Green radiation astronomy

Jump to navigation Jump to search

Editor-In-Chief: Henry A. Hoff

File:Northern Lights 02.jpg
Northern Lights are usually green, but in this image there is the very rare blue light. Credit: Varjisakka.

Green astronomy is the study and use of emission and absorption lines or bands in the wavelength range 495-570 nm. The use of filters with respect to this wavelength range are common when studying the Sun. Green astronomy also studies green sources and objects.

Green radiation

File:Color icon green.svg
These color squares show a variety of greens. Credit: FedericoMP.
File:RGB illumination.jpg
Green, blue and red are additive colors. All the colors you see on your computer screen are made by mixing them in different intensities. Credit: Bb3cxv.
The word green has the same Germanic root as the words for grass and grow and is a common color reflected by leaves on Earth. Credit: The cat.
File:Azurite-Malachite-23ub.jpg
Malachite is a mineral occurring on Earth, like many greens, is colored by the presence of copper, specifically by basic copper(II) carbonate.[1] Credit: Rob Lavinsky.

Green has a wavelength range of approximately 520–570 nm, a frequency range of ~575–525 THz, with color coordinates of (0, 255, 0) and a hexagonal triplet of #00FF00 from sRGB source of sRGB approximation to NCS S 2060-G.[2]

Def. the colour of growing foliage, as well as other plant cells containing chlorophyll; the colour between yellow and blue in the visible spectrum; one of the primary additive colour for transmitted light; the colour obtained by subtracting red and blue from white light using cyan and yellow filters is called green.

"...in nature chiefly conspicuous as the colour of growing herbage and leaves..."[3]

Green is the color of emeralds, jade, and growing grass.[3] In the continuum of colors of visible light it is located between yellow and blue. Green is the color most commonly associated with nature and the environmental movement, Islam, spring, hope and envy.[4]

Green is the color you see when you look at light with a wavelength of roughly 520–570 nanometers.

It is one of the three additive colors, along with red and blue, which are combined on computer screens and color televisions to make all other colors.

In the subtractive color system, used in printing, it is not a primary color, but is created out of a mixture of yellow and blue, or yellow and cyan.

On the HSV color wheel, also known as the RGB color wheel, the complement of green is magenta; that is, a purple color corresponding to an equal mixture of red and blue light. On a color wheel based on traditional color theory (RYB), the complementary color to green is considered to be red.[5]

The perception of greenness (in opposition to redness forming one of the opponent mechanisms in human color vision) is evoked by light which triggers the medium-wavelength M cone cells in the eye more than the long-wavelength L cones. Light which triggers this greenness response more than the yellowness or blueness of the other color opponent mechanism is called green. A green light source typically has a spectral power distribution dominated by energy with a wavelength of roughly 487–570 nm. More specifically, "blue green" 487–493 nm, "bluish green" 493–498 nm, "green" 498–530 nm, "yellowish green" 530–559 nm, "yellow green" 559–570 nm.[6]

Green earth is a natural pigment. It s composed of clay colored by iron oxide, magnesium, aluminum silicate, or potassium. Large deposits were found in the South of France near Nice, and in Italy around Verona, on Cyprus, and in Bohemia. The clay was crushed, washed to remove impurities, then powdered. It was sometimes called Green of Verona.[7]

Entities

File:Osiris-tomb-of-Nefertari.jpg
This is a detail of a frieze on a wall of tomb QV66, the burial place of Nefertari (c. 1295-1255 B.C.), royal wife of Ramesses the Great, featuring the Egyptian god Osiris. Credit: Mrgoodgame.
File:Rahu graha.JPG
This is an image of Rahu the head of the demonic snake. Credit: E. A. Rodrigues.

When any effort to acquire a system of laws or knowledge focusing on an astr, aster, or astro, that is, any natural body in the sky especially at night,[8] discovers an entity emitting, reflecting, or fluorescing green, succeeds even in its smallest measurement, green astronomy is the name of the effort and the result. Once an entity, source, or object has been detected as emitting, reflecting, or fluorescing green, it may be necessary to determine what the mechanism is. Usually this information provides understanding of the same entity, source, or object. Green is a color that suggests the sense of sight. Most people associate astronomy with the sense of seeing, what can be termed visual astronomy. As telescope optics transmit green well, green astronomy is also a field within optical astronomy.

Does the entity transit, continue occupation of a position for longer periods, or apparently change course or position in the sky? Is it a wanderer?

Do green entities fall from the sky? Are there green meteorites?

Are biological life forms able to perform astronomy in the green portion of the visible spectrum? If so the organisms may be conducting astrobiology.

Have observers recorded images of sky entities in the green?

At left is Rahu. Rahu is God of the Ascending / North lunar node. Rahu is the head of the demonic snake that swallows the sun or the moon causing eclipses, according to Hindu scriptures. He is depicted in art as a dragon with no body riding a chariot drawn by eight black horses. He is a Tamas Asura who does his best to plunge any area of one's life he controls into chaos.

Planetary sciences

File:Newgrange ireland 750px.jpg
The rising Sun illuminates the inner chamber of Newgrange, Ireland, only at the winter solstice. Credit: Richard Gallagher.
File:Polarlicht.jpg
The Aurora Borealis, or Northern Lights, shines above Bear Lake, Alaska. Credit: Senior Airman Joshua Strang, United States Air Force.

Of the planets in orbit around the Sun, only the Earth occasionally is seen to have green over large areas of its surface. This is usually from the presence of chlorophyll which obtains its green color from magnesium. The first image at right has green from this organic pigment.

Both the image at the page top and the second one at right show green colors in or above the atmosphere of the Earth.

The Earth of the early Archean (3,800 to 2,500 million years ago) may have had a different tectonic style. During this time, the Earth's crust cooled enough that rocks and continental plates began to form. Some scientists think because the Earth was hotter, that plate tectonic activity was more vigorous than it is today, resulting in a much greater rate of recycling of crustal material. This may have prevented cratonisation and continent formation until the mantle cooled and convection slowed down. Others argue that the subcontinental lithospheric mantle is too buoyant to subduct and that the lack of Archean rocks is a function of erosion and subsequent tectonic events.

In contrast to the Proterozoic, Archean rocks are often heavily metamorphized deep-water sediments, such as graywackes, mudstones, volcanic sediments and banded iron formations. Greenstone belts are typical Archean formations, consisting of alternating high- and low-grade metamorphic rocks. The high-grade rocks were derived from volcanic island arcs, while the low-grade metamorphic rocks represent deep-sea sediments eroded from the neighboring island arcs and deposited in a forearc basin. In short, greenstone belts represent sutured protocontinents.[9]

Theoretical green astronomy

"Some of the [green] color [in sediment] may ... come from phytoplankton, tiny plant-like organisms that live in the sun-lit surface waters of the [Earth's] ocean."[10]

Are the Earth auroras due to chlorophyll-containing phytoplankton aloft in the upper atmosphere?

Sources

Chlorophyll is a common source of green on Earth's surface. Malachite is a mineral that occurs in rocks at or near the interface between Earth's atmosphere and crust.

"[T]he typical moonless night-sky optical spectrum for the [Calar Alto observatory] ... shows a strong contamination by different [light] pollution lines, in particular from Mercury lines ... Light pollution arises principally from [tropospheric] scattering of light emitted by sodium and mercury vapour and incandescent street lamps ... the mercury lamps produce [a] narrow [line] at 5461 Å ... the 5460Å Hg line lies in the centre of the y-band of the ubvy Strömgren photometric system, which may affect the programs devoted to the study of stellar populations that use this filter set."[11]

Continua

"[H]igh-resolution (0.1") observations of the Seyfert 2 galaxy NGC 5728 with the Hubble Space Telescope [have produced images in the] green and red continua."[12]

Backgrounds

"[E]xposures were taken through the F492M filter (4906 Å/364 Å) to cover the [O III] λλ4959, 5007 emission. A single exposure was taken through each of the filters F547M (5454 Å/438 Å) and F718M (7160 Å/595 Å) to record the off-band continua."[12]

Heliums

File:Helium spectra.jpg
The spectrum shows the lines in the visible due to emission from elemental helium. Credit: Teravolt.

