990s.[82] After the equinox in 2007, the southern polar collar started to fade away as well, while the northern polar collar located at 45° to 50° latitude (first appeared in 2007) have grown more conspicuous since then.[83]

The atmosphere of Uranus is calm compared to those of other giant planets. Only a limited number of small bright clouds at middle latitudes in both hemispheres[16] and one Uranus Dark Spot have been observed since 1986.[49] One of those bright cloud features, located at −34° of latitude and called Berg, probably existed continuously since at least 1986.[84] Nevertheless, the Uranian atmosphere has rather strong zonal winds blowing in the retrograde (counter to the rotation) direction near the equator, but switching to the prograde direction poleward of ±20° latitude.[85] The wind speeds are from −50 to −100 m/s at the equator increasing up to 240 m/s near 50° latitude.[82] The wind profile measured before the equinox of 2007 was slightly asymmetric with winds stronger in the southern hemisphere, although it turned out to be a seasonal effect as this hemisphere was continuously illuminated by the Sun before 2007.[82] After 2007 winds in the northern hemisphere accelerated while those in the southern one slowed down.

Uranus exhibits a considerable seasonal variation over its 84-year orbit. It is generally brighter near solstices and dimmer at equinoxes.[50] The variations are largely caused by changes in the viewing geometry: a bright polar region comes into view near solstices, while the dark equator is visible near equinoxes.[86] Still there exist some intrinsic variations of the reflectivity of the atmosphere: periodically fading and brightening polar caps as well as appearing and disappearing polar collars.[86]

See also edit

Notes edit

  1. ^ a b Negative altitudes refer to locations below the nominal surface at 1 bar.
  2. ^ Indeed, a recent analysis based on a new data set of the methane absorption coefficients shifted the clouds to 1.6 and 3 bar, respectively.[45][46]
  3. ^ In 1986 the stratosphere was poorer in hydrocarbons at the poles than near the equator;[26] at the poles the hydrocarbons were also confined to much lower altitudes.[56] Temperatures in the stratosphere may increase at the solstices and decrease at equinoxes by as much as 50 K.[57]
  4. ^ At these altitudes the temperature has local maxima, which may be caused by absorption of solar radiation by haze particles.[18]
  5. ^ The total power input into the aurora is 3–7 × 1010 W—insufficient to heat up the thermosphere.[70]
  6. ^ The hot thermosphere of Uranus produces hydrogen quadrupole emission lines in the near-infrared part of the spectrum (1.8–2.5 μm) with the total emitted power of 1–2 × 1010 W. The power emitted by molecular hydrogen in the far infrared part of the spectrum is about 2 × 1011 W.[72]
  7. ^ The scale height sh is defined as sh = RT/(Mgj), where R = 8.31 J/mol/K is the gas constant, M ≈ 0.0023 kg/mol is the average molar mass in the Uranian atmosphere,[18] T is temperature and gj ≈ 8.9 m/s2 is the gravitational acceleration at the surface of Uranus. As the temperature varies from 53 K in the tropopause up to 800 K in the thermosphere, the scale height changes from 20 to 400 km.
  8. ^ The corona contains a significant population of supra-thermal (energy of up to 2 eV) hydrogen atoms. Their origin is unclear, but they may be produced by the same mechanism that heats the thermosphere.[79]

