Jump to content

Cloud: Difference between revisions

From Wikipedia, the free encyclopedia
[pending revision][pending revision]
Content deleted Content added
m Reverted edits by ROBLOcasno (talk) (HG 3)
Marabruma (talk | contribs)
added section concerning cloud properties (relevant parameters used in research)
Line 425: Line 425:
=== Global brightening ===
=== Global brightening ===
New research indicates a [[Global dimming#Recent reversal of the trend|global brightening]] trend.<ref>{{cite journal|author=Martin Wild, Hans Gilgen, Andreas Roesch, Atsumu Ohmura, Charles N. Long, Ellsworth G. Dutton, Bruce Forgan, Ain Kallis, Viivi Russak, and Anatoly Tsvetkov|doi=10.1126/science.1103215|title=From Dimming to Brightening: Decadal Changes in Solar Radiation at Earth's Surface|year=2005|journal=Science|volume=308|issue=5723|pages=847–50|pmid=15879214 |bibcode = 2005Sci...308..847W }}</ref> The details are not fully understood, but much of the [[global dimming]] (and subsequent reversal) is thought to be a consequence of changes in [[aerosol]] loading in the atmosphere, especially sulfur-based aerosol associated with biomass burning and urban pollution.<ref>{{cite journal|author=Costantino, L. and F.-M. Bréon |doi=10.1029/2009GL041828|title=Analysis of aerosol-cloud interaction from multi-sensor satellite observations|year=2010|journal=Geophysical Research Letters|volume=37|issue=11|pages=n/a|bibcode = 2010GeoRL..3711801C }}</ref> Changes in aerosol burden can have indirect effects on clouds by changing the droplet size distribution<ref>{{cite journal|author=S. A. Twomey|doi=10.1016/0004-6981(74)90004-3|title=Pollution and the planetary albedo|year=1974|journal=Atmospheric Environment (1967)|volume=8|issue=12|pages=1251}}</ref> or the lifetime and precipitation characteristics of clouds.<ref>{{cite journal|author=B. Stevens and G. Feingold|doi=10.1038/nature08281|title=Untangling aerosol effects on clouds and precipitation in a buffered system|year=2009|journal=Nature|volume=461|issue=7264|pages=607–13|pmid=19794487|bibcode = 2009Natur.461..607S }}</ref>
New research indicates a [[Global dimming#Recent reversal of the trend|global brightening]] trend.<ref>{{cite journal|author=Martin Wild, Hans Gilgen, Andreas Roesch, Atsumu Ohmura, Charles N. Long, Ellsworth G. Dutton, Bruce Forgan, Ain Kallis, Viivi Russak, and Anatoly Tsvetkov|doi=10.1126/science.1103215|title=From Dimming to Brightening: Decadal Changes in Solar Radiation at Earth's Surface|year=2005|journal=Science|volume=308|issue=5723|pages=847–50|pmid=15879214 |bibcode = 2005Sci...308..847W }}</ref> The details are not fully understood, but much of the [[global dimming]] (and subsequent reversal) is thought to be a consequence of changes in [[aerosol]] loading in the atmosphere, especially sulfur-based aerosol associated with biomass burning and urban pollution.<ref>{{cite journal|author=Costantino, L. and F.-M. Bréon |doi=10.1029/2009GL041828|title=Analysis of aerosol-cloud interaction from multi-sensor satellite observations|year=2010|journal=Geophysical Research Letters|volume=37|issue=11|pages=n/a|bibcode = 2010GeoRL..3711801C }}</ref> Changes in aerosol burden can have indirect effects on clouds by changing the droplet size distribution<ref>{{cite journal|author=S. A. Twomey|doi=10.1016/0004-6981(74)90004-3|title=Pollution and the planetary albedo|year=1974|journal=Atmospheric Environment (1967)|volume=8|issue=12|pages=1251}}</ref> or the lifetime and precipitation characteristics of clouds.<ref>{{cite journal|author=B. Stevens and G. Feingold|doi=10.1038/nature08281|title=Untangling aerosol effects on clouds and precipitation in a buffered system|year=2009|journal=Nature|volume=461|issue=7264|pages=607–13|pmid=19794487|bibcode = 2009Natur.461..607S }}</ref>

== Cloud properties ==

=== Detection ===
Data sets regarding cloud properties are gathered using satellites, such as [[Moderate-Resolution_Imaging_Spectroradiometer | MODIS]], [[POLDER]], [[CALIPSO]] or [[European_Remote-Sensing_Satellite | ATSR]] The instruments measure the [[Radiance | radiances]] of the clouds, from which the relevant parameters can be retrieved. This is usually done by using [[Inverse problem | inverse theory]]. <ref name = "2013stubenrauch">{{cite journal

| last1 = Stubenrauch | first1 = C. J.
| last2 = Rossow | first2 = W. B.
| last3 = Kinne | first3 = S.
| last4 = Ackerman | first4 = S.
| last5 = Cesana | first5 = G.
| last6 = Chepfer | first6 = H
| last7 = Di Girolamo | first7 = L.
| last8 = Getzewich | first8 = B.
| last9 = Guignard | first9 = A.
| last10 = Heidinger | first10 = A.
| last11 = Maddux | first11 = B. C.
| last12 = Menzel | first12 = W.P
| last13 = Minnis | first13 = P.
| last14 = Pearl | first14 = C.
| last15 = Platnick | first15 = S.
| last16 = Poulsen | first16 = C.
| last17 = Reidi | first17 = J.
| last18 = Sun-Mack | first18 = S
| last19 = Walther | first19 = A.
| last20 = Winker | first20 = D.
| last21 = Zeng | first21 = S.
| last22 = Zhao | first22 = G.
| title = Assessment of global cloud datasets from satellites: Project and Database initiated by GEWEX Radiation Panel
| journal = Bulletin of the American Meteorological Society
| date = 2013
| volume = 94 | pages = 1031-1049
| doi = 10.1175/BAMS-D-12-00117.1
| url=http://dx.doi.org/10.1175/BAMS-D-12-00117.1
| format = pdf

}}</ref>.

The method of detection is based on the fact that the clouds tend to appear brighter and colder than the land surface. Because of this, difficulties rise in detecting clouds above bright (highly [[Reflection_(physics) | reflective]]) surfaces, such as oceans and ice<ref name = "2013stubenrauch"></ref>.

=== Parameters ===

The value of a certain parameter is more reliable the more satellites are measuring the said parameter. This is due to the fact that the range of errors and neglected details varies from instrument to instrument. Thus, if the analysed parameter has similar values for different instruments, it is accepted that the true value lies in the range given by the corresponding data sets.

The [[Global Energy and Water Cycle Experiment]] uses the following quantities in order to compare data quality from different satellites in order to establish a reliable quantification of the properties of the clouds<ref name = "2013stubenrauch"></ref>:

*the '''[[Cloud_cover | Cloud Amount]]''' with values between 0 and 1
*the '''cloud temperature at [[Cloud top | top]]''' ranging from 150 to 340 K
*the '''cloud pressure at top''' 1013 - 100 [[Hecto- | hPa]]
*the '''cloud height''', measured above sea level, ranging from 0 to 20 km
*the '''cloud IR emissivity''', with values between 0 and 1, with a global average around 0.7.
*the '''effective cloud amount''' is the cloud amount weighted by the cloud IR emissivity, with a global average of 0.5
*the '''cloud (visible) [[optical depth]])''' varies within a range of 4 and 10.
*the '''cloud water path''' for the liquid and solid (ice) phases of the cloud particles
*the '''cloud effective particle size''' for both liquid and ice, ranging from 0 to 200 μm

=== Vertical Distribution ===
Data from satellites suggests the following vertical distribution of the cloud amount<ref name = "2013stubenrauch"></ref>:
*High-level clouds, 40%-50%
*Mid-level clouds, around 10%-20%
*low-level clouds, around 40%


== Extraterrestrial ==
== Extraterrestrial ==

Revision as of 11:42, 5 August 2014

Stratocumulus stratiformis cumulogenitus

In meteorology, a cloud is a visible mass of liquid droplets or frozen crystals made of water or various chemicals suspended in the atmosphere above the surface of a planetary body.[1] These suspended particles are also known as aerosols and are studied in the cloud physics branch of meteorology.

Terrestrial cloud formation is the result of air in Earth's atmosphere becoming saturated due to either or both of two processes; cooling of the air and adding water vapor. With sufficient saturation, precipitation will fall to the surface; an exception is virga, which evaporates before reaching the surface.

Clouds in the troposphere, the atmospheric layer closest to Earth's surface, have Latin names due to the universal adaptation of Luke Howard's nomenclature. It was introduced in December 1802 and became the basis of a modern international system that classifies these tropospheric aerosols into several physical forms, then cross-classifies them as low-, middle- and high-étage according to cloud-base altitude range above Earth's surface. Clouds with significant vertical extent occupying more than one étage are often considered a distinct group or sub-group. One physical form shows free-convective upward growth into low or vertical heaps of cumulus. Other forms appear as non-convective layered sheets like low stratus, and as limited-convective rolls or ripples as with stratocumulus. Both of these layered forms have middle- and high-étage variants identified respectively by the prefixes alto- and cirro-. Thin fibrous wisps of cirrus are a physical form found only at high altitudes of the tropopshere. In the case of clouds with vertical extent, prefixes are used whenever necessary to express variations or complexities in their physical structures. These include cumulo- for complex highly convective vertical nimbus storm clouds, and nimbo- for thick stratiform layers with sufficient vertical depth to produce moderate to heavy precipitation. This process of cross-classification produces ten basic genus-types or genera, most of which can be subdivided into species and varieties. Synoptic surface weather observations use code numbers to record and report any type of tropospheric cloud visible at scheduled observation times based on its height and physical appearance.

While a majority of clouds form in Earth's troposphere, there are occasions when they can be observed at much higher altitudes in the stratosphere and mesosphere. Clouds that form above the troposphere have common names for their main types, but are sub-classified alpha-numerically rather than with the elaborate system of Latin names given to cloud types in the troposphere. These three main atmospheric layers that can produce clouds, along with the lowest part of the cloudless thermosphere, are collectively known as the homosphere. Above this lies the heterosphere (which includes the rest of the thermosphere and the exosphere) that marks the transition to outer space. Clouds have been observed on other planets and moons within the Solar System, but, due to their different temperature characteristics, they are often composed of other substances such as methane, ammonia, and sulfuric acid as well as water.

Tropospheric class

Formation: how the air becomes saturated

Cooling air to its dew point

Cloud evolution in under a minute.
Late-summer rainstorm in Denmark. Nearly black color of base indicates main cloud in foreground probably cumulonimbus.
Adiabatic cooling

All weather-related clouds form in the troposphere, the lowest layer of Earth's atmosphere. This generally happens when one or more lifting agents causes air containing invisible water vapor to rise and cool to its dew point, the temperature at which the air becomes saturated. The main mechanism behind this process is adiabatic cooling.[2] Atmospheric pressure decreases with altitude, so the rising air expands in a process that expends energy and causes the air to cool, which reduces its capacity to hold water vapor. If the air is cooled to its dew point and becomes saturated, it normally sheds vapor it can no longer retain which condenses into cloud.[3]

Lifted condensation level

The altitude at which this begins to happen is called the lifted condensation level, which roughly determines the height of the cloud base. Water vapor in saturated air is normally attracted to condensation nuclei such as salt particles that are small enough to be held aloft by normal circulation of the air. If the condensation process occurs below the freezing level in the troposphere, the nuclei help transform the vapor into very small water droplets. Clouds that form just above the freezing level are composed mostly of supercooled liquid droplets, while those that condense out at higher altitudes where the air is much colder generally take the form of ice crystals. An absence of sufficient condensation particles at and above the condensation level causes the rising air to become supersaturated and the formation of cloud tends to be inhibited.[3]

Frontal and cyclonic lift

There are three main agents of vertical lift. One comprises two closely related processes which work together. Frontal lift and cyclonic lift occur when stable or slightly unstable air, which has been subjected to little or no surface heating, is forced aloft at weather fronts and around centers of low pressure.[3] Cloud droplets form when the air is lifted beyond the condensation level where water vapor condenses on so-called nuclei; (small particles) that grow to a size of typically 0.002 mm (.00008 in).[4] In a cloud the droplets collide to form larger droplets. These larger droplets remain aloft as long as the drag force of the air dominates over the gravitational force for small particles. If the cloud droplets continue to grow past this size, they become too heavy to be held aloft as the gravitational force overcomes the atmospheric drag, and they fall from the cloud as rain.[5] When this process takes place just above the freezing level, the vapor tends to condense into supercooled water droplets, which with additional lifting and growth in size, can eventually turn into freezing rain. At temperatures well below freezing, the vapor desublimates into ice crystals that average about 0.25 mm in length.[6] Continuing lift and desublimation will tend to increase the number of ice crystals which may combine until they are too heavy to be supported by the vertical air currents and fall out as snow.

Convective lift

Another agent is the buoyant convective upward motion caused by significant daytime solar heating at surface level, or by relatively high absolute humidity.[3] Air warmed in this way becomes increasingly unstable. This causes it to rise and cool until temperature equilibrium is achieved with the surrounding air aloft. If air near the surface becomes extremely warm and unstable, its upward motion can become quite explosive resulting in towering clouds that can break through the tropopause or cause severe weather. Strong convection upcurrents may allow the droplets to grow to nearly .08 mm (.003 in) before precipitating as heavy rain from an active thundercloud. More occasionally, very warm unstable air is present around fronts and low-pressure centers. As with non-frontal convective lift, increasing instability promotes upward vertical cloud growth and raises the potential for severe weather.