"C. FRIEDLÄNDER and H. Kayser have independently claimed to have found helium in the [Earth's] atmosphere. On examination of some photographs of the spectrum of neon I have identified six of the principal lines of helium, which thus establishes beyond question the presence of this gas in the air. The amount present in the neon it is, of course, impossible to estimate, but the green line (wave-length 5016 [501.6 nm]) is the brightest, as would be expected from the low pressure of the helium in the neon."[13]

Helium has green lines at 501 and 505 nm, while a line at 493 nm is in the cyan and 587 nm is in the yellow.[14]

Carbons

File:Carbon Spectra.jpg
The spectrum shows the lines in the visible due to emission from elemental carbon. Credit: Teravolt.

Carbon has an emission line in plasmas at 529.053 nm from C VI.[15]

"The [oxygen] green line occurs at 5577.339Å, in the middle of the strong C2 (1,2) band, and thus has seldom been observed."[16]

Nitrogens

File:Nitrogen.Spectrum.Vis.jpg
The spectrum shows the lines in the visible due to emission from elemental nitrogen. Credit: Kurgus.

A nitrogen green emission line occurs in plasmas at 566.934 nm from N VIII.[15]

"[T]he 5198-, 5201-A lines of nitrogen [occur] in the [Earth's] nightglow."[17]

Oxygens

File:Oxygen spectre.jpg
The spectrum shows the lines in the visible due to emission from elemental oxygen. Credit: Teravolt.

"[A]irglow emissions [have been] measured by using vertical-viewing photometers [for the] O(1S) green line at 557.7 nm [with a] background at 566 nm"[18].

The O III emission lines are at 495.9 and 500.7 nm.[12]

Oxygen has an emission line that occurs in plasmas at 527.62 nm from O IV.[15]

"Since the early work of Ångström,* we have the published records of over a hundred investigations on the spectrum, and many others on the origin or other phenomena characteristic of the aurora."[19] "Babcock, using a Fabry and Perot interferometer, determined very accurately the wave-length of the auroral green line 5577. ... After a careful examination of all the results obtained in these reports, we may only say that the exact nature of the cosmical rays, responsible for the aurora, remains a mystery. ... The origin of the most prominent and interesting line of the auroral spectrum, the line 5577, has hitherto remained unexplained. Vegard* has recently obtained a luminescent band from solid nitrogen, that he supposes, under very special conditions, may coincide with the auroral green line. ... spectra of pure helium and of pure oxygen were taken at different pressures and with various excitations, but no trace of 5577 or of any other new lines was obtained. ... Mixtures of helium, oxygen and nitrogen were excited, and it was found that the line 5577 could be photographed on the same plate with the nitrogen band system, thus reproducing in the laboratory practically the entire auroral spectrum. In ... mixtures of neon and oxygen ... neon enhanced the line 5577 in the same manner as helium. ... From Plate 20 it will be seen that all the lines except 5577 have been identified as strong lines in the spectrum of helium, hydrogen, oxygen, or mercury. ... It has been shown that this line must be attributed to some hitherto unknown spectrum of oxygen, and that it is not a limiting member of the ordinary band spectrum of oxygen. It has been observed faintly in highly purified oxygen when currents of high density have been used."[19]

Fluorines

File:Figure-9.-The-light-signature-of-Fluorine.png
This diagram contains the emission and absorption lines for the element fluorine. Credit: Alex Petty.

Fluorine has green emission lines that occur in plasmas at 526.83, 528.56, 529.76 and 530.27 nm from F VI.[15]

Magnesiums

File:Magnesium Spectra.jpg
This image shows the optical emission lines of magnesium. Credit: teravolt.
File:Green sand close up.jpg
This image is a visual close up of green sand which is actually olivine crystals that have been eroded from lava rocks. Credit: Brocken Inaglory.
File:Peridot2.jpg
This is a visual image of a forsterite crystal. Credit: Azuncha.
File:Basanite&Dunite bomb.JPG
This is a small volcanic bomb of (black) basanite with (green) dunite. Credit: B.navez.

At right is a visual close up of green sand which is actually olivine crystals that have been eroded from lava rocks. Some olivine crystals are still inside the lava rock.

Forsterite (Mg2SiO4) is the magnesium rich end-member of the olivine solid solution series.

Forsterite is associated with igneous and metamorphic rocks and has also been found in meteorites. In 2005 it was also found in cometary dust returned by the Stardust probe.[20] In 2011 it was observed as tiny crystals in the dusty clouds of gas around a forming star.[21]

Two polymorphs of forsterite are known: wadsleyite (also orthorhombic) and ringwoodite (isometric). Both are mainly known from meteorites.

At lower right is an image of a small volcanic bomb of (black) basanite with (green) dunite.

Dunite is an igneous, plutonic rock, of ultramafic composition, with coarse-grained or phaneritic texture. The mineral assemblage is greater than 90% olivine, with minor amounts of other minerals such as pyroxene, chromite and pyrope. Dunite is the olivine-rich end-member of the peridotite group of mantle-derived rocks. Dunite and other peridotite rocks are considered the major constituents of the Earth's mantle above a depth of about 400 kilometers. Dunite is rarely found within continental rocks, but where it is found, it typically occurs at the base of ophiolite sequences where slabs of mantle rock from a subduction zone have been thrust onto continental crust by obduction during continental or island arc collisions (orogeny). It is also found in alpine peridotite massifs that represent slivers of sub-continental mantle exposed during collisional orogeny. Dunite typically undergoes retrograde metamorphism in near-surface environments and is altered to serpentinite and soapstone.

Argons

File:Argon Spectrum.jpg
This shows an Argon spectra using a 600lpm diffraction grating. Credit: teravolt.

Argon has three emission lines that occur in an electron cyclotron resonance (ECR) heated plasmas: 497.216, 500.9334, and 506.204 nm from Ar II.[15]

Argon has an emission line occurring in the solar corona at 553.6 nm from Ar X.[22]

Potassiums

File:Potassium spectrum visible.png
This is a potassium emission-line spectrum over the range 400 nm - 700 nm. Credit: McZusatz.

"The group [of potassium lines] at λλ 535, 510, and 495 Å "showed no trace of structure even in an arc of but half an ampere."[23]

Calciums

File:Calcium spectrum visible.png
This is a calcium emission spectrum covering the range: 400 nm - 700 nm. Credit: McZusatz.

Calcium (Ca) has green emission lines at 526.170 nm and 526.556 nm as observed in solar limb faculae.[24]

Titaniums

File:Titanium spectrum visible.png
This is an emission spectrum that covers the visible range: 400 nm - 700 nm. Credit: McZusatz.

Titanium (Ti) has green emission lines at 521.97, 522.268, 522.413, 524.729, and 526.596 nm as observed in solar limb faculae.[24]

Vanadiums

File:Vanadiumoxidationstates.jpg
Oxidation states of vanadium are shown from left +2 (lilac), +3 (green), +4 (blue) and +5 (yellow). Credit: Steffen Kristensen.
File:Vanadium spectrum visible.png
This is an emission spectrum that covers the visible range: 400 nm - 700 nm. Credit: McZusatz.

The chemistry of vanadium is noteworthy for the accessibility of the four adjacent oxidation states 2-5. In aqueous solution the colours are lilac V2+(aq), green V3+(aq), blue VO2+(aq) and, at high pH, yellow VO42-.

Chromiums

File:Chromium spectrum visible.png
This is a visible emission-line spectrum for chromium over the range: 400-700 nm. Credit: McZusatz.