Citations edit

  1. ^ Williams, Matt (December 16, 2014). "What is the average surface temperature of the planets in our solar system?". phys.org. Retrieved 2022-04-20.
  2. ^ a b c d Lunine 1993, pp. 219–222.
  3. ^ a b c de Pater Romani et al. 1991, p. 231, Fig. 13.
  4. ^ a b c d e f g Fegley Gautier et al. 1991, pp. 151–154.
  5. ^ Lockyer 1889.
  6. ^ Huggins 1889.
  7. ^ a b Perrotin, Henri (1 May 1884). "The Aspect of Uranus". Nature. 30: 21. Retrieved 4 November 2018.
  8. ^ a b Adel & Slipher 1934.
  9. ^ Kuiper 1949.
  10. ^ Herzberg 1952.
  11. ^ Pearl Conrath et al. 1990, pp. 12–13, Table I.
  12. ^ a b Smith 1984, pp. 213–214.
  13. ^ Stone 1987, p. 14,874, Table 3.
  14. ^ Fegley Gautier et al. 1991, pp. 155–158, 168–169.
  15. ^ Smith Soderblom et al. 1986, pp. 43–49.
  16. ^ a b c Sromovsky & Fry 2005, pp. 459–460.
  17. ^ Sromovsky & Fry 2005, p. 469, Fig.5.
  18. ^ a b c d e f g Lunine 1993, pp. 222–230.
  19. ^ a b c Tyler Sweetnam et al. 1986, pp. 80–81.
  20. ^ Conrath Gautier et al. 1987, p. 15,007, Table 1.
  21. ^ Lodders 2003, pp. 1, 228–1, 230.
  22. ^ Conrath Gautier et al. 1987, pp. 15, 008–15, 009.
  23. ^ NASA NSSDC, Uranus Fact Sheet Archived 2011-08-04 at the Wayback Machine (retrieved 7 Oc 2015)
  24. ^ a b c d Lunine 1993, pp. 235–240.
  25. ^ a b c d Lindal Lyons et al. 1987, pp. 14, 987, 14, 994–14, 996.
  26. ^ a b c d Bishop Atreya et al. 1990, pp. 457–462.
  27. ^ a b c Atreya & Wong 2005, pp. 130–131.
  28. ^ a b c de Pater Romani et al. 1989, pp. 310–311.
  29. ^ Encrenaz 2005, pp. 107–110.
  30. ^ Encrenaz 2003, pp. 98–100, Table 2 on p. 96.
  31. ^ Feuchtgruber Lellouch et al. 1999.
  32. ^ Burgdorf Orton et al. 2006, pp. 634–635.
  33. ^ a b Bishop Atreya et al. 1990, p. 448.
  34. ^ a b c Summers & Strobel 1989, pp. 496–497.
  35. ^ Encrenaz 2003, p. 93.
  36. ^ a b c d e f Burgdorf Orton et al. 2006, p. 636.
  37. ^ Encrenaz 2003, p. 92.
  38. ^ a b c Encrenaz Lellouch et al. 2004, p. L8.
  39. ^ Herbert Sandel et al. 1987, p. 15,097, Fig. 4.
  40. ^ Lunine 1993, pp. 240–245.
  41. ^ a b Hanel Conrath et al. 1986, p. 73.
  42. ^ Pearl Conrath et al. 1990, p. 26, Table IX.
  43. ^ Sromovsky Irwin et al. 2006, pp. 591–592.
  44. ^ Sromovsky Irwin et al. 2006, pp. 592–593.
  45. ^ Fry & Sromovsky 2009.
  46. ^ Irwin Teanby et al. 2010, p. 913.
  47. ^ Irwin Teanby et al. 2007, pp. L72–L73.
  48. ^ Sromovsky & Fry 2005, p. 483.
  49. ^ a b Hammel Sromovsky et al. 2009, p. 257.
  50. ^ a b Hammel & Lockwood 2007, pp. 291–293.
  51. ^ a b c Herbert Sandel et al. 1987, pp. 15, 101–15, 102.
  52. ^ a b c d Lunine 1993, pp. 230–234.
  53. ^ Young 2001, pp. 241–242.
  54. ^ a b Summers & Strobel 1989, pp. 497, 502, Fig. 5a.
  55. ^ a b c d e Herbert & Sandel 1999, pp. 1, 123–1, 124.
  56. ^ Herbert & Sandel 1999, pp. 1, 130–1, 131.
  57. ^ Young 2001, pp. 239–240, Fig. 5.
  58. ^ Encrenaz 2005, p. 111, Table IV.
  59. ^ Pollack Rages et al. 1987, p. 15,037.
  60. ^ Lunine 1993, p. 229, Fig. 3.
  61. ^ Bishop Atreya et al. 1990, pp. 462–463.
  62. ^ a b c Smith Soderblom et al. 1986, pp. 43–46.
  63. ^ a b Herbert & Sandel 1999, pp. 1, 122–1, 123.
  64. ^ Miller Aylward et al. 2005, p. 322, Table I.
  65. ^ Herbert Sandel et al. 1987, pp. 15, 107–15, 108.
  66. ^ a b Tyler Sweetnam et al. 1986, p. 81.
  67. ^ a b Lindal Lyons et al. 1987, p. 14,992, Fig. 7.
  68. ^ a b c Trafton Miller et al. 1999, pp. 1, 076–1, 078.
  69. ^ Encrenaz Drossart et al. 2003, pp. 1, 015–1, 016.
  70. ^ a b Herbert & Sandel 1999, pp. 1, 133–1, 135.
  71. ^ Lam Miller et al. 1997, pp. L75–76.
  72. ^ a b Trafton Miller et al. 1999, pp. 1, 073–1, 076.
  73. ^ Miller Achilleos et al. 2000, pp. 2, 496–2, 497.
  74. ^ Herbert & Sandel 1999, pp. 1, 127–1, 128, 1, 130–1, 131.
  75. ^ a b Herbert & Hall 1996, p. 10,877.
  76. ^ Herbert & Hall 1996, p. 10,879, Fig. 2.
  77. ^ a b Herbert & Sandel 1999, p. 1,124.
  78. ^ a b Herbert Sandel et al. 1987, pp. 15, 102–15, 104.
  79. ^ Herbert & Hall 1996, pp. 10, 880–10, 882.
  80. ^ Herbert & Hall 1996, pp. 10, 879–10, 880.
  81. ^ Rages Hammel et al. 2004, p. 548.
  82. ^ a b c Sromovsky & Fry 2005, pp. 470–472, 483, Table 7, Fig. 6.
  83. ^ Sromovsky Fry et al. 2009, p. 265.
  84. ^ Sromovsky & Fry 2005, pp. 474–482.
  85. ^ Smith Soderblom et al. 1986, pp. 47–49.
  86. ^ a b Hammel & Lockwood 2007, pp. 293–296.