Orographic lift

A third source of lift is wind circulation forcing air over a physical barrier such as a mountain (orographic lift).[3] If the air is generally stable, nothing more than lenticular cap clouds will form. However, if the air becomes sufficiently moist and unstable, orographic showers or thunderstorms may appear.[7]

Windy evening twilight enhanced by the Sun's angle, can visually mimic a tornado resulting from orographic lift
Non-adiabatic cooling

Along with adiabatic cooling that requires a lifting agent, there are three other main mechanisms for lowering the temperature of the air to its dew point, all of which occur near surface level and do not require any lifting of the air. Conductive, radiational, and evaporative cooling can cause condensation at surface level resulting in the formation of fog.[8] Conductive cooling takes place when air from a relatively mild source area comes into contact with a colder surface, as when mild marine air moves across a colder land area. Radiational cooling occurs due to the emission of infrared radiation, either by the air or by the surface underneath.[9] This type of cooling is common during the night when the sky is clear. Evaporative cooling happens when moisture is added to the air through evaporation, which forces the air temperature to cool to its wet-bulb temperature, or sometimes to the point of saturation.[10]

Adding moisture to the air

There are five main ways water vapor can be added to the air. Increased vapor content can result from wind convergence over water or moist ground into areas of upward motion.[11] Precipitation or virga falling from above also enhances moisture content.[12] Daytime heating causes water to evaporate from the surface of oceans, water bodies or wet land.[13] Transpiration from plants is another typical source of water vapor.[14] Lastly, cool or dry air moving over warmer water will become more humid. As with daytime heating, the addition of moisture to the air increases its heat content and instability and helps set into motion those processes that lead to the formation of cloud or fog.[15]

Cohesion and dissolution

There are forces in the atmosphere such as wind shear and downdrafts that can impact the structural integrity of a cloud. However, as long as the air remains saturated, the natural force of cohesion that hold the molecules of a substance together acts to keep the cloud from breaking up.[16] Dissolution of the cloud can occur when the process of adiabatic cooling ceases after the passage of a weather disturbance or following the loss of daytime heating of the lower troposphere. Upward lift of the air is replaced by subsidence. This leads to at least some degree of adiabatic warming of the air which can result in the cloud droplets evaporating and turning back into invisible water vapor.[17]

Distribution: variable global prevalence

Convergence

Atmospheric convergence is a process that involves the horizontal inflow and accumulation of air at a given location, as well as the rate at which this happens. This accumulation causes air to rise. If higher altitude divergence (horizontal outflow) of an equal amount occurs simultaneously above the same location, the surface atmospheric pressure is theoretically not affected. However, much of the surface convergence that occurs in the atmosphere is caused by the drawing in of air in the form of wind currents towards areas of low pressure that are the product of unequal heating of the Earth's surface.

An idealised view of three large circulation cells.
Low-pressure zones

Although the local distribution of clouds can be significantly influenced by topography, the global prevalence of cloud cover tends to vary more by latitude. This is the result of atmospheric motion driven by the uneven horizontal distribution of net incoming radiation from the sun. Cloudiness reaches maxima close to the equator and near the 50th parallels of latitude in the northern and southern hemispheres.[18] These are zones of low pressure that encircle the Earth as part of a system of large latitudinal cells that influence atmospheric circulation. In both hemispheres working away from the equator, they are the tropical Hadley cells, the mid-latitude Ferrel, and the polar cells. The 50th parallels coincide roughly with bands of low pressure situated just below the polar highs. These extratropical convergence zones are occupied by the polar fronts where air masses of polar origin meet and clash with those of tropical or subtropical origin.[19] This leads to the formation of weather-making extratropical cyclones composed of cloud systems that may be stable or unstable to varying degrees according to the stability characteristics of the various airmasses that are in conflict.[20]

February position of the ITCZ and monsoon trough in the Pacific Ocean, depicted by area of convergent streamlines offshore Australia and in the equatorial eastern Pacific
Intertropical convergence zone

Near the equator, increased cloudiness is due to the presence of the low-pressure Intertropical Convergence Zone or monsoon trough. Monsoon troughing in the western Pacific reaches its latitudinal zenith in each hemisphere above and below the equator during the late summer when the wintertime surface high-pressure ridge in the opposite hemisphere is strongest. The trough can reach as far as the 40th parallel north in East Asia during August and the 20th parallel south in Australia during February. Its poleward progression is accelerated by the onset of the summer monsoon which is characterized by the development of lower air pressure of greater instability over the warmest parts of the various continents. Cloud cover formed in this way tends to be unstable and free-convective in nature. The resulting weather systems often produce heavy showers and thunderstorms[21] These can result in the formation of tropical storms and hurrucanes composed mainly of towering thunderclouds. In the southern hemisphere, the trough associated with the Australian monsoon reaches its most southerly latitude in February, oriented along a west-northwest to east-southeast axis.[22]

Divergence

Divergence is the opposite of convergence. In the Earth's atmosphere, it involves the hoizontal outflow of air from the upper part of a rising column of air, or from the lower part of a subsiding column often associated with an area or ridge of high pressure.

Subtropical ridge

Cloudiness reaches minima near the poles and in the subtropics close to the 20th parallels, north and south. The latter are sometimes referred to as the horse latitudes. The presence of a large-scale high-pressure subtropical ridge on each side of the equator reduces cloudiness at these low latitudes. Heating of the Earth near the equator leads to large amounts of upward motion and convection along the monsoon trough or intertropical convergence zone. These rising air currents diverge in the upper troposphere and move away from the equator at high altitude in both northerly and southerly directions. As it moves towards the mid-latitudes on both sides of the equator, the air cools and sinks. The resulting air mass subsidence creates a subtropical ridge near the 30th parallel of latitude in both hemispheres where the formation of cloud is minimal. At surface level, the sinking air diverges again with some moving back to the equator and completing the vertical cycle. This circulation on each side of the equator is known as the Hadley cell in the tropics. Many of the world's deserts are caused by these climatological high-pressure areas.[23]

Polar high

Similar patterns also occur at higher latitudes in both hemispheres. Upward currents of air along the polar fronts diverge at high tropospheric altitudes. Some of the diverging air moves to the poles where air mass subsidence inhibits cloud formation and leads to the creation of the polar areas of high pressure. Divergence occurs near surface level resulting in a return of the circulating air to the polar fronts where rising air currents can create extensive cloud cover and precipitation.[19] This vertical cycle comprises the polar cell in each latitudinal hemisphere. Some of the air rising at the polar fronts diverges away from the poles and moves in the opposite direction to the high level zones of convergence and subsidence at the subtropical ridges on each side of the equator. These mid-latitude counter-circulations create the Ferrel cells that encircle the globe in the northern and southern hemispheres.

Classification

Latin nomenclature: historical background

Luke Howard and Jean-Baptiste Lamarck

Luke Howard, a methodical observer with a strong grounding in the Latin language, used his background to classify the various tropospheric cloud types during December 1802. He believed that the changing cloud forms in the sky could unlock the key to weather forecasting. Jean-Baptiste Lamarck worked independently on cloud classification and came up with a different naming scheme that failed to make an impression even in his home country of France because it used unusual French names for cloud types. His system of nomenclature included twelve categories of clouds, with such names as (translated from French) hazy clouds, dappled clouds and broom-like clouds. Howard used universally accepted Latin, which caught on quickly. As a sign of the popularity of the naming scheme, the German dramatist and poet Johann Wolfgang von Goethe composed four poems about clouds, dedicating them to Howard. Classification systems would be proposed by Heinrich Dove of Germany in 1828 and Elias Loomis of the United States in 1841, but neither became the international standard that Howard's system became. It was formally adopted by the International Meteorological Commission in 1929.[24]

First comprehensive classification

Howard's original system established three general cloud forms based on physical appearance and process of formation: cirriform (mainly detached and wispy), cumuliform or convective (mostly detached and heaped, rolled, or rippled), and non-convective stratiform (mainly continuous layers in sheets).[25] These were cross-classified into lower and upper étages. Cumuliform clouds forming in the lower level were given the genus name cumulus, and low stratiform clouds the genus name stratus. Physically similar clouds forming in the upper étage were given the genus names cirrocumulus (generally showing more limited convective activity[26] than low level cumulus) and cirrostratus, respectively. Cirriform clouds were identified as always upper level and given the genus name cirrus. To these, Howard added the genus nimbus for clouds of complex structure producing significant precipitation[27] that came to be identified as a distinct nimbiform physical category.[28]

Howard's successors

Around 1840–41, German meteorologist Ludwig Kaemtz added stratocumulus as a mostly detached low-étage genus of limited convection with both cumuliform- and stratiform characteristics similar to upper-level cirrocumulus. This had the effect of creating a stratocumuliform[29] type that included rolled and rippled clouds separately from the more freely convective heap clouds.[27] About fifteen years later, Emilien Renou, director of the Parc Saint-Maur and Montsouris observatories, began work on an elaboration of Howard's classifications that would lead to the introduction during the 1870s of altocumulus (physically more closely related to stratocumulus than to cumulus[26]) and altostratus. These were respectively stratocumuliform and stratiform cloud genera of a newly defined middle étage above stratocumulus and stratus but below cirrocumulus and cirrostratus, with free convective cumulus and non-convective nimbus occupying more than one étage as clouds with vertical extent. In 1880, Philip Weilbach, secretary and librarian at the Art Academy in Copenhagen, and like Luke Howard, an amateur meteorologist, proposed and had accepted by the permanent committee of the International Meteorological Organization (IMO), a forerunner of the present-day World Meteorological Organization (WMO), the designation of a new free-convective vertical or multi-étage genus type, cumulonimbus, which would be distinct from cumulus and nimbus and identifiable by its often very complex structure (frequently including a cirriform top and what are now recognized as multiple accessory clouds), and its ability to produce thunder. With this addition, a canon of ten cloud genera was established that came to be officially and universally accepted. At about the same time, several cloud specialists proposed variations that came to be accepted as species subdivisions and varieties determined by more specific variable aspects of the structure of each genus. A further modification of the genus classification system came when an IMC commission for the study of clouds put forward a refined and more restricted definition of the genus nimbus which was effectively reclassified as a stratiform cloud type. It was then renamed nimbostratus and published with the new name in the 1932 edition of the International Atlas of Clouds and of States of the Sky.[27] This left cumulonimbus as the only nimbiform type as indicated by its root-name. In 1976, the National Aeronautics and Space Administration (NASA) published a cloud classification that showed a change in name of the nimbiform type to cumulonimbiform,[29] although some other agencies have continued to recognize the earlier name.

Physical forms

As established by Howard and his successors, clouds are commonly grouped into physical forms that can be up to five in number: cirriform, cumuliform, cumulonimbiform, stratocumuliform, and stratiform. These designations distinguish a cloud's physical structure and process of formation.

Cirriform

Genus cirrus (high)

Cirriform clouds generally have a wispy fibrous appearance and form at high tropospheric altitudes along the very leading edges of a frontal or low-pressure weather disturbance and often along the fringes of its other borders. In general, they are non-convective[26] but occasionally acquire a tufted or turreted appearance caused by small-scale high-altitude convection. These high clouds do not produce precipitation as such[30] but are often accompanied or followed by lower-based clouds that do.

Cumuliform

Genus cumulus (low or vertical/multi-level)

Cumuliform clouds typically have flat bases and puffy domed tops.[31][32] They are the product of localized but generally free-convective lift[26] and can vary in vertical extent depending on the stability characteristics of the air mass where they are forming.[33] The smallest fair weather cumuliform types occur with only minimal instability and can therefore be considered clouds of limited convection.[3] Incoming short-wave radiation generated by the sun is re-emitted as long-wave radiation when it reaches Earth's surface. This process warms the air closest to ground and increases air mass instability by creating a steeper temperature gradient from warm or hot at surface level to cold aloft. Moderate instability allows for the formation of cumuliform clouds of moderate size that can produce light showers if the airmass is sufficiently moist. The more the air is heated from below, the more unstable it tends to become. This may cause large towering cumuliform clouds to form in the lower half of the troposphere with tops growing into the upper levels. These buildups can cause moderate to occasionally heavy showers. They tend to be more concentrated and intense when they are associated with fast-moving unstable cold fronts.

Cumulonimbiform

Genus cumulonimbus (vertical/multi-level)

The largest free-convective[26] cumuliform clouds occur in very unstable air[3] and often have complex structures that include cirriform tops and multiple accessory clouds and are sometimes classified separately as cumulonimbiform. At maturity, they have very strong updrafts that can penetrate the tropopause. They can produce thunderstorms and a variety of types of lightning including cloud-to-ground that can cause wildfires.[34] Other convective severe weather may or may not be associated with thunderstorms and include heavy rain or snow showers, hail,[35] strong wind shear, downbursts,[36] and tornadoes.[37]

Stratiform

Genera cirrostratus (high), altostratus (middle), stratus (low), nimbostratus (vertical/multi-level)

In general, stratiform clouds have a flat sheet-like structure and form at any altitude in the troposphere where there is sufficient condensation as the result of non-convective[26] lift of relatively stable air, especially along warm fronts, around areas of low pressure, and sometimes along stable slow moving cold fronts.[3] In general, precipitation falls from stratiform clouds in the lower half of the troposphere. If the weather system is well-organized, the precipitation is generally steady and widespread. The intensity varies from light to heavy according to the thickness of the stratiform layer as determined by moisture content of the air and the intensity of the weather system creating the clouds and weather. Unlike free convective cumuliform and cumulonimbiform clouds that tend to grow upward, stratiform clouds achieve their greatest thickness when precipitation that forms in the middle étage of the troposphere triggers downward growth of the cloud base to near surface level. Stratiform clouds can also form in precipitation below the main frontal cloud deck where the colder air is trapped under the warmer air mass being forced above by the front. Non-frontal low stratiform cloud can form when advection fog is lifted above surface level during breezy conditions.