A chromium emission line occurs in plasmas at 520.84 nm from Cr II.[15]

Chromium (Cr) has green emission lines at 522.091, 522.407, 524.336, 524.757, and 526.572 nm as observed in solar limb faculae.[24]

Cr+ has an emission line at 524.677 nm as observed in solar limb faculae.[24]

Irons

File:Fraunhofer lines FR.svg
This digitally altered image shows the Fraunhofer lines. Credit: Saperaud.
File:Iron spectrum visible.png
This is an emission-line spectrum for iron over the visible range: 400-700 nm. Credit: McZusatz.

Fraunhofer's original (1817) designations of absorption lines in the solar spectrum include the E line for neutral iron at 526.96 nm.

Iron (Fe) has green emission lines at 522.319, 522.553, 524.705, 524.911, 525.021, 525.065, 526.262, 526.288, 526.331, 526.368, 526.656, and 526.727 nm as observed in solar limb faculae.[24]

"[G]round-based observations (around the solar limb) [are] of the 530.3 nm coronal line. This green solar corona is emitted by Fe XIV ions ... Green coronal intensities (Fe XIV, 530.3 nm) obtained from different coronal stations vary greatly among themselves."[25]

Fe+ has a green emission line at 526.480 nm as observed in solar limb faculae.[24]

Iron hydride

"Carroll and McCormack (1972) in Dublin reported complex spectra in the blue and green wavelength regions of both FeH and FeD".[26]

"Carroll et al. (1976) detected a number of coincidences between laboratory lines of FeH and weak unidentified solar lines, again in the blue and green wavelength region, in addition to the infrared."[27]

Nickels

File:Nickel spectrum visible.png
This is an emission-line spectrum for nickel over the visible range: 400-700 nm. Credit: McZusatz.

Nickel has an emission line occurring in the solar corona at 511.603 nm from Ni XIII.[22]

Mercuries

File:Mercury Spectra.jpg
This is a visible emission spectrum of mercury (Hg). Credit: teravolt.

"The green mercury line λ 5461 appears to be quite unique in the possession of its satellites."[23]

There are "telluric mercury lines".[28]

There are an additional pair of lines for mercury, one at 496.03 nm in the green and the second at 491.6 nm in the cyan.[29]

Thalliums

File:Thallium spectrum visible.png
This is an emission-line spectrum for thallium over the visible: 400-700 nm. Credit: McZusatz.

There is a green thallium line that shows up in arc spectra using "two to eight amperes at 120 volts, usually between ordinary arc carbons."[23]

Sun

File:Green rim of the setting sun.jpg
The upper rim is green while the lower one is red, as the Sun sets behind the Golden Gate Bridge. Credit: Brocken Inaglory.
File:Green rim and green flashes in SF.jpg
The green rim and flashes of a setting Sun are imaged. Credit: Brocken Inaglory.

As an astronomical object sets or rises in relation to the horizon, the light it emits travels through Earth's atmosphere, which works as a prism separating the light into different colors. The color of the upper rim of an astronomical object could go from green to blue to violet depending on the decrease in concentration of pollutants, as they spread throughout an increasing volume of atmosphere.[30] The lower rim of an astronomical object is always red.
A green rim is very thin, and is difficult or impossible to see with the naked eye. In usual conditions, a green rim of an astronomical object gets fainter, when an astronomical object is very low above the horizon because of atmospheric reddening,[31] but sometimes the conditions are right to see a green rim just above the horizon.

A number of emission lines have been detected in solar limb faculae.[24]

Limb brightness

"An excess brightness [at or near the "edge" of the Sun] can be expected to have a pronounced color dependence, whereas a geometrical oblateness cannot depend on color."[32]

There is "a time varying, excess equatorial brightness"[33]

Def. "the radial distance q from the Sun's center such that the following finite Fourier transform is zero:

<math>F(G; q, a) = \int_{-1/2}^{+1/2} G(q + a \sin \pi s) \cos(2\pi s) ds,</math>

where s is a dummy variable, G is the observed solar intensity as a function of the radius, and the parameter a determines the extent of the solar limb used"[33] is called the solar edge.

"When F(G; q, a) = 0, the a dependence of q can be used to choose different points as the edge."[33]

"The wavelength dependence of the facular excess brightness ... is ... normalized to unity for the [intensity of transmission of] the green filter [(525±70 nm)]. ... An unweighted least-squares fit to the five points [(five filter transmission intensities)] gives

R(λ) = -0.0568 ± 0.0634 + (0.582 ± 0.038)λ-1"[32].

"[T]he blue contrast lies about 1 σ above the λ-1 curve. ... [I]f real, [this] may be the result of an increasing opacity in the blue and an increasing ΔT/T in the upper layers of the photosphere. An increasing opacity in the blue may be due to line haze since the blue filter has a 78 nm width."[32]

Solar coronal clouds

File:LASCO C1 coronagraph of solar corona.png
A picture of the solar corona taken with the LASCO C1 coronagraph. The image is color coded for the doppler shift of the FeXIV 530.8 nm line. Credit: NASA and NRL.

In the image at right the iron (Fe XIV) green line is followed by doppler imaging to show associated relative coronal plasma velocity towards (-7 km/s side) and away from (+7 km/s side) the large angle spectrometric coronagraph LASCO satellite camera.

Coronal clouds

"The striking absence of green emission above both polar regions at activity minimum led Waldmeier (1957) to use the German term 'Koronalöcher', ie, coronal holes."[34] "Here we restrict ourselves to a qualitative study of large scale structures of the green emission line corona."[34]

The image of the solar coronal cloud at above right shows the polar coronal holes.

Solar cycles

"The course of the solar cycle can ... be tracked by changes in the line blanketing. The line-blanketing variations during solar cycle 21 have been measured by Mitchell & Livingston (1991) who found that in disk-integrated spectra the spectral lines in the 500 to 560 nm range are on average 1.4% shallower and have 0.8% smaller equivalent widths at solar maximum than at solar minimum."[35]

Venus

"Venus at times shows [the oxygen] green line emission with an intensity equal to terrestrial values [Slanger et al., 2001]. Furthermore, the intensity is quite variable, as is true for the much stronger O2( a-X) 1.27 μ emission."[36]

"In 1999, observations of the Venus nightglow with the Keck I telescope showed that the 5577 Å oxygen green line was a significant feature, comparable in intensity to the terrestrial green line. Subsequent measurements have been carried out at the Apache Point Observatory (APO) and again at Keck I, confirming the presence of the line with substantially varying intensity."[37]

"Ground-based studies suggest that the [557.7 nm oxygen green line] emission is correlated with the solar cycle."[38]

Airglows

File:Love and Joy for the New Year - NASA Earth Observatory.jpg
In this International Space Station image, you can see green and yellow airglow paralleling the Earth’s horizon line (or limb) before it is overwhelmed by the light of the rising Sun. Credit: NASA Earth Observatory.

In the International Space Station image at right, you can "see green and yellow airglow paralleling the Earth’s horizon line (or limb) before it is overwhelmed by the light of the rising Sun. Airglow is the emission of light by atoms and molecules in the upper atmosphere after they are excited by ultraviolet radiation. ... Astronaut photograph ISS030-E-015491 was acquired on December 22, 2011, with a Nikon digital camera, and is provided by the ISS Crew Earth Observations experiment and Image Science & Analysis Laboratory, Johnson Space Center."[39]

Airglow (also called nightglow) is the very weak emission of light by a planetary atmosphere. In the case of Earth's atmosphere, this optical phenomenon causes the night sky to never be completely dark (even after the effects of starlight and diffused sunlight from the far side are removed).

Airglow is caused by various processes in the upper atmosphere, such as the recombination of ions which were photoionized by the sun during the day, luminescence caused by cosmic rays striking the upper atmosphere, and chemiluminescence caused mainly by oxygen and nitrogen reacting with hydroxyl ions at heights of a few hundred kilometers. It is not noticeable during the daytime because of the scattered light from the Sun.