References edit

  • Adel, A.; Slipher, V. (1934). "The Constitution of the Atmospheres of the Giant Planets". Physical Review. 46 (10): 902. Bibcode:1934PhRv...46..902A. doi:10.1103/PhysRev.46.902.
  • Atreya, Sushil K.; Wong, Ah-San (2005). "Coupled Clouds and Chemistry of the Giant Planets — A Case for Multiprobes" (PDF). Space Science Reviews. 116 (1–2): 121–136. Bibcode:2005SSRv..116..121A. doi:10.1007/s11214-005-1951-5. hdl:2027.42/43766. S2CID 31037195.
  • Bishop, J.; Atreya, S. K.; Herbert, F.; Romani, P. (December 1990). "Reanalysis of voyager 2 UVS occultations at Uranus: Hydrocarbon mixing ratios in the equatorial stratosphere" (PDF). Icarus. 88 (2): 448–464. Bibcode:1990Icar...88..448B. doi:10.1016/0019-1035(90)90094-P. hdl:2027.42/28293.
  • Burgdorf, M.; Orton, G.; Vancleve, J.; Meadows, V.; Houck, J. (October 2006). "Detection of new hydrocarbons in Uranus' atmosphere by infrared spectroscopy". Icarus. 184 (2): 634–637. Bibcode:2006Icar..184..634B. doi:10.1016/j.icarus.2006.06.006.
  • Conrath, B.; Gautier, D.; Hanel, R.; Lindal, G.; Marten, A. (1987). "The Helium Abundance of Uranus from Voyager Measurements". Journal of Geophysical Research. 92 (A13): 15003–15010. Bibcode:1987JGR....9215003C. doi:10.1029/JA092iA13p15003.
  • Encrenaz, Thérèse (February 2003). "ISO observations of the giant planets and Titan: what have we learnt?". Planetary and Space Science. 51 (2): 89–103. Bibcode:2003P&SS...51...89E. doi:10.1016/S0032-0633(02)00145-9.
  • Encrenaz, T.; Drossart, P.; Orton, G.; Feuchtgruber, H.; Lellouch, E.; Atreya, S. K. (December 2003). "The rotational temperature and column density of H3+ in Uranus" (PDF). Planetary and Space Science. 51 (14–15): 1013–1016. Bibcode:2003P&SS...51.1013E. doi:10.1016/j.pss.2003.05.010.
  • Encrenaz, T.; Lellouch, E.; Drossart, P.; Feuchtgruber, H.; Orton, G. S.; Atreya, S. K. (January 2004). "First detection of CO in Uranus" (PDF). Astronomy and Astrophysics. 413 (2): L5–L9. Bibcode:2004A&A...413L...5E. doi:10.1051/0004-6361:20034637.
  • Encrenaz, Thérèse (January 2005). "Neutral Atmospheres of the Giant Planets: An Overview of Composition Measurements". Space Science Reviews. 116 (1–2): 99–119. Bibcode:2005SSRv..116...99E. doi:10.1007/s11214-005-1950-6. S2CID 119681087.
  • Fegley, Bruce Jr.; Gautier, Daniel; Owen, Tobias; Prinn, Ronald G. (1991). "Spectroscopy and chemistry of the atmosphere of Uranus" (PDF). In Bergstrahl, Jay T.; Miner, Ellis D.; Matthews, Mildred Shapley (eds.). Uranus. University of Arizona Press. ISBN 978-0-8165-1208-9. OCLC 22625114.
  • Feuchtgruber, H.; Lellouch, E.; Bézard, B.; Encrenaz, Th.; de Graauw, Th.; Davis, G. R. (January 1999). "Detection of HD in the atmospheres of Uranus and Neptune: a new determination of the D/H ratio". Astronomy and Astrophysics. 341: L17–L21. Bibcode:1999A&A...341L..17F.
  • Fry, Patrick M.; Sromovsky, L. A. (September 2009). Implications of New Methane Absorption Coefficients on Uranus Vertical Structure Derived from Near-IR Spectra. DPS meeting #41, #14.06. American Astronomical Society. Bibcode:2009DPS....41.1406F.