Stratocumuliform

Cirrocumulus (high), altocumulus (middle), stratocumulus (low)

Clouds of this physical structure have both cumuliform and stratiform characteristics and generally form as a result of limited convection[26] in slightly unstable air. They can form at any altitude in the troposphere wherever and whenever there is sufficient moisture and lift. High stratocumuliform clouds also tend show some cirriform characteristics or form in association with cirriform clouds. If a poorly organized low-pressure weather system is present, virga or weak intermittent precipitation may fall from those stratocumuliform clouds that form mostly in the low and lower-middle étages of the troposphere.

Etages and cross-classification into genera

Cloud classification by altitude of occurrence. Towering vertical cumulus congestus not shown.

The individual genus types result from the physical forms being cross-classified by étage within the troposphere. A general consensus exists as to the designation of high, middle, and low étages, the makeup of the basic canon of ten cloud genera that results from this cross-classification, and the étage designations of non-vertical genus types. Several but not all methods of altitude classification treat clouds with significant vertical extent as a separate group or sub-group.[38][39][40] These very thick clouds can spread vertically into more than one étage and therefore have multiple affiliations depending on the criteria used by various authorities, and can have low or middle-étage bases depending on the moisture content of the air. The base-height range for each étage varies depending on the latitudinal geographical zone.[41]

High

Clouds of the high-étage form at altitudes of 3,000 to 7,600 m (10,000 to 25,000 ft) in the polar regions, 5,000 to 12,200 m (16,500 to 40,000 ft) in the temperate regions and 6,100 to 18,300 m (20,000 to 60,000 ft) in the tropical region.[41] All cirriform clouds are classified as high and thus constitute a single genus cirrus (Ci). Stratocumuliform and stratiform clouds in the high-étage carry the prefix cirro-, yielding the respective genus names cirrocumulus (Cc) and cirrostratus (Cs). Strato- is excluded from cirrocumulus to avoid double prefixing. Most high cloud forms as a result of natural atmospheric processes. However, contrails formed from the exhaust of high-flying aircraft can persist and spread into formations resembling cirrus, cirrocumulus, or cirrostratus. This variant has no special WMO designation, but is sometimes given the faux-Latin name Aviaticus.

Middle

The middle-étage typically comprises one stratocumuliform and one stratiform genus. They are prefixed by alto-, yielding the genus names altocumulus (Ac) and altostratus (As). Strato- is also excluded from altocumulus. These clouds can form as low as 2,000 m (6,500 ft) above surface at any latitude, but may be based as high as 4,000 m (13,000 ft) near the poles, 7,000 m (23,000 ft) at mid latitudes, and 7,600 m (25,000 ft) in the tropics.[41]

Low

Low-étage clouds are found from near surface up to 2,000 m (6,500 ft).[41] This group mainly includes one stratocumuliform and one stratiform genus whenever vertical or multi-étage clouds are classified separately.[38] When a low stratiform cloud contacts the ground, it is called fog, although radiation and advection types of fog do not form from stratus layers. Genus types in this étage either have no prefix or carry one that refers to a characteristic other than altitude. Of the two main low cloud types, the prefixed genus is stratocumulus (Sc), a low-étage cloud of limited convection, and the non-prefixed genus is non-convective stratus (St) that usually forms into a comparatively thin layer. Small fair weather cumulus (Cu) of limited convection is also often included with this family due to its lack of vertical development compared to moderate and towering cumulus types.

Vertical or multi-level

Upward-growing free-convective clouds have low to middle-étage bases that form anywhere from near surface to about 2,400 m (8,000 ft) in temperate climates, and often much higher in arid regions, even to the very top of the middle altitude range of the troposphere. This group, when recognized as such, includes the singular cumuliform and cumulonimbiform genus types, and one very thick stratiform genus. The first of these is free-convective cumulus (Cu) that carries no prefix. It usually forms in the low étage except during conditions of very low relative humidity when the clouds bases can rise into the middle altitude range. The other two types have non height-related prefixes. Cumulonimbus (Cb) is prefixed according to its free-convective characteristics. Nimbostratus (Ns) is a deep non-convective stratiform genus with great horizontal extent. It normally forms from middle-étage altostratus and achieves vertical extent[38][39] as it thickens during precipitation with the base subsiding into the low altitude range.[42] The nimbo- prefix refers to its ability to produce continuous rain or snow over a wide area, aspecially ahead of a warm front.[43]

Some methods of height classification limit the term vertical to upward-growing free-convective cumuliform and cumulonimbiform genera whose vertical thickness exceeds their horizontal base-width.[40][44] Downward-growing nimbostratus can be as thick as most upward-growing vertical cumulus, but its horizontal extent tends to be even greater. This sometimes leads to the exclusion of this genus type from the group of vertical clouds. Authorities who follow this approach usually classify nimbostratus either as low-étage[40][45] to denote its normal base height range, or as middle,[41] based on the altitude range at which it normally forms. Sometimes the terms multi-level or multi-étage are used for all very thick or tall cloud types including nimbostratus to avoid the connotation of 'vertical' with free-convective cumuliform only.[39] Alternatively, some classifications do not recognize a vertical or multi-étage designation and include all vertical free-convective cumuliform and cumulonimbiform types with the low-étage clouds.[30]

Nimbostratus and some cumulus in this family usually only achieve comparatively moderate vertical extent. However, with sufficient airmass instability, upward-growing cumuliform clouds can grow to towering proportions. Although genus types with vertical extent are often considered a single group,[38] the International Civil Aviation Organization (ICAO) further distinguishes towering vertical clouds as a separate group or sub-group by specifying that these very large cumuliform and cumulonimbiform types must be identified by their standard names or abbreviations in all aviation observations (METARS) and forecasts (TAFS) to warn pilots of possible severe weather and turbulence.[46] When towering vertical types are considered separately, they comprise the aforementioned cumulonimbus genus and one cumulus species, cumulus congestus (Cu con). The latter is a sub-type of the genus cumulus. This species is designated towering cumulus (Tcu) by ICAO. There is no stratiform type in this group because by definition, even very thick stratiform clouds cannot have towering vertical structure, although they may be accompanied by embedded towering cumuliform or cumulonimbiform types.[39][47]

Species

Genus types are divided into species that indicate specific structural details. However, because these latter types are not always restricted by étage, some species can be common to several genera that are differentiated mainly by altitude.[48]

Mostly stable stratocumuliform
Altocumulus lenticularis forming over mountains in Wyoming with lower layer of cumulus mediocris and higher layer of cirrus spissatus.

Good examples of species common to more than one genus are the stratiformis and lenticularis types, each of which is common to mostly stable stratocumuliform genera in the high, middle, and low étages. (cirrocumulus, altocumulus, and stratocumulus, respectively). Stratiformis species normally occur in extensive sheets or in smaller patches where there is only minimal convective activity. Lenticularis species tend to have lens-like shapes tapered at the ends. They are most commonly seen as orographic mountain-wave clouds, but can occur anywhere in the troposphere where there is strong wind shear combined with sufficient airmass stability to maintain a generally flat cloud structure.[38][48]

Stable cirriform and stratiform

Cirrus clouds have a couple of species that are unique to the wispy structures of this genus and an additional species which is also seen with high-étage stratiform clouds. Uncinus filaments with upturned hooks and spissatus filaments that merge into dense patches are both considered cirriform species. However the species fibratus can be seen with cirrus and with cirrostratus that is transitional to or from cirrus. Cirrostratus at its most characteristic tends to be mostly of the stratiform species nebulosus, which creates a rather diffuse appearance lacking in structural detail. Altostratus and nimbostratus clouds always have this physical appearance without significant variation or deviation and, therefore, do not need to be subdivided into species. Low-étage stratus is also of the species nebulosus except when broken up into ragged sheets of stratus fractus.[30][38][48]

Partly or mostly unstable cirriform and stratocumuliform

With increasing airmass instability, castellanus structures, which resemble the turrets of a castle when viewed from the side, can be found with any stratocumuliform genus. This species is also sometimes seen with convective patches of cirrus, as are the more detached tufted floccus species, which are common to cirrus, cirrocumulus, and altocumulus, but not stratocumulus.[48]

Mostly unstable cumuliform and cumulonimbiform

With the exception of stratocumulus castellanus, local airmass instability in the low étage tends to produce clouds of the more freely convective cumulus and cumulonimbus genera, whose species are mainly indicators of degrees of vertical development. A cumulus cloud initially forms as a cloudlet of the species fractus or humilis that shows only slight vertical development. If the air becomes more unstable, the cloud tends to grow vertically into the species mediocris, then congestus, the tallest cumulus species. With further instability, the cloud may continue to grow into cumulonimbus calvus (essentially a very tall congestus cloud that produces thunder), then ultimately capillatus when supercooled water droplets at the top turn into ice crystals giving it a cirriform appearance.[48]

Varieties

Genus and species types are further subdivided into varieties whose names can appear after the species name to provide a fuller description of a cloud. Some cloud varieties are not restricted to a specific étage or physical structure, and can therefore be common to more than one genus or species.[30][48]

Opacity-based

All cloud varieties fall into one of two main groups. One group identifies the opacities of particular low and middle étage cloud structures and comprises the varieties translucidus (translucent), perlucidus (opaque with translucent breaks), and opacus (opaque). These varieties are always identifiable for cloud genera and species with variable opacity. All three are associated with the stratiformis species of altocumulus and stratocumulus. However, only two are seen with altostratus and stratus nebulosus whose uniform structures prevent the formation of a perlucidus variety. Opacity-based varieties are not applied to high-étage clouds because they are always translucent, or in the case of cirrus spissatus, always opaque.[49] Similarly, these varieties are also not associated with moderate and towering vertical clouds because they are always opaque.

Pattern-based
Sunrise scene giving a shine to an altocumulus stratiformis perlucidus cloud.

A second group describes the occasional arrangements of cloud structures into particular patterns that are discernable by a surface-based observer (cloud fields usually being visible only from a significant altitude above the formations). These varieties are not always present with the genera and species with which they are otherwise associated, but only appear when atmospheric conditions favor their formation. Intortus and vertebratus varieties occur on occasion with cirrus fibratus. They are respectively filaments twisted into irregular shapes, and those that are arranged in fishbone patterns, usually by uneven wind currents that favor the formation of these varieties. The variety radiatus is associated with cloud rows of a particular type that appear to converge at the horizon. It is sometimes seen with the fibratus and uncinus species of cirrus, the stratiformis species of altocumulus and stratocumulus, all species of cumulus,[50] and with the genus altostratus. Another variety, duplicatus (closely spaced layers of the same type, one above the other), is sometimes found with cirrus of both the fibratus and uncinus species, and with altocumulus and stratocumulus of the species stratiformis and lenticularis. The variety undulatus (having a wavy undulating base) can occur with any clouds of the species stratiformis or lenticularis, and with altostratus. It is only rarely observed with stratus nebulosus. Under conditions of strong atmospheric wind-shear and instability, this wave-like formation may break into regularly spaced crests. This variant has no separate WMO Latin designation, but is sometimes known informally as Kelvin-Helmholtz. The variety lacunosus is caused by localized downdrafts that create circular holes in the form of a honeycomb or net. It is occasionally seen with cirrocumulus and altocumulus of the species stratiformis, castellanus, and floccus, and with stratocumulus of the species stratiformis and castellanus.[30][49]

Combinations

It is possible for some species to show combined varieties at one time, especially if one variety is opacity-based and the other is pattern-based. An example of this would be an opaque layer of altocumulus stratiformis arranged in seemingly converging rows. The full technical name of a cloud in this configuration would be altocumulus stratiformis opacus radiatus, which would identify respectively its genus, species, and two combined varieties.[48][49]

Accessory clouds and other supplementary features

Cumulus and stratocumulus made orange by the sun rising

Supplementary features are not further subdivisions of cloud types below the species and variety level. Rather, they are either hydrometeors or special cloud formations with their own Latin names that form in association with certain cloud genera, species, and varieties.[48][49]

Precipitation-based

One group of supplementary features are not actual cloud formations but rather precipitation that falls when water droplets that make up visible clouds have grown too heavy to remain aloft. Virga is a feature seen with clouds producing precipitation that evaporates before reaching the ground, these being of the genera cirrocumulus, altocumulus, altostratus, nimbostratus, stratocumulus, cumulus, and cumulonimbus. When the precipitation reaches the ground without completely evaporating, it is designated as the feature praecipitatio.[51] This normally occurs with altostratus opacus, which can produce widespread but usually light precipitation, and with thicker clouds that show significant vertical development. Of the latter, upward-growing cumulus mediocris produces only isolated light showers, while downward growing nimbostratus is capable of heavier, more extensive precipitation. Towering vertical clouds have the greatest ability to produce intense precipitation events, but these tend to be localized unless organized along fast-moving cold fronts. Showers of moderate to heavy intensity can fall from cumulus congestus clouds. Cumulonimbus, the largest of all cloud genera, has the capacity to produce very heavy showers. Low stratus clouds usually produce only light precipitation, but this always occurs as the feature praecipitatio due to the fact this cloud genus lies too close to the ground to allow for the formation of virga.[48][49]