Auroras

File:Aurora Borealis.jpg
Aurora Borealis is photographed by NASA astronaut Donald R. Pettit. Credit: NASA.
File:Red and green aurora.jpg
This image shows both red and green aurora over Fairbanks, Alaska. Credit: Photograped by Brocken Inaglory.
File:ISS003-ESC-6152.jpg
"As geomagnetic storms cause beautiful displays of aurora across the United States ..., astronauts onboard the International Space Station also have the opportunity to take a look. Credit: Earth Sciences and Image Analysis Laboratory, Johnson Space Center, NASA.

At right is a natural color photograph of the Aurora Borealis or northern lights and the Manicouagan Impact Crater reservoir (foreground) in Quebec, Canada. They are featured in this photograph taken by astronaut Donald R. Pettit, Expedition Six NASA ISS science officer, on board the International Space Station (ISS).

The second image contains both red and green auroral components.

Auroras result from emissions of photons in the Earth's upper atmosphere, above 80 km (50 mi), from ionized nitrogen atoms regaining an electron, and oxygen and nitrogen atoms returning from an excited state to ground state.[40] They are ionized or excited by the collision of solar wind and magnetospheric particles being funneled down and accelerated along the Earth's magnetic field lines; excitation energy is lost by the emission of a photon, or by collision with another atom or molecule:

oxygen emissions
green or brownish-red, depending on the amount of energy absorbed
nitrogen emissions
blue or red; blue if the atom regains an electron after it has been ionized, red if returning to ground state from an excited state.

Oxygen is unusual in terms of its return to ground state: it can take three quarters of a second to emit green light and up to two minutes to emit red. Collisions with other atoms or molecules will absorb the excitation energy and prevent emission. Because the very top of the atmosphere has a higher percentage of oxygen and is sparsely distributed such collisions are rare enough to allow time for oxygen to emit red. Collisions become more frequent progressing down into the atmosphere, so that red emissions do not have time to happen, and eventually even green light emissions are prevented.

This is why there is a color differential with altitude; at high altitude oxygen red dominates, then oxygen green and nitrogen blue/red, then finally nitrogen blue/red when collisions prevent oxygen from emitting anything. Green is the most common of all auroras. Behind it is pink, a mixture of light green and red, followed by pure red, yellow (a mixture of red and green), and lastly, pure blue.

The third image is taken from the International Space Station. "Green colors of the aurora are dominant in this image captured by a digital still camera on October 4, 2001. Auroras are caused when high-energy electrons pour down from the Earth’s magnetosphere and collide with atoms. Green aurora occurs from about 100 km to 250 km altitude and is caused by the emission of 5577 Angstrom wavelength light from oxygen atoms. The light is emitted when the atoms return to their original unexcited state."[41]

Coral reefs

File:Coral Reefs in the Persian Gulf.jpg
Pale green when viewed through the water of the Persian Gulf, coral reefs fringe the shoreline and islands of the United Arab Emirates in this Landsat satellite image. Credit: Jesse Allen, NASA, United States Geological Survey.

"Pale green when viewed through the water of the Persian Gulf, coral reefs fringe the shoreline and islands of the United Arab Emirates in this Landsat satellite image. The reefs are more extensive than the islands they surround in many cases. The natural-color image was made from data collected by the Landsat 7 satellite on July 3 and July 10, 2002."[42]

"The reefs are an oasis of biological diversity, and they are among the most extensive reefs in the Gulf. Starting in January 2005, a division of the World Wildlife Federation in Abu Dhabi used Landsat images such as this one to create the first map of coral reefs off Abu Dhabi in the United Arab Emirates and Qatar. The map will help resource managers assess the quality and quantity of marine resources in the region and to make decisions that protect them."[42]

Forests

File:Spring Paints the Piedmont Green - NASA Earth Observatory.jpg
From a satellite's vantage point, the wash of green that appears across the forests of the eastern United States is one of the most noticeable signs of winter's passing. Credit: Jeff Schmaltz, LANCE/EOSDIS MODIS Rapid Response and NASA's Earth Observatory.

"At ground level, tulips and daffodils mark the arrival of spring. But from a satellite's vantage point, the wash of green that appears across the forests of the eastern United States is one of the most noticeable signs of winter's passing. The Moderate Resolution Imaging Spectroradiometer (MODIS) instrument on NASA's Aqua satellite captured this view of spring greening on April 7, 2012, an unusually cloud free day."[43]

"The deepest greens of lush foliage are most visible throughout the Piedmont, a forested plateau between the Appalachians and the lower elevation plains along the Atlantic coast. The Appalachians appear brown because cooler temperatures at higher elevations cause a lag in the greening. In this case, the trees at the higher elevations were likely still in bloom and hadn't started to produce leaves. The speckles of tan throughout the coastal plain are farmlands, where fields often stay bare or filled with dry crop stubble until late spring planting."[43]

Hydrogen sulfide emissions

File:Hydrogen Sulfide Emissions off of Africa - NASA Earth Observatory.jpg
Pale green patterns tinted the water along the Namibian coast in late February 2012. Credit: Jeff Schmaltz, LANCE/EOSDIS MODIS Rapid Response Team at NASA GSFC.

In the image at right, "[p]ale green patterns tinted the water along the Namibian coast in late February 2012. But unlike other bright hues that occasionally show up in the ocean, these colors didn’t result from a phytoplankton bloom.[44]

"[H]ydrogen sulfide gas is emitted periodically along the Namibian coast. Ocean currents carry oxygen-poor water to the region, and chemical and biological processes can deplete what little oxygen is available. The sediments in the local seafloor are also rich with organic matter. When organic matter decays in an oxygen-poor environment, hydrogen sulfide emissions can result."[44]

"The Moderate Resolution Imaging Spectroradiometer (MODIS) on NASA’s Terra satellite captured this natural-color image on February 29, 2012. The pale-hued surface waters snake along the shore of the Namib Desert, stretching roughly 150 kilometers (90 miles)."[44]

"The milky-green colors along Namibia’s coast indicate high concentrations of sulfur and low concentrations of oxygen. Episodes like this aren’t just colorful, they are actually toxic to local marine organisms. Fish die in the low-oxygen water; however, what is deadly for the fish can be good for birds that feed on their carcasses. Likewise, lobsters crawling onto shore to escape the toxic seawater can make meals for locals. And some species of foraminifera—tiny shelled marine organisms—actually thrive in the oxygen-poor sea floor sediments off the Namibian coast."[44]

Phytoplanktons

File:Phytoplankton Bloom in the Arabian Sea.jpg
Phytoplankton swirling across the Arabian Sea are drawn into thin green ribbons by turbulent eddies in this image. Credit: Norman Kuring, NASA Earth Observatory, and GSFC Oceancolor Team.

"Phytoplankton swirled across the Arabian Sea on February 18, 2010, drawn into thin green ribbons by turbulent eddies. The bloom stretches from the shores of Pakistan (top) to the coast of Oman (lower left). The washed out appearance at the upper left of the image is due to sunglint, which is the mirror–like reflection of sunlight off the water. Some of the brightness may be caused by blowing dust. Pakistan’s coastal waters are tinged blue as well as green. The color may result from numerous influences, including phytoplankton and sediment."[45]

"The phytoplankton blooms in the northern Arabian Sea are strongly influenced by the seasonal wind shifts (the monsoon) that dominate the area’s climate. Because the Arabian Sea is landlocked to the north, it is less influenced by large-scale ocean circulation and more strongly influenced by the monsoon winds. Large blooms of phytoplankton occur in the summer, when strong southwesterly winds blow from the ocean toward land, mixing the water. Blooms also happen in the winter, however, when northeast winds blow offshore."[45]

Sediments

File:Sediment off the Yucatan Peninsula.jpg
As sediment disperses into the Gulf of Mexico, the water fades to green and then black. Credit: NASA image courtesy the MODIS Rapid Response Team at NASA GSFC.