  • Hammel, H. B.; Lockwood, G. W. (January 2007). "Long-term atmospheric variability on Uranus and Neptune". Icarus. 186 (1): 291–301. Bibcode:2007Icar..186..291H. doi:10.1016/j.icarus.2006.08.027.
  • Hammel, H. B.; Sromovsky, L. A.; Fry, P. M.; Rages, K.; Showalter, M.; de Pater, I.; van Dam, M. A.; LeBeau, R. P.; Deng, X. (May 2009). "The Dark Spot in the atmosphere of Uranus in 2006: Discovery, description, and dynamical simulations" (PDF). Icarus. 201 (1): 257–271. Bibcode:2009Icar..201..257H. doi:10.1016/j.icarus.2008.08.019. Archived from the original (PDF) on 2011-07-19.
  • Hanel, R.; Conrath, B.; Flasar, F. M.; Kunde, V.; Maguire, W.; Pearl, J.; Pirraglia, J.; Samuelson, R.; Cruikshank, D. (4 July 1986). "Infrared Observations of the Uranian System". Science. 233 (4759): 70–74. Bibcode:1986Sci...233...70H. doi:10.1126/science.233.4759.70. PMID 17812891. S2CID 29994902.
  • Herbert, F.; Sandel, B. R.; Yelle, R. V.; Holberg, J. B.; Broadfoot, A. L.; Shemansky, D. E.; Atreya, S. K.; Romani, P. N. (December 30, 1987). "The Upper Atmosphere of Uranus: EUV Occultations Observed by Voyager 2" (PDF). Journal of Geophysical Research. 92 (A13): 15, 093–15, 109. Bibcode:1987JGR....9215093H. doi:10.1029/JA092iA13p15093.
  • Herbert, F.; Hall, D. T. (May 1996). "Atomic hydrogen corona of Uranus". Journal of Geophysical Research. 101 (A5): 10, 877–10, 885. Bibcode:1996JGR...10110877H. doi:10.1029/96JA00427.
  • Herbert, Floyd; Sandel, Bill R. (August–September 1999). "Ultraviolet observations of Uranus and Neptune". Planetary and Space Science. 47 (8–9): 1, 119–1, 139. Bibcode:1999P&SS...47.1119H. doi:10.1016/S0032-0633(98)00142-1.
  • Herzberg, G. (May 1952). "Spectroscopic evidence of molecular hydrogen in the atmospheres of Uranus and Neptune". The Astrophysical Journal. 115: 337–340. Bibcode:1952ApJ...115..337H. doi:10.1086/145552.
  • Huggins, William (June 1889). "The spectrum of Uranus". Monthly Notices of the Royal Astronomical Society. 49: 404. Bibcode:1889MNRAS..49Q.404H. doi:10.1093/mnras/49.8.403a.
  • Irwin, P. G. J.; Teanby, N. A.; Davis, G. R. (2007-08-10). "Latitudinal Variations in Uranus' Vertical Cloud Structure from UKIRT UIST Observations". The Astrophysical Journal. 665 (1). The American Astronomical Society: L71–L74. Bibcode:2007ApJ...665L..71I. doi:10.1086/521189.
  • Irwin, P. G. J.; Teanby, N. A.; Davis, G. R. (August 2010). "Revised vertical cloud structure of Uranus from UKIRT/UIST observations and changes seen during Uranus' Northern Spring Equinox from 2006 to 2008: Application of new methane absorption data and comparison with Neptune". Icarus. 208 (2): 913–926. Bibcode:2010Icar..208..913I. doi:10.1016/j.icarus.2010.03.017. S2CID 123043940.
  • Kuiper, G. P. (May 1949). "New absorptions in the Uranian atmosphere". The Astrophysical Journal. 109: 540–541. Bibcode:1949ApJ...109..540K. doi:10.1086/145161.