Cumulonimbus dissipating at dusk
Cloud-based

The heavier precipitating clouds, nimbostratus, towering cumulus (cumulus congestus), and cumulonimbus, also typically see the formation in precipitation of the pannus feature, low ragged clouds of the genera and species cumulus fractus or stratus fractus.[30][52] These formations, along with several other cloud-based supplementary features, are also known as accessory clouds.[48][49]

After the pannus types, the remaining supplementary features comprise cloud formations that are associated mainly with upward-growing cumuliform and cumulonimbiform clouds of free convection. Incus is the most type-specific supplementary feature, seen only with cumulonimbus of the species capillatus. A cumulonimbus incus cloud top is one that has spread out into a clear anvil shape as a result of rising air currents hitting the stability layer at the tropopause where the air no longer continues to get colder with increasing altitude.[53] The mamma feature forms on the bases of clouds as downward-facing bubble-like protuberances caused by localized downdrafts within the cloud. It is also sometimes called mammatus, an earlier version of the term used before a standardization of Latin nomenclature brought about by the World Meterorological Organization during the 20th century. The best-known is cumulonimbus with mammatus, but the mamma feature is also seen occasionally with cirrus, cirrocumulus, altocumulus, altostratus, and stratocumulus. Pileus is a cap cloud that can form over a cumulonimbus or large cumulus cloud,[54] whereas a velum feature is a thin horizontal sheet that sometime forms like an apron around the middle or in front of the parent cloud. An arcus feature is a roll or shelf cloud that forms along the leading edge of a squall line or thunderstorm outflow.[55] Some arcus clouds form as a consequence of interactions with specific geographical features. Perhaps the strangest geographically specific arcus cloud in the world is the Morning Glory, a rolling cylindrical cloud that appears unpredictably over the Gulf of Carpentaria in Northern Australia. Associated with a powerful "ripple" in the atmosphere, the cloud may be "surfed" in glider aircraft. A tuba feature is a cloud column that may hang from the bottom of a cumulus or cumulonimbus. A newly formed or poorly organized column might be comparatively benign, but can quickly intensify into a funnel cloud or tornado.[30][56][57]

Cumulus mediocris and congestus over Swifts Creek, Australia

Mother clouds

Cumulus partly spreading into stratocumulus cumulogenitus over the port of Piraeus in Greece

Clouds initially form in clear air or become clouds when fog rises above surface level. The genus of a newly formed cloud is determined mainly by air mass characteristics such as stability and moisture content. If these characteristics change over a period of time, the genus tends to change accordingly. When this happens, the original genus is called a mother cloud. If the mother cloud retains much of its original form after the appearance of the new genus, it is termed a genitus cloud. One example of this is stratocumulus cumulogenitus, a stratocumulus cloud formed by the partial spreading of a cumulus type when there is a loss of convective lift. If the mother cloud undergoes a complete change in genus, it is considered to be a mutatus cloud.[30] It is theoretically possible for some lengthy terminologies to emerge by combining the names of all applicable genera, species, varieties, and supplementary features to provide a complete description of an active and evolving genitus or mutatus cloud formation. As an extreme example, a flat opaque layer of altocumulus formed by the spreading of cumulus arranged in parallel bands accompanied by precipitation not reaching the ground could be termed altocumulus stratiformis opacus radiatus cumulogenitus virga.[48]

Stratocumulus fields

Stratocumulus clouds can be organized into "fields" that take on certain specially classified shapes and characteristics. In general, these fields are more discernable from high altitudes than from ground level. They can often be found in the following forms:

  • Actinoform, which resembles a leaf or a spoked wheel.
  • Closed cell, which is cloudy in the center and clear on the edges, similar to a filled honeycomb.[58]
  • Open cell, which resembles an empty honeycomb, with clouds around the edges and clear, open space in the middle.[59]

Cloud symbols used on weather office maps

Low cloud weather map symbols: Includes low and upward-growing vertical.
Middle cloud weather map symbols: Includes middle and downward growing vertical.
High cloud weather map symbols.

Selection of symbols

Weather maps plotted and analyzed at weather forecasting centers employ special symbols to denote various cloud étages, genera, species, varieties, mutations, and cloud movements that are considered important to identify conditions in the troposphere that will assist in preparing the forecasts. The cloud symbols are translated from numerical codes included with other meteorological data that make up the contents of international synoptic messages transmitted at regular intervals by professionally trained staff at major weather stations. In a couple of cases, an entire genus like cirrocumulus is represented by one cloud symbol, regardless of species, varieties, or any other considerations. In general though, the codes and their symbols are used to identify cloud types at the species level. A number of varieties and supplementary features are also deemed important enough to have their own weather map symbols. For the sake of economy, a particular genus, species, or variety may share a numerical reporting code and symbol with another similar cloud type. Sometimes, a separate symbol is used to indicate whether or not a particular genus has transformed or emerged from a mother cloud of another genus, or is increasing in amount or invading the sky (usually in the form of parallel bands in a radiatus configuration) ahead of an approaching weather disturbance.[30]

International synoptic code

The international synoptic code (or SYNOP) provides for reporting the three basic étages for tropospheric clouds, but makes no special provision for vertical or multi-level clouds that can occupy more than one étage at a particular time. Consequently, cloud genera with significant vertical development are coded as low when they form in the low étage of the troposphere and achieve vertical extent by growing upward into the middle or high étage, as is the case with cumulus and cumulonimbus. Conversely, nimbostratus is coded as middle because it usually initially forms in the middle étage of the troposphere and becomes vertically developed by growing downward into the low étage, and ofter upward in to the high étage as well.[38][39] Because of the structure of the SYNOP code, a maximum of three cloud symbols can be plotted for each reporting station that appears on the weather map; one symbol each for a low (or upward growing vertical) cloud type, a middle (or downward growing vertical) type, and one for a high cloud type.

The symbol used on the map for each of these étages at a particular observation time is for the genus, species, variety, mutation, or cloud motion that is considered most important according to criteria set out by the World Meteorological Organization (WMO). If these elements for any étage at the time of observation are deemed to be of equal importance, then the type which is predominant in amount is coded by the observer and plotted on the weather map. Although the SYNOP code has no separate formal group classification for vertical or multi-level clouds, the observer procedure for selecting numerical codes is designed to give high reporting priority to those genera or species that show significant vertical development.[30]

Clouds and weather forecasting

The identification and reporting of clouds contribute to the process of weather forecasting. Satellite pictures used in conjunction with the cloud symbols plotted on weather maps provide the forecaster with important information about conditions within the troposphere and the weather systems that form as a result.

Warm front or low-pressure area

The presence of significant high-étage cloud cover indicates an organized low-pressure disturbance or an associated warm front is about 300 km away from the point of observation. Clouds associated with warm fronts tend to be mostly high cirriform at first, changing to stratiform at progressively lower altitude levels as the front approaches. However, if cirrocumulus also appears, there is greater airmass instability arriving with the front[60] which increases the risk that thunderstorms may accompany the system.[61] When these high clouds progressively invade the sky and the barometric pressure begins to fall, precipitation associated with the disturbance is likely about 6 to 8 hours away.[62] A thickening and lowering of cirrostratus into middle-étage altostratus is a good sign the warm front or low has moved closer and precipitation may begin within less than six hours hours. A further thickening of the altostratus is often accompanied by virga and the arrival of precipitation is imminent. The cloud layer achieves significant vertical extent as it lowers and changes into nimbostratus.[42] Rain or snow begins to reach surface level at the beginning of a precipitation event that can last up to 36 hours depending and the size of the weather system and its speed of movement. As the low and the warm front pass, the nimbostratus thins out into low-étage stratus and the precipitation tapers off.

Cold front

A cold front tends to give less warning of its approach because it usually moves faster than a warm front and has a narrower band of clouds and weather. If the cold front is active enough to produce thunderstorms, anvil cirrus clouds[63] may spread ahead of the front as a warning of its approach. The other cloud types associated with a cold front depend on atmospheric conditions such as air mass stability and wind shear,[64][65] but are mostly cumuliform or stratocumuliform, with midddle-étage altocumulus giving way to low-level stratocumulus and intermittent light precipitation if there is only slight airmass instability. With significant instability, vertically developed cumulus or cumulonimbus with showers and thunderstorms will form along the front.[62]

High-pressure area

After the passage of the front, the sky usually clears as high pressure builds in behind the system, although significant amounts of cumulus or stratocumulus, often in the form of long bands called cloud streets may persist if the air mass behind the front remains humid.[66] Small and unchanging amounts of cumulus or cirrus clouds in an otherwise clear sky are usually indications of continuing fair weather as long as the barometric pressure remains comparatively high.[67]

Rainmaking bacteria

There is evidence that clouds, especially in the weather-making troposphere, contain biological ice nuclei that may play a key role in the formation of precipitation. Bioprecipitation, the concept of rain-making bacteria, was proposed by David Sands from Bozeman Campus, Montana State University, USA. Such microbes – called ice nucleators – are found in rain, snow, and hail throughout the world. These bacteria may be part of a constant feedback between terrestrial ecosystems and tropospheric clouds and may even have evolved the ability to promote rainstorms as a means of dispersal. They may rely on the rainfall to spread to new habitats, much as some plants rely on windblown pollen grains.[68][69]

Summary of étages, forms, genera, species, varieties, supplementary features, mother clouds, and associated weather

High cirriform, stratocumuliform, and stratiform

High cirrus uncinus and cirrus fibratus upper-left merging into cirrostratus fibratus with some higher cirrocumulus floccus upper right.
These are mostly fibrous wisps of delicate white cirriform ice crystal cloud that show up clearly against the blue sky. Cirrus are generally non-convective except castellanus and floccus species which show limited convection. They often form along a high altitude jetstream and at the very leading edge of a frontal or low-pressure disturbance where they may merge into cirrostratus.
  • Species: This genus is divided into five species which are grouped to form the basis of reporting cirrus in the SYNOP code.[48][49] Cirrus fibratus (Ci fib) consists of thin fibrous streaks with no tufts or hooks. Cirrus uncinus (Ci unc) is similar except that the filaments are hooked at the ends. Both species are coded CH1. Cirrus spissatus (Ci spi) consists of patchy dense high cloud. The castellanus species (Ci cas) has convective buildups that give the cloud a partly or mainly turreted appearance, especially when viewed from the side. Cirrus with a tufted appearance is designated cirrus floccus (Ci flo).[48][49] All three of these dense cirrus species are coded CH2. Although classified as high, dense cirrus, especially castellanus, can achieve some vertical extent.
  • Varieties: Certain cirrus species can sometimes be divided into pattern-based varieties. The filaments of cirrus fibratus intortus are twisted into irregular patterns. Cirrus fibratus vertebratus sees the filaments arranged in a pattern that resembles the backbone of a fish. Another pattern-based variety can be found with fibratus and uncinus species. Cirrus radiatus consists of parallel bands that appear to converge at the horizon. This pattern is often seen when the high cloud is invading the sky or increasing in amount.[48][49] It is then reported on the SYNOP observation code as CH4, or as CH5 or 6 (depending on how much of the sky is covered) if accompanied by cirrostratus. Cirrus duplicatus is observable when the fibratus or uncinus filaments are arranged in closely spaced layers, one above the other. Pattern-based varieties are not commonly associated with the species spissatus, castellanus, or floccus.[49] Opacity-based varieties are not associated with cirrus of any types because the wispy or fibrous species are always translucent while the more dense species are inherently opaque.
  • Precipitation-based supplementary features: These are not associated with cirrus clouds because they do not produce any precipitation.
  • Accessory cloud: Mamma is cloud-based supplementary feature that can be seen with cirrus spissatus cumulonimbogenitus (CH3). It appears as bubble-like downward protuberances from the cloud base and is caused by localized downdrafts in the cloud.
  • Genitus mother clouds: Apart from the aforementioned cumulonimbus mother cloud, cirrus fibratus cirrocumulogenitus or altocumulogenitus can form when cirrocumulus or very high altocumulus mother clouds lose some of their stratocumuliform structure and take on a more wispy or fibrous appearance.
  • Mutatus mother cloud: Cirrus fibratus cirrostratomutatus forms from a cirrostratus mother cloud when mostly continuous sheets of high cloud break up into more detached wispy or fibrous streaks.[48][49]
A A large field of cirrocumulus clouds in a blue sky, beginning to merge near the upper left.
A large field of cirrocumulus stratiformis
This is a pure white high-étage stratocumuliform layer of limited convection. It is composed of ice crystals or supercooled water droplets appearing as small unshaded round masses or flakes in groups or lines with ripples like sand on a beach.[70] Cc occasionally form alongside cirrus or cirrostratus clouds at the very leading edge of an active weather system. It is coded CH9 for all species.
  • Species: Cirrocumulus stratiformis (Cc str) is one of four species and appears in the form of relatively flat stratocumuliform sheets or patches. The species lenticularis (Cc len) takes its name from the lens-shaped structure of this cloud which is tapered at each end. Cirrocumulus castellanus (Cc cas) has cumuliform buildups that give the cloud a partly or mainly turreted appearance. When the cumuliform parts have more of a tufted appearance, it is given the species name floccus (Cc flo)[48][49] Cirrocumulus with castellanus buildups can show some vertical extent, but are not usually classified as vertical or multi-étage clouds.
  • Varieties: This genus type is always translucent and so has no opacity-based varieties. However, like cirrus, certain cirrocumulus species can sometimes be divided into pattern-based varieties. The undulatus variety has a wavy undulating base and is seen mostly with the stratiformis and lenticularis species types. The lacunosus variety contains circular holes caused by downdrafts in the cloud and is associated mainly with the species stratiformis, castellanus and floccus.[49]
  • Precipitation-based supplementary feature: Cirrocumlus occasionally produces virga, precipitation that evaporates before reaching the ground..
  • Accessory cloud: Mamma in the form of downward forming bubbles is infrequently seen as a cloud-based supplementary feature.
  • Mother clouds: This genus type has no recognized genitus mother clouds. However cirrocumulus stratiformis cirromutatus or cirrostratomutatus can result from sheets or filaments of high cloud taking on a stratocumuliform structure as a result of high altitude convection. A high layer of white or light grey altocumulus of a particular species can thin out into pure white cirrocumulus altocumulomutatus of the same species.[48][49]
Cirrostratus is a thin non-convective stratiform ice crystal veil that typically gives rise to halos caused by refraction of the sun's rays. The sun and moon are visible in clear outline. Cirrostratus often thickens into altostratus ahead of a warm front or low-pressure area.
  • Species: Cirrostratus fibratus (Cs fib) is a high fibrous sheet similar to cirrus but with less detached semi-merged filaments.[48][49] It is reported in the SYNOP code as CH8 or as CH5 or 6 (depending on the amount of sky covered) if increasing in amount. If the high cloud covers the entire sky and takes on the form of a featureless veil, it is classified as cirrostratus of the species nebulosus (Cs neb) and is coded CH7.
  • Varieties: Cirrostratus species have no opacity-based varieties as they are always translucent. Two pattern-based varieties are sometimes seen with the species fibratus. These are the closely spaced duplicatus and wavy undulatus types similar to those seen with cirrus fibratus. Pattern-based varieties are not commonly associated with the species nebulosus due to its lack of features.[49]
  • Supplementary features: Cirrostratus produces no precipitation or virga, and is not accompanied by any accessory clouds.
  • Genitus mother clouds: Cirrostratus fibratus cirrocumulogenitus sometimes appears as the latter cloud flattens and loses some of its stratocumuliform structure. Cirrostratus fibratus cumulonimbogenitus may form if the cirriform top of a mature thundercloud spreads and flattens sufficiently to become a high stratiform cloud.[48][49]
  • Mutatus mother clouds: Cirrostratus fibratus cirromutatus or cirrocumulomutatus are the result of a complete transformation from cirrus and cirrocumulus genus types. Cirrostratus nebulosis altostratomutatus results when a high grey nebulous altostratus layer thins out into a whitish layer of featureless high cloud.[30][48][49]