"A burst of color lit the shallow waters of the Gulf of Mexico off the Yucatan Peninsula on December 14, 2008, when the Moderate Resolution Imaging Spectroradiometer (MODIS) on NASA’s Terra satellite captured this image [at right]. The swirls of tan, green, blue, and white are most likely sediment in the water. The sediment scatters light, giving the water its color. The sediment comes from two sources: the land and the sea floor. Some of the color may also come from phytoplankton, tiny plant-like organisms that live in the sun-lit surface waters of the ocean."[10]

"Near the shore, the water is tan where rivers carry dirt from land to the ocean. As the sediment disperses, the water fades to green and then black. To the north, the water is more blue and white than tan and green. In these regions, the sediment has likely come from the sea floor. Made up of chalky white calcium carbonate from shell-building marine life like coral, sea floor sediment gives the water a white or bright blue color. The sediment was probably brought to the surface in shallow waters by strong waves. A few days before the image was taken, strong winds churned the Gulf. The blue-green cloud in this image roughly matches the extent of the shallow continental shelf west of the peninsula."[10]

Volcanoes

File:Tremors, Eruption at El Hierro Subsiding - NASA Earth Observatory.jpg
The Advanced Land Imager (ALI) on NASA’s Earth Observing-1 (EO-1) satellite captured this natural-color image of the island and its offshore eruption on December 16, 2011. Credit: esse Allen and Robert Simmon, NASA EO-1 team, and NASA's Earth Observatory.

"After two months of eruptions and six months of quakes and tremors, the underwater volcano off the coast of El Hierro appears to be quieting down. A mass of new crust has been building by tens to perhaps a hundred meters, but it is still nowhere near rising above the ocean surface and extending the Canary Island chain."[46]

"The Advanced Land Imager (ALI) on NASA’s Earth Observing-1 (EO-1) satellite captured this natural-color image of the island and its offshore eruption on December 16, 2011. The eruption is located about one kilometer south-southwest of the town of La Restinga."[46]

"The milky green swirls in the Atlantic Ocean are a volcanic brew of steaming lava fragments, bits of rock, heated gas, and other debris that are carried to the west and north by currents. Meanwhile, tremors have been occurring on the north side of the island at depths of 17 to 23 kilometers. Red circles mark the locations of three tremors reported on December 18. The white puffs over the island are clouds, not volcanic emissions."[46]

"El Hierro is a shield volcano growing along the southwestern edge of the ancient El Golfo volcano. El Golfo collapsed about 130,000 years ago, but the ruddy brown and rugged terrain of the island attests to many years of volcanism in the area. El Hierro last erupted in 1793, according to some historical records, and the area has the greatest concentration of young vents in the Canary Islands."[46]

Greenstones

File:Temagami greenstone belt pillow lava.jpg
This image shows weathered Precambrian pillow lava in the Temagami greenstone belt of the Canadian Shield in Eastern Canada. Credit: Black Tusk.

Greenstone belts are zones of variably metamorphosed mafic to ultramafic volcanic sequences with associated sedimentary rocks that occur within Archaean and Proterozoic cratons between granite and gneiss bodies.

The name comes from the green hue imparted by the colour of the metamorphic minerals within the mafic rocks. Chlorite, actinolite and other green amphiboles are the typical green minerals.

A greenstone belt is typically several dozens to several thousand kilometres long and although composed of a great variety of individual rock units, is considered a 'stratigraphic grouping' in its own right, at least on continental scales.

"Greenstone belts" are distributed throughout geological history from the Phanerozoic Franciscan belts of California where blueschist, whiteschist and greenschist facies are recognised, through to the Palaeozoic greenstone belts of the Lachlan Fold Belt, Eastern Australia, and a multitude of Proterozoic and Archaean examples.

Archaean greenstones are found in the Slave craton, northern Canada, Pilbara craton and Yilgarn Craton, Western Australia, Gawler Craton in South Australia, and in the Wyoming Craton in the US. Examples are found in South and Eastern Africa, namely the Kaapvaal craton and also in the cratonic core of Madagascar, as well as West Africa and Brazil, northern Scandinavia and the Kola Peninsula (see Baltic Shield).

Phanerozoic ophiolite belts and greenstone belts occur in the Franciscan Complex of south-western North America, within the Lachlan Fold Belt, the Gympie Terrane of Eastern Australia, the ophiolite belts of Oman and around the Guiana Shield.

Meteors

Template:Radiation astronomy/Meteors "[T]he first identification of the OI 5577 Å green line [in a meteor spectrum was] (Halliday, 1959). ... the emission is a major contributor to wake spectra. ... emission occurs only from the upper part of the meteor light curve being generally confined to the height interval 100 - 110 km, and that its occurrence is rare in meteors whose velocities are less than about 40 kms-1."[47]

Meteorites

File:Moldavite Besednice.jpg
This image shows Moldavite from Besednice, Bohemia. Credit: H. Raab (Vesta).

Moldavite is an olive-green or dull greenish vitreous substance possibly formed by a meteorite impact. It is one kind of tektite.

Because of their difficult fusibility, extremely low water content, and its chemical composition, the current overwhelming consensus among Earth scientists is that moldavites were formed 15 million years ago during the impact of a giant meteorite in present-day Nördlinger Ries. Splatters of material that was melted by the impact cooled while they were actually airborne and most fell in central Bohemia—traversed by [the] Vltava river. Currently, moldavites have been found in [an] area that includes southern Bohemia, western Moravia, the Cheb Basin (northwest Bohemia), Lusatia (Germany), and Waldviertel (Austria).[48] Isotope analysis of samples of moldavites have shown a beryllium-10 isotope composition similar to the composition of Australasian tektites (Australites)and Ivory Coast tektites (Ivorites). Their similarity in beryllium-10 isotope composition indicates that moldavites, Australites, and Ivorites consist of near surface and loosely consolidated terrestrial sediments melted by hypervelocity impacts.[49]

Zodiacal clouds

File:EZL-Faulkes.jpg
The Zodiacal Light is over the Faulkes Telescope, Haleakala, Maui. Credit: 808caver.

The Zodiacal light is a faint, roughly triangular, diffuse white glow seen in the night sky that appears to extend up from the vicinity of the Sun along the ecliptic or zodiac.[50] Caused by sunlight scattered by space dust in the zodiacal cloud, the zodiacal light covers the entire sky, being responsible for major part[51] of the total skylight on a moonless night.

"[T]he distribution [is] over the sky of the linearly polarized component of the zodiacal light at λλ5080 Å and 5300 Å. In the antisolar hemisphere the distribution shows three clear effects: (i) There is a slight (~ 3°) clockwise overall rotation of the orientation of the polarization plane relative to that calculated for spherical interplanetary dust; (ii) near the antisolar point the polarization plane tends to be normal to the ecliptic, but the distribution of the polarization vectors is asymmetric with respect to the ecliptic plane ("north-south asymmetry"); (iii) significant day-to-day and month-to-month changes occur in the zodiacal light but without clear trends."[52]

Earthshines

File:Earthshine Karonen.jpg
This visual image of the Moon shows sunlight reflected as the bright crescent and earthlight reflected from the dark side. Credit: Ilmari Karonen.

Def. sunlight reflected from the earth's surface is called earthlight.

Def. reflected earthlight visible on the Moon's night side is called earthshine.