  • Lam, H. A.; Miller, S.; Joseph, R. D.; Geballe, T. R.; Trafton, L. M.; Tennyson, J.; Ballester, G. E. (1997-01-01). "Variation in the H3+ Emission of Uranus" (PDF). The Astrophysical Journal. 474 (1). The American Astronomical Society: L73–L76. Bibcode:1997ApJ...474L..73L. doi:10.1086/310424.
  • Lindal, G. F.; Lyons, J. R.; Sweetnam, D. N.; Eshleman, V. R.; Hinson, D. P.; Tyler, G. L. (December 30, 1987). "The Atmosphere of Uranus: Results of Radio Occultation Measurements with Voyager 2". Journal of Geophysical Research. 92 (A13). American Geophysical Union: 14, 987–15, 001. Bibcode:1987JGR....9214987L. doi:10.1029/JA092iA13p14987.
  • Lockyer, J. N. (June 1889). "Note on the Spectrum of Uranus". Astronomische Nachrichten. 121 (24): 369. Bibcode:1889AN....121..369L. doi:10.1002/asna.18891212402.
  • Lodders, Katharina (July 10, 2003). "Solar System Abundances and Condensation Temperatures of the Elements" (PDF). The Astrophysical Journal. 591 (2). The American Astronomical Society: 1220–1247. Bibcode:2003ApJ...591.1220L. doi:10.1086/375492. S2CID 42498829. Archived from the original (PDF) on November 7, 2015. Retrieved September 2, 2015.
  • Lunine, Jonathan I. (September 1993). "The Atmospheres of Uranus and Neptune". Annual Review of Astronomy and Astrophysics. 31: 217–263. Bibcode:1993ARA&A..31..217L. doi:10.1146/annurev.aa.31.090193.001245.
  • Miller, Steven; Achilleos, Nick; Ballester, Gilda E.; Geballe, Thomas R.; Joseph, Robert D.; Prangé, Renee; Rego, Daniel; Stallard, Tom; Tennyson, Jonathan; Trafton, Laurence M.; Waite, J. Hunter Jr (15 September 2000). "The role of H3+ in planetary atmospheres" (PDF). Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences. 358 (1774): 2485–2502. Bibcode:2000RSPTA.358.2485M. doi:10.1098/rsta.2000.0662. S2CID 124490318.
  • Miller, Steve; Aylward, Alan; Millward, George (January 2005). "Giant Planet Ionospheres and Thermospheres: The Importance of Ion-Neutral Coupling". Space Science Reviews. 116 (1–2): 319–343. Bibcode:2005SSRv..116..319M. doi:10.1007/s11214-005-1960-4. S2CID 119906560.
  • Rages, K. A.; Hammel, H. B.; Friedson, A. J. (11 September 2004). "Evidence for temporal change at Uranus' south pole". Icarus. 172 (2): 548–554. Bibcode:2004Icar..172..548R. doi:10.1016/j.icarus.2004.07.009.
  • de Pater, I.; Romani, P. N.; Atreya, S. K. (December 1989). "Uranus Deep Atmosphere Revealed" (PDF). Icarus. 82 (2): 288–313. Bibcode:1989Icar...82..288D. doi:10.1016/0019-1035(89)90040-7. hdl:2027.42/27655.
  • de Pater, Imke; Romani, Paul N.; Atreya, Sushil K. (June 1991). "Possible microwave absorption by H2S gas in Uranus' and Neptune's atmospheres" (PDF). Icarus. 91 (2): 220–233. Bibcode:1991Icar...91..220D. doi:10.1016/0019-1035(91)90020-T. hdl:2027.42/29299.
  • Pearl, J. C.; Conrath, B. J.; Hanel, R. A.; Pirraglia, J. A.; Coustenis, A. (March 1990). "The albedo, effective temperature, and energy balance of Uranus, as determined from Voyager IRIS data". Icarus. 84 (1): 12–28. Bibcode:1990Icar...84...12P. doi:10.1016/0019-1035(90)90155-3.