Middle stratocumuliform and stratiform

File:Misc clouds santa clarita.JPG
Middle clouds over Santa Clarita, CA. Altocumulus floccus producing virga near top and middle of image merging into altostratus translucidus near horizon.
This is a middle-étage stratocumuliform cloud layer of limited convection that is usually appears in the form of irregular patches or rounded masses in groups, lines, or waves. High altocumulus may resemble cirrocumulus but is usually thicker and composed of water droplets so that the bases show at least some light-grey shading. Opaque altocumulus associated with a weak frontal or low-pressure disturbance can produce virga, very light intermittent precipitation that evaporates before reaching the ground. If the altocumulus is mixed with moisture-laden altostratus, the precipitation may reach the ground.
  • Species: Altocumulus has the same four species as cirrocumulus. The stratiformis species (Ac str) is composed of sheets or relatively flat patches of stratocumuliform cloud.[48][49] The synoptic coding is determined by the predominant variety or occasionally by the genitus mother cloud. Altocumulus lenticularis (Ac len) is a lens-shaped middle cloud similar to its cirrocumulus counterpart but usually with at least some grey shading. It is coded CM4 on the SYNOP weather observation. Grey shading is also seen with altocumulus castellanus (Ac cas), a turreted middle cloud that can achieve significant vertical development and signals increasing airmass instability. It is nevertheless usually classified as middle rather than vertical and is coded CM8. The floccus species (Ac flo) is a tufted middle cloud which is also associated with greater instability.[48][49] It shares the same code CM8. Chaotic altocumulus, which is typically poorly defined with multiple species or transitional forms arranged in several layers, is coded CM9.
  • Opacity-based varieties: Altocumulus stratiformis has three opacity-based varieties; Translucidus (CM3), perlucidus (CM3 or 7 depending on predominant opacity), and opacus (CM7). Varieties based on opacity are not commonly associatied with the species lenticularis, castellanus, or floccus.[48][49]
  • Pattern-based varieties: Radiatus (arranged in parallel bands) is sometimes seen with the stratformis and castellanus species. Altocumulus stratiformis radiatus of any opacity is coded CM5 if it is increasing in amount. The duplicatus or undulatus varieties are occasionally seen with the stratiformis and lenticularis species. Altocumulus stratiformis duplicatus is coded CM7 if it is not overridden by another coding of higher importance. Lacunosus is very occasionally associated with altocumulus of the species stratiformis, castellanus, or floccus.[49]
  • Supplementary feature: Altocumulus often produces virga but usually not the precipitatio feature that reaches the ground.
  • Accessory cloud: Mamma caused by localized downdrafts in the cloud layer are occasionally seen with altocumulus.
  • Genitus Mother clouds: Altocumulus stratiformis cumulogenitus or cumulonimbogenitus can form when the middle or upper part of a towering free convective cloud begins to spread horizontally due to a loss of convective lift. It is coded CM6.
  • Mutatus mother clouds: Altocumulus can form due to the complete transformation of cirrocumulus, altostratus, nimbostratus, or stratocumulus.[48][49]
Altostratus translucidus near top of photo merging into altostratus opacus near bottom
Altostratus is a mid-level opaque or translucent stratiform or non-convective veil of grey/blue-grey cloud that often forms along warm fronts and around low-pressure areas where it may thicken into nimbostratus. Altostratus is usually composed of water droplets but may be mixed with ice crystals at higher altitudes. Widespread opaque altostratus can produce light continuous or intermittent precipitation.
  • Species: Altostratus is not divided into species because it is always nebulous or featureless in structure.
  • Opacity-based varieties: Altocumulus translucidus is relatively thin so that the sun or moon is always visible as if seen through frosted glass. It is strictly a middle cloud and is coded CM1 in the SYNOP report. The opacus variety is sufficiently thick to obscure the sun or moon and can extend vertically into the high étage. It is therefore sometimes classified as vertical or multi-level, but is still coded CM2 as a middle cloud.
  • Pattern-based varieties: Radiatus, duplicatus, and undulatus are all occasionally associated with Altostratus.[49]
  • Precipitation-based supplementary features: Altostratus opacus can be thick enough to produce both virga or praecipitatio features.
  • Accessory clouds: Pannus forming in precipitation is the most common cloud-based supplementary feature and is coded CL7. Mamma may be occasionally seen with altostratus, especially if it is associated with or changing to or from altocumulus.
  • Genitus mother clouds: Altostratus altocumulogenitus forms due to the fusing of altocumulus elements. Altostratus cumulonimbogenitus results from the spreading of the middle part of a cumulonimbus cloud.[48][49]
  • Mutatus mother clouds: Cirrostratus can thicken into altostratus cirrostratomutatus, usually ahead of an approaching disturbance. Nimbostratus associated with an organized weather system may lift and change into to altostratus nimbostratomutatus, especially if the disturbance is weakening or moving out of a particular area.[30]

Low stratocumuliform, stratiform, and cumuliform

Stratocumulus stratiformis perlucidus over Galapagos, Tortuga Bay
This genus type is a stratocumuliform cloud layer of limited convection, usually in the form of irregular patches or rounded masses similar to altocumulus but having larger elements with deeper-gray shading. Opaque stratocumulus associated with a weak frontal or low-pressure disturbance can produce very light intermittent precipitation. This cloud often forms under a precipitating deck of altostratus or high-based nimbostratus associated with a well-developed warm front, slow-moving cold front, or low-pressure area. This can create the illusion of continuous precipitation of more than very light intensity falling from stratocumulus. All species are coded CL5 except when formed from free convective mother clouds (CL4) or when formed separately from co-existing cumulus (CL8).
  • Species: Stratocumulus has three species which it shares in common with the other stratocumuliform genus types. The stratiformis species (Sc str) consists of sheets or relatively flat patches of low cloud similar if thicker in structure to the higher altocumulus and cirrocumulus types. Stratocumulus lenticularis (Sc len) and castellanus (Sc cas) also have similar structures to their Ac and Cc counterparts.[48][49] As with the other stratocumuliform genus-types, castellanus species can show vertical development but are not usually grouped with the vertical or multi-étage clouds.
  • Opacity-based varieties: The translucidus, perlucidus, and opacus varieties are the same for stratocumulus stratiformis as for Ac stratiformis. Varieties based on opacity are not commonly associated with species lenticularis or castellanus.
  • Pattern-based varieties: Parallel bands of radiatus are occasionally seen with the stratiformis species. Duplicatus and undulatus varieties are sometimes associated with stratocumulus stratiformis and lenticularis. With increased airmass instability, lacunosus downdraft holes may appear in layers of stratocumulus stratiformis and castellanus.[49]
  • Precipitation-based supplementary features: Virga or praecipitatio features of weak intensity may be seen with stratocumulus.
  • Accessory cloud: Mamma in the form of downward facing bubble-like protuberances may form as a result of localized downdrafts in the cloud layer.[48][49]
  • Genitus mother clouds: Stratocumulus may form from the spreading of cumulus or cumulonimbus (CL4), or the partial transformation of altostratus or nimbostratus.
  • Mutatus mother clouds: This genus type may also result from the complete transformation of altocumulus, nimbostratus, or stratus.
At level with stratus nebulosus translucidus clouds
This is a low-étage non-convective stratiform type resembling fog but not resting on the ground. Only very weak precipitation can fall from this cloud (usually drizzle or snow grains), although heavier rain or snow may fall through a stratus layer from a higher precipitating cloud deck.
  • Species: There are two species of stratus.[48][49] The nebulosus species (St neb) is a featureless veil of low cloud sometimes producing light drizzle that is coded CL6 in the SYNOP report. Stratus fractus (St fra) appears as a ragged broken up sheet that often forms as an accessory cloud in precipitation falling from a higher cloud deck. It is coded CL7 when associated with bad weather. This species may also result from a continuous sheet of stratus in clear air becoming broken up by the wind, in which case it is coded CL6, the same as for stratus nebulosis not associated with bad weather.
  • Opacity-based varieties: Stratus nebulosus can appear as translucidus or opacus depending on its thickness. The perlucidus variety is not usually associated with this species because of its relatively even structure.
  • Pattern-based variety: Stratus nebulosus usually shows no patterns. However a slightly disturbed gentle wind current can create a mild undulatus pattern, but this is rarely seen. Varieties of any kind are not commonly associated with stratus fractus because of the highly fragmented structure that identifies this species.[49]
  • Precipitation-based supplementary feature: Stratus cloud is too low to produce virga, but the praecipitatio feature can be seen in the form of drizzle or snow grains.
  • Accessory clouds: Stratus does not have any accessory clouds as such, but may form in precipitation as a cloud based supplementary feature associated with other precipitating clouds.
  • Genitus mother clouds: Stratus can form from the spreading or thinning of the base of clouds with significant vertical development, particularly nimbostratus, cumulus, or cumulonimbus.
  • Mutatus mother clouds: This genus type can form as the result of the fusing of stratocumulus elements into an even featureless sheet.[30][48][49]
  • Genus cumulus (Cu) – little vertical extent:
Low cumulus humilis and some moderate vertical cumulus mediocris in the foreground and background with some stratocumulus stratiformis perlucidus clouds mainly in the foreground.
These are small fair-weather cumuliform clouds of limited convection that do not grow vertically and generally do not produce rain showers. They are sometimes accompanied by low-étage stratocumulus or more vertically developed cumulus mediocris..
  • Species: Cumulus species are mainly indicators of degrees of vertical development.[48][49] The smallest type is cumulus fractus (Cu fra) which consists of cumulus broken up into ragged and changing fragments. Fair weather Cu fractus is coded CL1. It can also form in precipitation as a pannus accessory cloud which is coded CL7. Cumulus humilis (Cu hum) is the smallest non-ragged cloud and usually shows a light-grey shading underneath. Fair weather Cu humilis is also coded CL1 in the SYNOP code. Cumulus fractus and humilis are two species that cannot be described as vertical in the true sense of the word. Being at or near the beginning of the convective cloud's daily life cycle, they lack the moderate vertical extent of cumulus mediocris. Consequently they are commonly classified as low clouds[41] despite the fact their bases can be in the middle étage when the moisture content of the air is very low. When cumulus fractus and cumulus humilis are classified as vertical,[38] it is on the basis of their potential for at least moderate upward growth during their daily cycle.
  • Opacity-based varieties: Cumulus fractus is inherently translucent and the humilis species is generally opaque, so these do not have opacity- based varieties.
  • Pattern-based varieties: Radiatus is occasionally seen with fair-weather cumulus when arranged in parallel rows.
  • Supplementary features: These are not commonly seen with small cumulus, but Cu fractus of bad weather may be seen as a pannus feature with precipitating clouds.
  • Genitus mother clouds: Cumulus fractus or humilis may form as the result of a partial transformation of altocumulus or stratocumulus.
  • Mutatus mother clouds: These cumulus species may also appear due to a complete transformation of stratocumulus or stratus.[48][49]

Vertical or multi-étage stratiform, cumuliform, and cumulonimbiform (low to middle cloud base)