"The search for life on extrasolar planets" requires a test of vegetation detectability from a single dot source.[53]

"The earthshine, or ashen light, is the glow of the dark part of the lunar disk visible to a night-time observer. ... [T]he light rays coming from different parts of the Earth are mixed together in the ashen light and mimic the Earth as a single dot."[53]

"[T]he vegetation spectrum which is unequivocal ... presents a bump at 0.5 µ in the green wavelength range, which implies that plants appear green".[53]

Moon

File:Earth's Moon.jpg
This colour mosaic was assembled from 18 images taken by Galileo's imaging system through a green filter. Credit: NASA/JPL/USGS.

"During its mission, the Galileo spacecraft returned a number of images of Earth's only natural satellite. Galileo surveyed the moon on Dec. 7, 1992, on its way to explore the Jupiter system in 1995-1997."[54]

"This color mosaic was assembled from 18 images taken by Galileo's imaging system through a green filter. On the upperleft is the dark, lava-filled Mare Imbrium, Mare Serenitatis (middle left), Mare Tranquillitatis (lower left), and Mare Crisium, the dark circular feature toward the bottom of the mosaic. Also visible in this view are the dark lava plains of the Marginis and Smythii Basins at the lower right. The Humboldtianum Basin, a 400-mile impact structure partly filled with dark volcanic deposits, is seen at the center of the image."[54]

Mars

"[O]bservations from the SPICAM spectrometer on the Mars Express spacecraft ... of the [oxygen] green line [are compared with simulations]".[55]

"O(1S) [green line] emission is seen in the atmospheres of all three terrestrial planets -Venus, Earth, Mars."[56]

Asteroids

File:PIA00136.jpg
This is the first image of the asteroid Ida using the green 559 nm filter onboard the Galileo spacecraft. Credit: NASA/JPL.

"This is the first full picture showing both asteroid 243 Ida and its newly discovered moon to be transmitted to Earth from the National Aeronautics and Space Administration's (NASA's) Galileo spacecraft--the first conclusive evidence that natural satellites of asteroids exist. Ida, the large object, is about 56 kilometers (35 miles) long. Ida's natural satellite is the small object to the right. This portrait was taken by Galileo's charge-coupled device (CCD) camera on August 28, 1993, about 14 minutes before the Jupiter-bound spacecraft's closest approach to the asteroid, from a range of 10,870 kilometers (6,755 miles). Ida is a heavily cratered, irregularly shaped asteroid in the main asteroid belt between Mars and Jupiter--the 243rd asteroid to be discovered since the first was found at the beginning of the 19th century. Ida is a member of a group of asteroids called the Koronis family. The small satellite, which is about 1.5 kilometers (1 mile) across in this view, has yet to be given a name by astronomers. It has been provisionally designated '1993 (243) 1' by the International Astronomical Union. ('1993' denotes the year the picture was taken, '243' the asteroid number and '1' the fact that it is the first moon of Ida to be found.) Although appearing to be 'next' to Ida, the satellite is actually in the foreground, slightly closer to the spacecraft than Ida is. Combining this image with data from Galileo's near-infrared mapping spectrometer, the science team estimates that the satellite is about 100 kilometers (60 miles) away from the center of Ida. This image, which was taken through a green filter, is one of a six-frame series using different color filters. The spatial resolution in this image is about 100 meters (330 feet) per pixel."[57]

Io

File:Io Aurorae color.jpg
Gases above Io's surface produced a ghostly glow that could be seen at visible wavelengths (red, green, and violet). Credit: NASA/JPL/University of Arizona.

At right is an "eerie view of Jupiter's moon Io in eclipse ... acquired by NASA's Galileo spacecraft while the moon was in Jupiter's shadow. Gases above the satellite's surface produced a ghostly glow that could be seen at visible wavelengths (red, green, and violet). The vivid colors, caused by collisions between Io's atmospheric gases and energetic charged particles trapped in Jupiter's magnetic field, had not previously been observed. The green and red emissions are probably produced by mechanisms similar to those in Earth's polar regions that produce the aurora, or northern and southern lights. Bright blue glows mark the sites of dense plumes of volcanic vapor, and may be places where Io is electrically connected to Jupiter."[58]

"North is to the top of the picture, and Jupiter is towards the right. The resolution is 13.5 kilometers (8 miles) per picture element. The images were taken on May 31, 1998 at a range of 1.3 million kilometers (800,000 miles) by Galileo's onboard solid state imaging camera system during the spacecraft's 15th orbit of Jupiter."[58]

Enceladus

File:Spray Above Enceladus PIA07760.jpg
The image at the left was taken in visible green light. The image at the right has been color-coded to make faint signals in the plume more apparent. Credit: NASA.

At right is a pair of images showing "[a] fine spray of small, icy particles emanating from the warm, geologically unique province surrounding the south pole of Saturn’s moon Enceladus[. It] was observed in a Cassini narrow-angle camera image of the crescent moon taken on Jan. 16, 2005. Taken from a high-phase angle of 148 degrees -- a viewing geometry in which small particles become much easier to see -- the plume of material becomes more apparent in images processed to enhance faint signals [right image of the pair]."[59]

"Though the measurements of particle abundance are more certain within 100 kilometers (60 miles) of the surface, the values measured there are roughly consistent with the abundance of water ice particles measured by other Cassini instruments (reported in September, 2005) at altitudes as high as 400 kilometers (250 miles) above the surface."[59]

"The image at the left was taken in visible green light. A dark mask was applied to the moon's bright limb in order to make the plume feature easier to see."[59]

"The image at the right has been color-coded to make faint signals in the plume more apparent. Images of other satellites (such as Tethys and Mimas) taken in the last 10 months from similar lighting and viewing geometries, and with identical camera parameters as this one, were closely examined to demonstrate that the plume towering above Enceladus' south pole is real and not a camera artifact. The images were acquired at a distance of about 209,400 kilometers (130,100 miles) from Enceladus. Image scale is about 1 kilometer (0.6 mile) per pixel."[59]

Iapetus

File:A Moon with Two Dark Sides.jpg
This is an image of Iapetus in polarized green light with the Cassini spacecraft narrow-angle camera. Credit: NASA/JPL/Space Science Institute.
File:PIA11689 Iapetus color dichotomy.jpg
The color seen here on this false-color composite of Iapetus is similar to that produced in (red, green and blue) natural color views. Credit: NASA / JPL / Space Science Institute.

"Although it is no longer uncharted land, the origin of the dark territory of Cassini Regio on Iapetus remains a mystery. Also puzzling is the equatorial ridge that bisects this terrain, and how it fits into the story of the moon's strange brightness dichotomy. The ridge is seen here, curving along the lower left edge of Iapetus. The view looks down onto the northern hemisphere of Iapetus (1,468 kilometers, or 912 miles across), and shows terrain on the moon's leading hemisphere."[60]

"The image was taken in polarized green light with the Cassini spacecraft narrow-angle camera on Nov. 12, 2005 at a distance of approximately 417,000 kilometers (259,000 miles) from Iapetus and at a Sun-Iapetus-spacecraft, or phase, angle of 95 degrees. Image scale is about 2 kilometers (1 mile) per pixel."[60]

In the second image are three "different false-color views of Saturn's moon Iapetus show the boundary of the global "color dichotomy" on the hemisphere of this moon facing away from Saturn. The "color dichotomy," which has been detected in images from the Cassini imaging team, is a second global pattern found on Iapetus besides the well-known global brightness dichotomy."[61]

"This image consists of three panels, each of which was contrast-enhanced in different ways to bring out surface features. Minimal enhancement was applied to the image on the left panel while those on the middle and right panels were enhanced more (with contrast increased by factors of two and four, respectively), making them appear brighter and overexposed."[61]

"In the case of the color dichotomy seen here, its boundary is quite well correlated with the boundary between the moon's leading and the trailing hemispheres. At near-infrared wavelengths, the bright terrain on the leading side is redder than on the trailing side. This pattern is visible in the panel on the left, which uses normal contrast enhancement. The characteristic reddish distribution also appears on the dark material, as seen in the middle and right-hand panels that have been adjusted with even higher contrast. Indeed, the otherwise uniformly dark material shows different color hues, depending on whether the viewer looks at the leading vs. the trailing side."[61]

Tethys

File:Tethys Cassini green large.jpg
This image of Tethys is taken in visible green light with the Cassini spacecraft narrow-angle camera on September 14, 2011. Credit: NASA Cassini.