  • Pollack, James B.; Rages, Kathy; Pope, Shelly K.; Tomasko, Martin G.; Romani, Paul N.; Atreya, Sushil K. (December 30, 1987). "Nature of the Stratospheric Haze on Uranus: Evidence for Condensed Hydrocarbons" (PDF). Journal of Geophysical Research. 92 (A13): 15, 037–15, 065. Bibcode:1987JGR....9215037P. doi:10.1029/JA092iA13p15037.
  • Smith, B. A. (October 1984). "Near infrared imaging of Uranus and Neptune". In JPL Uranus and Neptune. 2330: 213–223. Bibcode:1984NASCP2330..213S.
  • Smith, B. A.; Soderblom, L. A.; Beebe, A.; Bliss, D.; Boyce, J. M.; Brahic, A.; Briggs, G. A.; Brown, R. H.; Collins, S. A. (4 July 1986). "Voyager 2 in the Uranian System: Imaging Science Results". Science. 233 (4759): 43–64. Bibcode:1986Sci...233...43S. doi:10.1126/science.233.4759.43. PMID 17812889. S2CID 5895824.
  • Sromovsky, L. A.; Fry, P. M. (December 2005). "Dynamics of cloud features on Uranus". Icarus. 179 (2): 459–484. arXiv:1503.03714. Bibcode:2005Icar..179..459S. doi:10.1016/j.icarus.2005.07.022.
  • Sromovsky, L. A.; Irwin, P. G. J.; Fry, P. M. (June 2006). "Near-IR methane absorption in outer planet atmospheres: Improved models of temperature dependence and implications for Uranus cloud structure". Icarus. 182 (2): 577–593. Bibcode:2006Icar..182..577S. doi:10.1016/j.icarus.2006.01.008.
  • Sromovsky, L. A.; Fry, P. M.; Hammel, H. B.; Ahue, W. M.; de Pater, I.; Rages, K. A.; Showalter, M. R.; van Dam, M. A. (September 2009). "Uranus at equinox: Cloud morphology and dynamics". Icarus. 203 (1): 265–286. arXiv:1503.01957. Bibcode:2009Icar..203..265S. doi:10.1016/j.icarus.2009.04.015. S2CID 119107838.
  • Summers, M. E.; Strobel, D. F. (November 1, 1989). "Photochemistry of the atmosphere of Uranus". The Astrophysical Journal. 346: 495–508. Bibcode:1989ApJ...346..495S. doi:10.1086/168031.
  • Stone, E. C. (December 30, 1987). "The Voyager 2 Encounter with Uranus" (PDF). Journal of Geophysical Research. 92 (A13): 14, 873–14, 876. Bibcode:1987JGR....9214873S. doi:10.1029/JA092iA13p14873.
  • Trafton, L. M.; Miller, S.; Geballe, T. R.; Tennyson, J.; Ballester, G. E. (October 1999). "H2 Quadrupole and H3+ Emission from Uranus: The Uranian Thermosphere, Ionosphere, and Aurora". The Astrophysical Journal. 524 (2): 1, 059–1, 083. Bibcode:1999ApJ...524.1059T. doi:10.1086/307838.
  • Tyler, G. L.; Sweetnam, D. N.; Anderson, J. D.; Campbell, J. K.; Eshleman, V. R.; Hinson, D. P.; Levy, G. S.; Lindal, G. F.; Marouf, E. A.; Simpson, R. A. (4 July 1986). "Voyager 2 Radio Science Observations of the Uranian System: Atmosphere, Rings, and Satellites". Science. 233 (4759): 79–84. Bibcode:1986Sci...233...79T. doi:10.1126/science.233.4759.79. PMID 17812893. S2CID 1374796.
  • Young, L. (2001). "Uranus after Solstice: Results from the 1998 November 6 Occultation" (PDF). Icarus. 153 (2): 236–247. Bibcode:2001Icar..153..236Y. doi:10.1006/icar.2001.6698.

External links edit

  Media related to Uranus (atmosphere) at Wikimedia Commons