Moderate or deep vertical
Moderate vertical nimbostratus cloud covering the sky with a scattered layer of low stratus fractus in the middle of the upper half of the image.
This is a diffuse dark-grey non-convective stratiform layer with great horizontal extent and moderate to great vertical development. It lacks towering structure and looks feebly illuminated from the inside. Ns normally forms from altostratus and develops at least moderate vertical extent when the base subsides into the low étage during precipitation that can reach moderate to heavy intensity. It commonly achieves deep vertical development when it simultaneously grows upward into the high étage due to large scale frontal or cyclonic lift. Nimbostratus can have multiple associations as low, middle, or multi-étage depending on the criteria used by various authorities. It is coded CM2 on the SYNOP report.
  • Species and varieties: Nimbostratus is very thick, opaque, and featureless, so this genus type is not subdivided into species or varieties.[48][49]
  • Precipitation-based supplementary features: Nimbostratus is a major precipition cloud and produces the virga or praecipitatio features. The latter can achieve heavy intensity due to the cloud's vertical depth.
  • Accessory cloud: Pannus frequently forms in precipitation and is coded CL7.
  • Genitus mother clouds: This genus type can form from cumulus and cumulonimbus.
  • Mutatus mother clouds: Nimbostratus can form due to the complete transformation of altocumulus, altostratus and stratocumulus.[30][48][49]
  • Genus cumulus (Cu) – moderate vertical extent:
These cumuliform clouds of free convection generally have clear-cut flat bases and domed tops and are capable of producing showers.
  • Species Cumulus mediocris (Cu med): This species achieves moderate vertical development, has medium-grey shading underneath, and can produce scattered showers of light intensity.[48][49] This larger species is also classified as a low cloud by some authorities and is coded CL2 on the synop report.
  • Varieties: Cumulus mediocris is always opaque and therefore has no opacity-based varieties. A single pattern-based variety, radiatus, is sometime seen when the individual clouds are arranged into parallel rows.
  • Precipitation-based supplementary features: Cumulus mediocris can produce virga and praecipitatio features.
  • Accessory clouds: The pannus supplementary feature is sometimes seen with precipitating Cu mediocris, however the CL7 reporting code normally used with this feature is overridden by the CL2 code that identifies cumulus with significant vertical development. Pileus (cap cloud), velum (apron), arcus (roll or shelf cloud) and tuba (vertical column) features are also occasionally seen with cumulus mediocris.
  • Genitus mother clouds: Cumulus mediocris may form as a result of a partial transformation of altocumulus or stratocumulus.
  • Mutatus mother clouds: This genus and species type may also be the result of a complete transformation of stratocumulus or stratus.[30][48][49]
Towering vertical
Towering vertical cumulus congestus embedded within a layer of cumulus mediocris. Higher layer of stratocumulus stratiformis perlucidus.

These clouds are sometimes classified separately from the other major types because of their ability to produce severe turbulance.[46]

  • Genus cumulus (Cu) – great vertical extent:
Increasing airmass instability can cause free-convective cumulus to grow very tall to the extent that the vertical height from base to top is greater than the base-width of the cloud.
  • Species: Cumulus congestus (Cu Con) is the largest of the cumulus species and is designated separately as towering cumulus (Tcu) by the International Civil Aviation Organization : They grow upward to great vertical size, usually with dark-grey bases, and are capable of producing severe turbulence and showers of moderate to heavy intensity. Cumulus congestus is classified as multi-étage and is coded CL2 in the synop report. They usually too large and opaque to have any opacity or pattern-based varieties.
  • Non-WMO variant: Pyrocumulus (No official abbreviation) is a free convective cloud associated with volcanic eruptions and large-scale fires. Pyrocumulus is not recognized by the WMO as a distinct genus or species, but is, in essence, cumulus congestus formed under special circumstances that can also cause severe turbulence.[30][48][49]
This genus type is a heavy towering cumulonimbiform mass of free convective cloud with a dark-grey to nearly black base that is associated with thunderstorms and showers. Thunderstorms can produce a range of severe weather that includes hail, tornadoes, a variety of other localized strong wind events, several types of lightning, and local very heavy downpours of rain that can cause flash floods, although lightning is the only one of these that requires a thunderstorm to be taking place. In general, cumulonimbus require moisture, an unstable air mass, and a lifting force (heat) in order to form. Cumulonimbus typically go through three stages: the developing stage, the mature stage (where the main cloud may reach supercell status in favorable conditions), and the dissipation stage.[71] The average thunderstorm has a 24 km (15 mi) diameter. Depending on the conditions present in the atmosphere, these three stages take an average of 30 minutes to go through.[72] Cumulonimbus can have more than one altitude association and is sometimes included with the family of low clouds.
Stages of a cumulonimbus cloud's life.
Towering vertical cumulonimbus capillatus with anvil-shaped incus supplementary feature. High layer of cirrus spissatus near top of image.
  • Species: Cumulonimbus calvus (Cb cal) has a very high clear-cut domed top similar to towering cumulus and is coded CL3. The capillatus species (Cb cap) has very high tops that have become fibrous due to the presence of ice crystals.[48][49] It is coded CL9 in the SYNOP report.
  • Varieties: Cumulonimbus is too large and opaque to show any opacity or pattern-based varieties.
  • Precipitation-based supplementary features: This is also a major precipitation cloud and can produce virga or praecipitatio features, of which the latter can reach heavy intensity.
  • Accessory clouds: The cloud-based supplementary features normally associated with cumulonimbus are pannus, incus (cirriform anvil top), mamma, pileus, velum, arcus, and tuba. As with precipitating cumulus, the CL7 coding for pannus is overridden by higher codes, in this case CL3 or 9 depending on the species of cumulonimbus. The tuba feature can develop into a funnel cloud, water spout, or tornado.[48][49]
  • Genitus mother clouds: Cumulonimbus can develop from altocumulus, altostratus, nimbostratus, stratocumulus, and cumulus.
  • Mutatus mother cloud: This genus type can also result from the complete transformation of cumulus undergoing rapid vertical growth.[48][49]

Polar stratospheric class

Formation and distribution

Stratospheric nacreous clouds over Antarctica

This class forms at altitudes of about 15,000–25,000 m (49,200–82,000 ft) during the winter when the stratosphere is coldest and has the best chance of triggering condensation caused by adiabatic cooling. It is typically very thin with an undulating cirriform appearance. Moisture is very scarce in the stratosphere, so cloud at this altitude range is rare and is usually restricted to polar regions where the air is coldest.[73]

Classification

The formation of Polar stratospheric cloud is limited to a single very high range of altitude, so this class is not divided into height-related families. Polar stratospheric has a generally cirriform structure and appearance and does not have separate genus types, species, or varieties. Instead, the classification is alpha-numeric and is based on chemical makeup rather than variations in physical appearance.

Very high cirriform

  • Type 1 (Non-nacreous): This type contains supercooled nitric acid and water droplets and lacks any special coloration. It is dividable into subtype 1A which is mostly made up of ice crystals and frozen nitric acid, and 1B which remains in a supercooled state and includes sulfuric acid and water in a ternary solution. Nacreous type 2 is sometimes associated or embedded.
  • Type 2 (Nacreous): Nacreous polar stratospheric cloud consists of ice crystals only and generally shows mother-of-pearl colors.

Polar mesospheric class

Formation and distribution

Noctilucent cloud over Estonia

Polar mesospheric clouds are the highest in the atmosphere and occur mostly at altitudes of 80 to 85 km (50 to 53 mi), which is about ten times the altitude of tropospheric high clouds.[74] From ground level, they can occasionally be seen illuminated by the sun during deep twilight. Ongoing research indicates that convective lift in the mesophere is strong enough during the polar summer to cause adiabatic cooling of small amount of water vapour to the point of saturation. This tends to produce the coldest temperatures in the entire atmosphere just below the mesopause. These conditions result in the best environment for the formation of polar mesospheric clouds.[75] There is also evidence that smoke particles from burnt-up meteors provide much of the condensation nuclei required for the formation of noctilucent cloud.[76]

Because of the need for maximum cooling of the water vapor to produce these clouds, their distribution tends to be restricted to polar regions of Earth during the respective summer seasons in the northern and southern hemispheres. Sightings are rare more than 40 degrees south of the north pole or north of the south pole.

Classification

Polar mesospheric clouds all tend to form at an extreme altitude range and are consequently not classified into height-related families. They are given the Latin name Noctilucent because of their illumination well after sunset and before sunrise. An alpha-numeric classification is used to identify variations in physical appearance.

Noctilucent (extremely high cirriform)

  • Type 1: The first type is characterized by very tenuous filaments resembling cirrus fibratus.
  • Type 2: This type comprises bands in the form of long streaks, often in groups or interwoven at small angles, similar to cirrus intortus. It is dividable into two subtypes; 2A where the streaks have diffuse, blurred edges, and 2B where they have sharply defined edges.
  • Type 3: Billows in the form of short streaks can be seen that are clearly spaced and roughly parallel. Subtype 3A has short, straight, narrow streaks while 3B has wave-like streaks similar to cirrus undulatus.
  • Type 4: This shows whirls in the form of partial or rarely complete rings with dark centers. With subtype 4A, the whirls are of small angular radius and have a similar appearance to surface water ripples. 4B is characterized by simple curves of medium angular radius with one or more bands. Subtype 4C has whirls with large-scale ring structure.

Throughout the homosphere

Coloration

An occurrence of altocumulus and cirrocumulus cloud iridescence
Sunset reflecting shades of pink onto grey stratocumulus clouds.
Effect of sunlight before sunset. Bangalore, India.

Striking cloud colorations can be seen at many altitudes in the homosphere, which includes the troposphere, stratosphere, and mesophere. The first recorded colored cloud was seen by Nathan Ingleton in 1651, he wrote the event in his diary but the records were destroyed in 1666, in the Great Fire of London. The color of a cloud, as seen from Earth, tells much about what is going on inside the cloud.

In the troposphere, dense, deep clouds exhibit a high reflectance (70% to 95%) throughout the visible spectrum. Tiny particles of water are densely packed and sunlight cannot penetrate far into the cloud before it is reflected out, giving a cloud its characteristic white color, especially when viewed from the top.[77] Cloud droplets tend to scatter light efficiently, so that the intensity of the solar radiation decreases with depth into the gases. As a result, the cloud base can vary from a very light to very-dark-grey depending on the cloud's thickness and how much light is being reflected or transmitted back to the observer. Thin clouds may look white or appear to have acquired the color of their environment or background. High tropospheric clouds appear mostly white if composed entirely of ice crystals or supercooled water droplets.

As a tropospheric cloud matures, the dense water droplets may combine to produce larger droplets. If the droplets become too large and heavy to be kept aloft by the air circulation, they will fall from the cloud as rain. By this process of accumulation, the space between droplets becomes increasingly larger, permitting light to penetrate farther into the cloud. If the cloud is sufficiently large and the droplets within are spaced far enough apart, a percentage of the light that enters the cloud is not reflected back out but is absorbed giving the cloud a darker look. A simple example of this is one's being able to see farther in heavy rain than in heavy fog. This process of reflection/absorption is what causes the range of cloud color from white to black.[78]

Other colors occur naturally in tropospheric clouds. Bluish-grey is the result of light scattering within the cloud. In the visible spectrum, blue and green are at the short end of light's visible wavelengths, whereas red and yellow are at the long end. The short rays are more easily scattered by water droplets, and the long rays are more likely to be absorbed. The bluish color is evidence that such scattering is being produced by rain-size droplets in the cloud. A cumulonimbus cloud that appears to have a greenish/bluish tint is a sign that it contains extremely high amounts of water; hail or rain. Supercell type storms are more likely to be characterized by this but any storm can appear this way. Coloration such as this does not directly indicate that it is a severe thunderstorm, it only confirms its potential. Since a green/blue tint signifies copious amounts of water, a strong updraft to support it, high winds from the storm raining out, and wet hail; all elements that improve the chance for it to become severe, can all be inferred from this. In addition, the stronger the updraft is, the more likely the storm is to undergo tornadogenesis and to produce large hail and high winds. Yellowish clouds may occur in the late spring through early fall months during forest fire season. The yellow color is due to the presence of pollutants in the smoke. Yellowish clouds caused by the presence of nitrogen dioxide are sometimes seen in urban areas with high air pollution levels.[79]

Particles in the atmosphere and the sun's angle enhance cloud colors at morning twilight

Within the troposphere, red, orange, and pink clouds occur almost entirely at sunrise/sunset and are the result of the scattering of sunlight by the atmosphere. When the angle between the sun and the horizon is less than 10 percent, as it is just after sunrise or just prior to sunset, sunlight becomes too red due to refraction for any colors other than those with a reddish hue to be seen.[80] The clouds do not become that color; they are reflecting long and unscattered rays of sunlight, which are predominant at those hours. The effect is much like if one were to shine a red spotlight on a white sheet. In combination with large, mature thunderheads, this can produce blood-red clouds. Clouds look darker in the near-infrared because water absorbs solar radiation at those wavelengths.

In high latitude regions of the stratosphere, nacreous clouds occasionally found there during the polar winter tend to display quite striking displays of mother-of-pearl colorations. This is due to the refraction and diffusion of the sun's rays through thin clouds with supercooled droplets that often contain compounds other than water. At still higher altitudes up in the mesospere, noctilucent clouds made of ice crystals are sometimes seen in polar regions in the summer. They typically have a silvery white coloration that can resemble brightly illuminated cirrus.