At right is an image of Saturn's moon Tethys in green light.

"On the top left of the image there is huge Odysseus Crater. See PIA07693 for a closer view. On the bottom right there is Ithaca Chasma, a series of scarps that runs north-south across the moon for more than 620 miles (1,000 kilometers). North on Tethys is up and rotated 25 degrees to the right."[62]

"This view looks toward the area between the leading hemisphere and Saturn-facing side of Tethys (660 miles, or 1,062 kilometers across)."[62]

"The image was taken in visible green light with the Cassini spacecraft narrow-angle camera on Sept. 14, 2011. The view was acquired at a distance of approximately 178,000 miles kilometers (287,000) from Tethys and at a Sun-Tethys-spacecraft, or phase, angle of 11 degrees. Image scale is about 1 mile (2 kilometers) per pixel."[62]

Uranus

File:Uranus in Green.jpg
This is a visual image of Uranus in the green. Credit: Heidi Hammel (Massachusetts Institute of Technology), NASA.

"Spring has finally come to the northern hemisphere of Uranus. The newest images, both the visible-wavelength ones described here and those taken a few days earlier with the Near Infrared and Multi-Object Spectrometer (NICMOS) by Erich Karkoschka (University of Arizona), show a planet with banded structure and detectable clouds."[63]

"The "aqua" image (on the right) is taken at 5,470 Angstroms, which is near the human eye's peak response to wavelength. Color has been added to the image to show what a person on a spacecraft near Uranus might see. Little structure is evident at this wavelength, though with image-processing techniques, a small cloud can be seen near the planet's northern limb (rightmost edge)."[63]

Comets

File:Rosetta.jpg
This is a 3D model of the Rosetta Spacecraft. The individual scientific payloads are highlighted in different colours. Credit: IanShazell.

For elongated dust particles in cometary comas an investigation is performed at 535.0 nm (green) and 627.4 nm (red) peak transmission wavelengths of the Rosetta spacecraft's OSIRIS Wide Angle Camera broadband green and red filters, respectively.[64] "In the green, the polarization of the pure silicate composition qualitatively appears a better fit to the shape of the observed polarization curves".[64] "[B]ut they are characterized by a high albedo."[64] The silicates used to model the cometary coma dust are olivene (Mg-rich is green) and the pyroxene, enstatite.[64]

"[U]nequivocal detections [occurred at McDonald Observatory on 10 nights from 25 June through 17 July 2000] of the O (1S) and O (1D) metastable lines in emission in the cometary [Comet C/1999 S4 (LINEAR)] spectrum. These lines are well separated from any telluric or cometary emission features."[65]

"[T]he presence of the [oxygen] green line can still be questioned, unless the 2972 Å trans-auroral line [1S - 3P] is detected (Herbig, 1976)."[28] "The transitions involved (allowed and forbidden) in the spectrum of the oxygen atoms in a cometary atmosphere" are 557.7 nm, 630.0 and 636.4 nm, 295.8 and 297.2 nm, 98.9 nm (a triplet), 799.0 nm, 844.7 nm, and 1304 nm (a triplet), 102.7 nm (a triplet) and 1128.7 nm.[28]

"When the green line is overwhelming (in faint comets like Encke), this emission is mainly due to the airglow, the red airglow emission being quenched and consequently weaker than the green."[28]

"The measured intensity on 10 January 1980, when the comet was 0.71 a.u. from the Sun and 0.615 a.u. from the Earth, is 30±15 Rayleighs.[28]

Protoplanetary disks

"Seven papers in [the journal] Science [...] (December 2006) discuss details of the sample analysis. Among their findings are discoveries of a wide range of organic compounds, including two that contain biologically usable nitrogen. Indigenous aliphatic hydrocarbons were found with longer chain lengths than those observed in the diffuse interstellar medium. The Stardust samples contain abundant amorphous silicates in addition to crystalline silicates such as olivine and pyroxene. The presence of crystalline silicates in Wild 2 is consistent with mixing of solar system and interstellar matter, something which had been deduced spectroscopically before (see quote above). No hydrous silicates or carbonate minerals were detected, which suggests a lack of aqueous processing of Wild 2 dust. Very few pure carbon (CHON) particles were found in the samples returned."[66]

Ghost Head Nebula

File:NGC2080.jpg
This is an image of NGC 2080, the Ghost Head Nebula. Credit: NASA, ESA and Mohammad Heydari-Malayeri (Observatoire de Paris, France).

At right is a Hubble Space Telescope image of the Ghost Head Nebula. "This nebula is one of a chain of star-forming regions lying south of the 30 Doradus nebula in the Large Magellanic Cloud. The red and blue light comes from regions of hydrogen gas heated by nearby stars. The green light comes from glowing oxygen, illuminated by the energy of a stellar wind. The white center shows a core of hot, massive stars."[67]

Stars

File:The bright star Alpha Centauri and its surroundings.jpg
This is a wide-field view of the sky around the bright G2V star Alpha Centauri A, created from photographic images forming part of the Digitized Sky Survey 2. Credit: ESO/Digitized Sky Survey 2.

Alpha Centauri A is the principal member, or primary, of the binary system, being slightly larger and more luminous [151.9% the luminosity of the Sun] than the Sun. It is a solar-like main sequence star with a similar yellowish color,[68] [surface temperature of 5790 K][69] whose stellar classification is spectral type G2 V.[70] From the determined mutual orbital parameters, Alpha Centauri A is about 10% more massive than the Sun, with a radius about 23% larger.[69] The projected rotational velocity ( v·sin i ) of this star is 2.7 ± 0.7 km·s−1, resulting in an estimated rotational period of 22 days,[71] which gives it a slightly faster rotational period than the Sun's 25 days.

Capella B

File:Capella B Aladin.jpg
This is an optical image of Capella B. Credit: Aladin at SIMBAD.

Capella B, the image at right, has a surface temperature of approximately 5700 K, a radius of approximately 9 solar radii, a mass of approximately 2.6 solar masses, and a luminosity, again measured over all wavelengths, approximately 78 times that of the Sun.[72]

Capella B is spectral type G0III star and an X-ray source from the catalog [FS2003] 0255 by ROSAT. Its surface temperature has an uncertainty of 100 K. It is part of a binary star with Capella A a G8III. From SIMBAD, the orbital period is 104.0217 d with an eccentricity is 0.001 and inclination of 137.2° to the line of sight.

Although the binary star Capella is not an eclipsing binary, it is a RS Canum Venaticorum variable. These are close binary stars[73] having active chromospheres which can cause large stellar spots. These spots are believed to cause variations in their observed luminosity. Systems can exhibit variations on timescales of years due to variation in the spot surface coverage fraction, as well as periodic variations which are, in general, close to the orbital period of the binary system. Typical brightness fluctuation is around 0.2 magnitudes.

Galaxies

File:Ngc5728-hst-RmirvisGB330.jpg
This is a visual astronomy image of NGC 5728 from the Hubble Space Telescope. Credit: Fabian RRRR.
File:HST NGC 5728 Green Continuum.jpg
This is a Hubble Space Telescope image of the green continuum in the nuclear region of NGC 5728. The observation was made on September 4, 1992, using the Planetary Camera with appropriate filters. Credit: A. S. Wilson, J. A. Braatz, T. M. Heckman, J. H. Krolik, & G. K. Miley/NASA-Hubble.
File:HST NGC 5728 -O III- emission-line image.jpg
This is a Hubble Space Telescope image of the green [O III] emission line in the active galactic nuclear region of NGC 5728. The observation was made on September 4, 1992, using the Planetary Camera with appropriate filters. Credit: A. S. Wilson, J. A. Braatz, T. M. Heckman, J. H. Krolik, & G. K. Miley/NASA-Hubble.