Effects on climate

Global cloud cover, averaged over the month of October 2009. NASA composite satellite image; larger image available here.
These maps display the fraction of Earth's area that was cloudy on average during each month from January 2005 to August 2013. The measurements were collected by the Moderate Resolution Imaging Spectroradiometer (MODIS) on NASA’s Terra satellite. Colors range from blue (no clouds) to white (totally cloudy). Like a digital camera, MODIS collects information in gridded boxes, or pixels. Cloud fraction is the portion of each pixel that is covered by clouds. Colors range from blue (no clouds) to white (totally cloudy).[81] (click for more detail)

The role of tropospheric clouds in regulating weather and climate remains a leading source of uncertainty in projections of global warming.[82][83] This uncertainty arises because of the delicate balance of processes related to clouds, spanning scales from millimeters to planetary. Hence, interactions between the large-scale (synoptic meteorology) and clouds becomes difficult to represent in global models. The complexity and diversity of clouds, as outlined above, adds to the problem. On the one hand, white-colored cloud tops promote cooling of Earth's surface by reflecting short-wave radiation from the sun. Most of the sunlight that reaches the ground is absorbed, warming the surface, which emits radiation upward at longer, infrared, wavelengths. At these wavelengths, however, water in the clouds acts as an efficient absorber. The water reacts by radiating, also in the infrared, both upward and downward, and the downward long-wave radiation results in some warming at the surface. This is analogous to the greenhouse effect of greenhouse gases and water vapor.[84]

High-étage tropospheric genus-types, cirrus, cirrocumulus, and cirrostratus, particularly show this duality with both short-wave albedo cooling and long-wave greenhouse warming effects. On the whole though, ice-crystal clouds in the upper troposphere tend to favor net warming.[85][86] However, the cooling effect is dominant with low-étage stratocumuliform and stratiform clouds made of very small water droplets that have an average diameter of about 0.002 mm (.00008 in).,[4] especially when they form in extensive sheets that block out more of the sun. These include middle-étage layers of altocumulus and altostratus as well as low stratocumulus, and stratus. Small-droplet aerosols are not good at absorbing long-wave radiation reflected back from Earth, so there is a net cooling with almost no long-wave effect. This effect is particularly pronounced with low-étage clouds that form over water.[85] Low and vertical heaps of cumulus, towering cumulus, and cumulonimbus are made of larger water droplets ranging in diameter from .005 to about 0.3mm. Nimbostratus cloud droplets can also be quite large, up to 0.3mm in diameter.[87] These larger droplets associated with vertically developed clouds are better able to trap the long-wave radiation thus migitating the cooling effect to some degree. However, these large often precipitating clouds are variable or unpredictable in their overall effect because of variations in their concentration, distribution, and vertical extent. Measurements taken by NASA indicate that on the whole, the effects of low and middle étage clouds that tend to promote cooling are outweighing the warming effects of high layers and the variable outcomes associated with multi-level or vertically developed clouds.[85]

As difficult as it is to evaluate the effects of current cloud cover characteristics on climate change, it is even more problematic to predict the outcome of this change with respect to future cloud patterns and events. As a consequence, much research has focused on the response of low and vertical clouds to a changing climate. Leading global models can produce quite different results, however, with some showing increasing low-étage clouds and others showing decreases.[88][89]

Polar stratospheric and mesospheric clouds are not common or widespread enough to have a significant effect on climate.[75] However an increasing frequency of occurrence of noctilucent clouds since the 19th century may be the result of climate change.

Global brightening

New research indicates a global brightening trend.[90] The details are not fully understood, but much of the global dimming (and subsequent reversal) is thought to be a consequence of changes in aerosol loading in the atmosphere, especially sulfur-based aerosol associated with biomass burning and urban pollution.[91] Changes in aerosol burden can have indirect effects on clouds by changing the droplet size distribution[92] or the lifetime and precipitation characteristics of clouds.[93]

Cloud properties

Detection

Data sets regarding cloud properties are gathered using satellites, such as MODIS, POLDER, CALIPSO or ATSR The instruments measure the radiances of the clouds, from which the relevant parameters can be retrieved. This is usually done by using inverse theory. [94].

The method of detection is based on the fact that the clouds tend to appear brighter and colder than the land surface. Because of this, difficulties rise in detecting clouds above bright (highly reflective) surfaces, such as oceans and ice[94].

Parameters

The value of a certain parameter is more reliable the more satellites are measuring the said parameter. This is due to the fact that the range of errors and neglected details varies from instrument to instrument. Thus, if the analysed parameter has similar values for different instruments, it is accepted that the true value lies in the range given by the corresponding data sets.

The Global Energy and Water Cycle Experiment uses the following quantities in order to compare data quality from different satellites in order to establish a reliable quantification of the properties of the clouds[94]:

  • the Cloud Amount with values between 0 and 1
  • the cloud temperature at top ranging from 150 to 340 K
  • the cloud pressure at top 1013 - 100 hPa
  • the cloud height, measured above sea level, ranging from 0 to 20 km
  • the cloud IR emissivity, with values between 0 and 1, with a global average around 0.7.
  • the effective cloud amount is the cloud amount weighted by the cloud IR emissivity, with a global average of 0.5
  • the cloud (visible) optical depth) varies within a range of 4 and 10.
  • the cloud water path for the liquid and solid (ice) phases of the cloud particles
  • the cloud effective particle size for both liquid and ice, ranging from 0 to 200 μm

Vertical Distribution

Data from satellites suggests the following vertical distribution of the cloud amount[94]:

  • High-level clouds, 40%-50%
  • Mid-level clouds, around 10%-20%
  • low-level clouds, around 40%

Extraterrestrial

Cloud cover has been seen on most other planets in the solar system. Venus's thick clouds are composed of sulfur dioxide and appear to be almost entirely stratiform.[95] They are arranged in three main layers at altitudes of 45 to 65 km that obscure the planet's surface and can produce virga.[96] No embedded cumuliform types have been identified, but broken stratocumuliform wave formations are sometimes seen in the top layer that reveal more continuous layer clouds underneath.[97] On Mars, cirrus, cirrocumulus and stratocumulus composed of water-ice have been detected mostly near the poles.[98] Water-ice fogs have also been detected on this planet.[99]

Both Jupiter and Saturn have an outer cirriform cloud deck composed of ammonia,[100][101] an intermediate stratiform haze-cloud layer made of ammonium hydrosulfide, and an inner deck of cumulus water clouds.[102][103] Embedded cumulonimbus are known to exist near the Great Red Spot on Jupiter.[104][105] The same category-types can be found covering Uranus, and Neptune, but are all composed of Methane.[106][107][108][109][110][111] Saturn's moon Titan has cirrus clouds believed to be composed largely of methane.[112][113] The Cassini–Huygens Saturn mission uncovered evidence of a fluid cycle on Titan, including lakes near the poles and fluvial channels on the surface of the moon.

In October 2013, the detection of high altitude optically thick clouds in the atmosphere of Kepler-7b was announced,[114][115] and, in December 2013, also in the atmospheres of GJ 436 b and GJ 1214 b.[116][117][118][119]

See also

References

  1. ^ "Weather Terms". National Weather Service. Retrieved 21 June 2013. {{cite web}}: Cite has empty unknown parameter: |1= (help)
  2. ^ R. Nave (2013). "Adiabatic Process". gsu.edu. Retrieved 18 November 2013.
  3. ^ a b c d e f g h Elementary Meteorology Online (2013). "Humidity, Saturation, and Stability". vsc.edu. Retrieved 18 November 2013.
  4. ^ a b Steve Horstmeyer (2008). "Cloud Drops, Rain Drops". Retrieved 19 March 2012.
  5. ^ Steve Horstmeyer (2008). "Why don't clouds fall?". Graz University of Technology. Retrieved 6 April 2012.
  6. ^ David R. Dowling and Lawrence F. Radke (1990). "A Summary of the Physical Properties of Cirrus Clouds". Journal of Applied Meteorology. 29 (9): 970–978 (977). Bibcode:1990JApMe..29..970D. doi:10.1175/1520-0450(1990)029<0970:ASOTPP>2.0.CO;2. OCLC 4643887498.
  7. ^ Pidwirny, M. (2006). "Cloud Formation Processes", chapter 8 in Fundamentals of Physical Geography, 2nd ed.
  8. ^ Steven A. Ackerman (2011). Meteorology:Fog Formation. Published by Jones & Bartlett. Retrieved 22 October 2013.
  9. ^ Glossary of Meteorology (2009). "Radiational cooling". American Meteorological Society. Retrieved 27 December 2008.
  10. ^ Robert Fovell (2004). "Approaches to saturation" (PDF). University of California in Los Angeles. Retrieved 7 February 2009.
  11. ^ Robert Penrose Pearce (2002). Meteorology at the Millennium. Academic Press. p. 66. ISBN 978-0-12-548035-2.
  12. ^ National Weather Service Office, Spokane, Washington (2009). "Virga and Dry Thunderstorms". National Oceanic and Atmospheric Administration. Retrieved 2 January 2009.{{cite web}}: CS1 maint: multiple names: authors list (link)
  13. ^ Bart van den Hurk and Eleanor Blyth (2008). "Global maps of Local Land-Atmosphere coupling" (PDF). KNMI. Retrieved 2 January 2009.
  14. ^ H. Edward Reiley, Carroll L. Shry (2002). Introductory horticulture. Cengage Learning. p. 40. ISBN 978-0-7668-1567-4.
  15. ^ JetStream (2008). "Air Masses". National Weather Service. Retrieved 2 January 2009.
  16. ^ American Heritage Science Dictionary (2010). "cohesion science definition". Retrieved 25 July 2012.
  17. ^ The Westminster review (2013). attraction cohesive cloud droplets. Retrieved 1 October 2013.
  18. ^ Kirill Iakovlevich Kondratʹev (2006). Atmospheric aerosol properties: formation, processes and impacts. Springer. p. 403. ISBN 978-3-540-26263-3.
  19. ^ a b Yochanan Kushnir (2000). "The Climate System: General Circulation and Climate Zones". Retrieved 13 March 2012.
  20. ^ Jack Williams (27 June 1997). "Extratropical storms are major weather makers". USA Today. Retrieved 13 March 2012.
  21. ^ National Centre for Medium Range Forecasting. Chapter-II Monsoon-2004: Onset, Advancement and Circulation Features. Retrieved on 3 May 2008.
  22. ^ Australian Broadcasting Corporation. Monsoon. Retrieved on 3 May 2008.
  23. ^ ThinkQuest team 26634 (1999). The Formation of Deserts. Oracle ThinkQuest Education Foundation. Retrieved on 16 February 2009.
  24. ^ John D. Cox (2002). Storm Watchers. John Wiley & Sons, Inc. pp. 13–17. ISBN 0471444863.
  25. ^ Met Office (U.K.) (2013). "Fact sheet No. 1 - Clouds" (PDF). Met Office (U.K.). Retrieved 21 November 2013.
  26. ^ a b c d e f g J. Koch (2013). "Cloud Types" (PDF). Brandon University. Retrieved 1 January 2014.
  27. ^ a b c Office National Meteorologique (1932). International Atlas of Clouds and of States of the Sky. Paris. p. 17.{{cite book}}: CS1 maint: location missing publisher (link)
  28. ^ Henry Glassford Bell, ed. (1827). Constable's miscellany of original and selected publications. Vol. XII. p. 320.
  29. ^ a b E.C. Barrett and C.K. Grant (1976). "The identification of cloud types in LANDSAT MSS images". NASA. Retrieved 22 August 2012.
  30. ^ a b c d e f g h i j k l m n o p World Meteorological Organization (1995). "WMO cloud classifications" (PDF). Retrieved 1 February 2012.
  31. ^ "Cumulus clouds". Weather. USA Today. 16 October 2005. Retrieved 16 October 2012.
  32. ^ Stommel 1947, p. 91
  33. ^ Mossop & Hallett 1974, pp. 632–634
  34. ^ Scott, A (2000). "The Pre-Quaternary history of fire". Palaeogeography Palaeoclimatology Palaeoecology. 164 (1–4): 281. doi:10.1016/S0031-0182(00)00192-9.
  35. ^ National Center for Atmospheric Research (2008). "Hail". University Corporation for Atmospheric Research. Retrieved 18 July 2009.
  36. ^ Ted Fujita (1985). "The Downburst, microburst and macroburst". SMRP Research Paper 210, 122 pp.
  37. ^ Nilton O. Renno (August 2008). "A thermodynamically general theory for convective vortices" (PDF). Tellus A. 60 (4): 688–99. Bibcode:2008TellA..60..688R. doi:10.1111/j.1600-0870.2008.00331.x.
  38. ^ a b c d e f g h Clouds Online (2012). "Cloud Atlas". Retrieved 1 February 2012.
  39. ^ a b c d e Jim Koermer (2011). "Plymouth State Meteorology Program Cloud Boutique". Plymouth State University. Retrieved 31 January 2011.
  40. ^ a b c "cloud: Classification of Clouds". Infoplease.com.
  41. ^ a b c d e f JetStream (5 January 2010). "Cloud Classifications". National Weather Service. Retrieved 31 January 2011.
  42. ^ a b American Meterological Society (2012). "Glossary of Meteorology". Retrieved 9 January 2014. Cite error: The named reference "nimbostratus" was defined multiple times with different content (see the help page).
  43. ^ Steven A. Ackerman (2011). Meteorology:Precipitating Clouds. Published by Jones & Bartlett. Retrieved 22 October 2013.
  44. ^ Becca Hatheway (2009). "Cloud Types". Windows to the Universe, US National Earth Science Teachers Association (NESTA). Retrieved 15 September 2011.
  45. ^ Paul Hancock and Brian Skinner (2000). "clouds. The Oxford Companion to the Earth". Retrieved 22 February 2011.
  46. ^ a b Paul de Valk, Rudolf van Westhrenen, and Cintia Carbajal Henken (2010). "Automated CB and TCU detection using radar and satellite data: from research to application" (PDF). Retrieved 15 September 2011.{{cite web}}: CS1 maint: multiple names: authors list (link)
  47. ^ Robert A. Houze (1994). Cloud Dynamics. Retrieved 21 November 2013.
  48. ^ a b c d e f g h i j k l m n o p q r s t u v w x y z aa ab ac ad ae af ag ah ai aj ak al am Morris (2008). "Clouds – Species and Varieties". University of Minnesota. Retrieved 4 February 2012.
  49. ^ a b c d e f g h i j k l m n o p q r s t u v w x y z aa ab ac ad ae af ag ah ai aj ak al am an ao Aerographer/Meteorology (2012). "Cloud Variety". meteorologytraining.tpub.com. Retrieved 2 July 2012.
  50. ^ Pretor-Pinney 2007, p. 20
  51. ^ Dunlop 2003, pp. 77–78
  52. ^ Allaby, Michael, ed. (2010). "Pannus". A Dictionary of Ecology (4 ed.). Oxford University Press. doi:10.1093/acref/9780199567669.001.0001/acref-9780199567669-e-4082. ISBN 9780199567669.
  53. ^ "Cumulonimbus Incus". Universities Space Research Association. 5 August 2009. Retrieved 23 October 2012.
  54. ^ Garret, et al. 2006, p. i
  55. ^ Ludlum 2000, p. 473
  56. ^ Aerographer/Meteorology (2012). "Roll cloud formation on cumulonimbus". Retrieved 5 July 2012.
  57. ^ Dunlop 2003, p. 79
  58. ^ Ilan Koren & Graham Feingold (2013). "Adaptive behavior of marine cellularclouds". Retrieved 1 October 2013.
  59. ^ "Cloud Formations off the West Coast of South America". NASA Earth Observatory. Retrieved 29 March 2013.
  60. ^ Steven A. Ackerman (2011). Meteorology:cirrocumulus. Published by Jones & Bartlett. Retrieved 22 October 2013.
  61. ^ Chris C. Park (2001). The environment: principles and applications. Psychology Press. p. 309. ISBN 978-0-415-21771-2. Retrieved 17 December 2010.
  62. ^ a b "Mackerel sky". Weather Online. Retrieved 21 November 2013.
  63. ^ Lee M. Grenci and Jon M. Nese (2001). A World of Weather: Fundamentals of Meteorology: A Text / Laboratory Manual (3 ed.). Kendall/Hunt Publishing Company. p. 212. ISBN 978-0-7872-7716-1. OCLC 51160155.
  64. ^ Preview Text, Holton, James R., 2004: An Introduction to Dynamic Meteorology. Academic Press, 277 pp.
  65. ^ Preview Text, Carlson, Toby N., 1991: Mid-Latitude Weather Systems. HarperCollins, 435 pp.
  66. ^ Weston 1980, pp. 437–438
  67. ^ "Weather Glossary". The Weather Channel. Retrieved 18 October 2012.
  68. ^ Brent Christner (28 February 2008). "LSU scientist finds evidence of 'rain-making' bacteria". American Association for the Advancement of Science. Retrieved 31 January 2011.
  69. ^ Christine Dell'Amore (12 January 2009). "Rainmaking Bacteria Ride Clouds to "Colonize" Earth?". National Geographic. Retrieved 31 January 2011.
  70. ^ Miyazaki et al. 2001, p. 364
  71. ^ Michael H. Mogil (2007). Extreme Weather. New York: Black Dog & Leventhal Publisher. pp. 210–211. ISBN 978-1-57912-743-5.
  72. ^ National Severe Storms Laboratory (15 October 2006). "A Severe Weather Primer: Questions and Answers about Thunderstorms". National Oceanic and Atmospheric Administration. Retrieved 1 September 2009.
  73. ^ Les Cowley (2011). "Nacreous clouds". Atmospheric optics, atoptics.co.uk. Retrieved 31 January 2012.
  74. ^ Michael Gadsden and Pekka Parviainen (September 2006). Observing Noctilucent Clouds (PDF). International Association of Geomagnetism & Aeronomy. p. 9. Retrieved 31 January 2011.
  75. ^ a b Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1016/0032-0633(82)90126-X, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1016/0032-0633(82)90126-X instead.
  76. ^ Karen C. Fox (2013). "NASA Sounding Rocket Observes the Seeds of Noctilucent Clouds". Retrieved 1 October 2013.
  77. ^ Increasing Cloud Reflectivity, Royal Geographical Society, 2010.
  78. ^ Bette Hileman (1995). "Clouds absorb more solar radiation than researchers previously thought". Chemical & Engineering News. 73 (7): 33. doi:10.1021/cen-v073n007.p033.
  79. ^ Cities and Air Pollution, Nature, 1998, chapter 10
  80. ^ Frank W. Gallagher, III. (October 2000). "Distant Green Thunderstorms – Frazer's Theory Revisited". Journal of Applied Meteorology. 39 (10). American Meteorological Society: 1754–1757. Bibcode:2000JApMe..39.1754G. doi:10.1175/1520-0450-39.10.1754.
  81. ^ http://earthobservatory.nasa.gov/GlobalMaps/view.php?d1=MODAL2_M_CLD_FR
  82. ^ D. Randall, R. Wood, S. Bony, R. Colman, T. Fichefet, J. Fyfe, V. Kattsov, A. Pitman, J. Shukla, J. Srinivasan, R. Stouffer, A. Sumi, and K. Taylor (2007) "Climate models and their evaluation" in S. Solomon, D. Qin, M. Manning, Z. Chen, M. Marquis, K. Averyt, M.Tignor, and H. Miller (eds.) Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA.
  83. ^ "Will Clouds Speed or Slow Global Warming?". National Science Foundation. Retrieved 23 October 2012.
  84. ^ "Cloud Climatology". International Satellite Cloud Climatology Program. National Aeronautics and Space Administration. Retrieved 12 July 2011.
  85. ^ a b c Steven A. Ackerman (2011). Meteorology:Clouds and the Greenhouse Effect. Published by Jones & Bartlett. Retrieved 23 October 2013.
  86. ^ Franks F. (2003). "Nucleation of ice and its management in ecosystems" (PDF). Philosophical Transactions of the Royal Society A. 361 (1804): 557–574. Bibcode:2003RSPTA.361..557F. doi:10.1098/rsta.2002.1141. PMID 12662454.
  87. ^ T. Okita (1961). "Size Distribution of Large Droplets in Precipitating Clouds". tellusa. Retrieved 24 November 2013.
  88. ^ S. Bony and J.-L. Dufresne (2005). "Marine boundary layer clouds at the heart of tropical cloud feedback uncertainties in climate models" (PDF). Geophysical Research Letters. 32 (20). Bibcode:2005GeoRL..3220806B. doi:10.1029/2005GL023851.
  89. ^ B. Medeiros, B. Stevens, I.M. Held, M. Zhao, D.L. Williamson, J.G. Olson, and C.S. Bretherton (2008). "Aquaplanets, Climate Sensitivity, and Low Clouds". Journal of Climate. 21 (19): 4974. Bibcode:2008JCli...21.4974M. doi:10.1175/2008JCLI1995.1.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  90. ^ Martin Wild, Hans Gilgen, Andreas Roesch, Atsumu Ohmura, Charles N. Long, Ellsworth G. Dutton, Bruce Forgan, Ain Kallis, Viivi Russak, and Anatoly Tsvetkov (2005). "From Dimming to Brightening: Decadal Changes in Solar Radiation at Earth's Surface". Science. 308 (5723): 847–50. Bibcode:2005Sci...308..847W. doi:10.1126/science.1103215. PMID 15879214.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  91. ^ Costantino, L. and F.-M. Bréon (2010). "Analysis of aerosol-cloud interaction from multi-sensor satellite observations". Geophysical Research Letters. 37 (11): n/a. Bibcode:2010GeoRL..3711801C. doi:10.1029/2009GL041828.
  92. ^ S. A. Twomey (1974). "Pollution and the planetary albedo". Atmospheric Environment (1967). 8 (12): 1251. doi:10.1016/0004-6981(74)90004-3.
  93. ^ B. Stevens and G. Feingold (2009). "Untangling aerosol effects on clouds and precipitation in a buffered system". Nature. 461 (7264): 607–13. Bibcode:2009Natur.461..607S. doi:10.1038/nature08281. PMID 19794487.
  94. ^ a b c d Stubenrauch, C. J.; Rossow, W. B.; Kinne, S.; Ackerman, S.; Cesana, G.; Chepfer, H; Di Girolamo, L.; Getzewich, B.; Guignard, A.; Heidinger, A.; Maddux, B. C.; Menzel, W.P; Minnis, P.; Pearl, C.; Platnick, S.; Poulsen, C.; Reidi, J.; Sun-Mack, S; Walther, A.; Winker, D.; Zeng, S.; Zhao, G. (2013). "Assessment of global cloud datasets from satellites: Project and Database initiated by GEWEX Radiation Panel" (pdf). Bulletin of the American Meteorological Society. 94: 1031–1049. doi:10.1175/BAMS-D-12-00117.1.
  95. ^ Bougher & Phillips 1997, pp. 127–129
  96. ^ Franck Montmessin (2013). "Clouds in the terrestrial planets" (PDF). Retrieved 5 November 2013.
  97. ^ David Shiga (2006). "Mysterious waves seen in Venus's clouds". New Scientist. Retrieved 5 November 2013.
  98. ^ "Clouds Move Across Mars Horizon". Phoenix Photographs. National Aeronautics and Space Administration. 19 September 2008. Retrieved 15 April 2011.
  99. ^ "NASA SP-441: Viking Orbiter Views of Mars". National Aeronautics and Space Administration. Retrieved 26 January 2013.
  100. ^ Phillips, Tony (20 May 2010). "Big Mystery: Jupiter Loses a Stripe". Nasa Headline News – 2010. National Aeronautics and Space Administration. Retrieved 15 April 2011.
  101. ^ Dougherty & Esposito 2009, p. 118
  102. ^ A.P. Ingersoll, T.E. Dowling, P.J. Gierasch, G.S. Orton, P.L. Read, A. Sanchez-Lavega, A.P. Showman, A.A. Simon-Miller, A.R. Vasavada. "Dynamics of Jupiter's Atmosphere" (PDF). Lunar & Planetary Institute. Retrieved 1 February 2007.{{cite web}}: CS1 maint: multiple names: authors list (link)
  103. ^ Monterrey Institute for Research in Astronomy (11 August 2006). "Saturn". Retrieved 31 January 2011.
  104. ^ "Thunderheads on Jupiter". Jet Propulsion Laboratory. National Aeronautics and Space Administration. Retrieved 26 January 2013.
  105. ^ Minard, Anne (14 October 2008). "Mysterious Cyclones Seen at Both of Saturn's Poles". National Geographic News. National Geographic. Retrieved 26 January 2013.
  106. ^ Nola Taylor Redd (2012). "Neptune's Atmosphere: Composition, Climate, & Weather". Space.com. Retrieved 5 November 2013.
  107. ^ Boyle, Rebecca (18 October 2012). "Check Out The Most Richly Detailed Image Ever Taken Of Uranus". Popular Science. Retrieved 26 January 2013.
  108. ^ Irwin 2003, p. 115
  109. ^ "Uranus". Scholastic. Archived from the original on 24 July 2011. Retrieved 16 April 2011.
  110. ^ Attention: This template ({{cite doi}}) is deprecated. To cite the publication identified by doi:10.1146/annurev.aa.31.090193.001245, please use {{cite journal}} (if it was published in a bona fide academic journal, otherwise {{cite report}} with |doi=10.1146/annurev.aa.31.090193.001245 instead.
  111. ^ Linda T. Elkins-Tanton (2006). Uranus, Neptune, Pluto, and the Outer Solar System. New York: Chelsea House. pp. 79–83. ISBN 0-8160-5197-6.
  112. ^ Athéna Coustenis and F.W. Taylor (2008). Titan: Exploring an Earthlike World. World Scientific. pp. 154–155. ISBN 978-981-270-501-3.
  113. ^ "Surprise Hidden in Titan's Smog: Cirrus-Like Clouds". Mission News. National Aeronautics and Space Administration. 3 February 2011. Retrieved 16 April 2011.
  114. ^ Chu, Jennifer (2 October 2013). "Scientists generate first map of clouds on an exoplanet". MIT. Retrieved 2 January 2014.
  115. ^ "Inference of Inhomogeneous Clouds in an Exoplanet Atmosphere". arXiv. 30 September 2013. arXiv:1309.7894. Retrieved 2 January 2014. {{cite journal}}: Unknown parameter |authors= ignored (help)
  116. ^ Harrington, J.D.; Weaver, Donna; Villard, Ray (31 December 2013). "Release 13-383 - NASA's Hubble Sees Cloudy Super-Worlds With Chance for More Clouds". NASA. Retrieved 1 January 2014.
  117. ^ Moses, Julianne (1 January 2014). "Extrasolar planets: Cloudy with a chance of dustballs". Nature (journal). 505: 31–32. doi:10.1038/505031a. Retrieved 1 January 2014.
  118. ^ "A featureless transmission spectrum for the Neptune-mass exoplanet GJ 436b". Nature (journal). 505: 66–68. 1 January 2014. doi:10.1038/nature12887. Retrieved 1 January 2014. {{cite journal}}: Unknown parameter |authors= ignored (help)
  119. ^ "Clouds in the atmosphere of the super-Earth exoplanet GJ 1214b". Nature (journal). 505: 69–72. 1 January 2014. doi:10.1038/nature12888. Retrieved 1 January 2014. {{cite journal}}: Unknown parameter |authors= ignored (help)

Bibliography

Template:Link GA Template:Link FA Template:Link FA Template:Link GA