"[H]igh-resolution (0.1") observations of the Seyfert 2 galaxy NGC 5728 with the Hubble Space Telescope",[12] in the images at right, show the full color of the galactic nucleus and a green continuum image of the active galactic nucleus.

At left is a Hubble Space Telescope image in the light of the green [O III] emission line in the active galactic nuclear region of NGC 5728. "The emission-line images reveal a spectacular biconical structure with overall extent [of] 1.8 kpc. The two cones share a common axis and apex. The cones' axis coincides to within ≃ 3° with that of the one-sided nuclear radio continuum emission but is not aligned with the rotation axis of the galaxy disk."[12]

These "ionization cones" are "conical regions of high-excitation emission-line gas extending from an active nucleus. ... The generally accepted interpretation is that partially collimated ionizing radiation shines out from the nucleus and ionizes gas in its vicinity. ... two exposures were taken through the F492M filter (4906 Å/364 Å) to cover the [O III] λλ4959, 5007 emission. ... The high-excitation gas can be traced to ≃ 8.5" (1.6 kpc) from the apex in the SE cone, but only to 1.5" (270 pc) in the NW one."[12]

Holocene

File:La Tombe de Horemheb cropped.jpg
The gods Osiris, Anubis, and Horus are shown from a tomb painting. Credit: A. Parrot.

The Holocene starts at ~11,700 b2k and extends to the present.

Osiris is the mythological father of the god Horus, whose conception is described in the Osiris myth, a central myth in ancient Egyptian belief. The myth described Osiris as having been killed by his brother Seth, who wanted Osiris' throne. Isis joined the fragmented pieces of Osiris, but the only body part missing was the phallus. Isis fashioned a golden phallus, and briefly brought Osiris back to life by use of a spell that she learned from her father. This spell gave her time to become pregnant by Osiris before he again died. Isis later gave birth to Horus. As such, since Horus was born after Osiris' resurrection, Horus became thought of as a representation of new beginnings and the vanquisher of the evil Set.

"The Phoenician El - Saturn - has four eyes, as does the Orphic Kronos (Saturn)."[74]

"The Chinese Yellow Emperor Huang-ti--identified as Saturn--is also four-eyed.74"[74]

"Osiris, as the Ram of Mendes, is the god of "four faces on one neck."62"[74]

Temperatures

<math>\frac{c2}{\lambda T}\frac{1}{ e^{\frac{c2}{\lambda T}} - 1}e^{\frac{c2}{\lambda T}} - 5 = 0.</math>

c2 = 1.438833 cm K may be used to approximate a pair such as (495 nm, 6100 K).

For a simple calculator, y=(1.48833/(0.0000570*5260))*exp(1.48833/(0.0000570*5260))/(exp(1.48833/(0.0000570*5260))-1)-5, followed by print y, yields a value close to zero (-1.006663E-03). The closer to zero the more accurate the estimate.

Although Planck's equation is not an exact fit to a star's spectral radiance, it may be close enough to suggest if a star is an astronomical green source using the above derivative.

From a Planckian spectrum peaked in the green radiation band the wavelength temperature pairs are approximately (495 nm, 6100 K) and (570 nm, 5260 K).

As the Sun has a surface temperature of 5777 K, it is a green radiation source as a Planckian radiator.

Hubble Space Telescope

File:HST-SM4.jpeg
The Hubble Space Telescope is seen from the departing Space Shuttle Atlantis, flying Servicing Mission 4 (STS-125), the fifth and final human spaceflight to visit the observatory. Credit: Ruffnax (Crew of STS-125).

The Hubble Space Telescope (HST) is an excellent example of a radiation astronomy satellite designed for more than one purpose: the various astronomies of optical astronomy.

The HST is an optical astronomy telescope that incorporated a set of 48 filters isolating spectral lines of particular astrophysical interest.

The Wide Field Planetary Camera (PC-1) was in use from about 1990 through 1993. It carried 48 filters on 12 filter wheels of four each. For the green band, these were the F492M, F502N, F517N, F547M, and the F555W. Those ending in 'N' are narrow-band filters.[75]

One of these filters is F492M which allows imaging with the [O III]λλ4959,5007 and its adjacent green continuum. The filter band pass is centered at 490.6 nm with a full-width at half maximum (FWHM) of 36.4 nm.[12] "The F492M filter also includes Hβ.[12]

The F502N is centered at 501.85 nm with a band pass of 2.97 nm. The F547M is centered at 546.1 nm with a band pass of 43.8 nm.

The Wide Field Planetary Camera (PC-2) replaced PC-1 and carried the following filters on the same filter wheels: F467M, F502N, F547M, F555W, and the F569W.[75] In December 1993 PC-1 was replaced with PC-2 and the HST was declared operational on January 13, 1994.

Onboard the HST is the Faint Object Camera (FOC) which carries filters for green astronomy: F470M, F480LP, F501N, F502N, and the F550M.[75]

The Wide Field Camera 3 (WFC3) is the Hubble Space Telescope's last and most technologically advanced instrument to take images in the visible spectrum. It was installed as a replacement for the Wide Field and Planetary Camera 2 during the first spacewalk of Space Shuttle mission STS-125 on May 14, 2009.

Filters

File:UHC filter spectrum.jpg
This image shows the three main spectral lines that Narrowband filters transmit. Credit: Vedran V.

Most astronomical filters work by blocking a specific part of the color spectrum above and below a bandpass, significantly increasing the signal to noise of the interesting wavelengths, and so making the object more visible, 'contrasty', or defined. ... The broadband and narrowband filters transmit the wavelengths that are emitted by the nebulae (by the Hydrogen and Oxygen atoms), and are frequently used for reducing light pollution.[76]

Color filters work by absorption/transmission, and can tell which part of the spectrum they are reflecting and transmitting. Filters can be used to increase contrast and enhance the details of the Moon and planets. All of the visible spectrum colors each have a filter, and every color filter is used to bring a certain lunar and planetary feature; for example, the #8 yellow filter is used to show Mars's maria and Jupiter's belts.[77] The Wratten system is the standard number system used to refer to the color filter types. Professional filters are also colored, but their bandpass centers are placed around other mid-points (such as in the UBVRI and Cousins systems).

The Dark green: Improves cloud patterns on Venus. Reduces sky brightness during daylight observation of Venus. Increases contrast of ice and polar caps on Mars. Improves visibility of the Great Red Spot on Jupiter and other features in Jupiter atmosphere. Enhances white clouds and polar regions on Saturn.

Narrowband filters are astronomical filters which transmit only a narrow band of spectral lines from the spectrum (usually 22 nm or less). It is mainly used for nebulae observation. Emission nebulae mainly radiate the doubly ionized oxygen in the visible spectrum, which emits near 500 nm wavelength. These nebulae also radiate weaker at 486 nm from the Hydrogen-beta atoms. There are three main types of Narrowband filters: Ultra-high contrast (UHC), Oxygen-III & Hydrogen-beta, and Hydrogen-alpha, the narrowest of the three filters with 8 nm range. The UHC filters range from 484 to 506 nm.[77] It transmits both the O-III and H-beta spectral lines, blocks a large fraction of light pollution, and brings the details of planetary nebulae and most of emission nebulae under a dark sky.[78]

The Broadband or light pollution reduction (LPR) filters are nebular filters that block the light pollution in the sky and transmit the H-alpha, H-beta, and O-III spectral lines, which makes observing nebulae from the city and light polluted skies possible.[76] These filters block the Sodium and Mercury vapor light, and also block the natural skyglow such as the auroral light.[79] The broadband filters differ from the narrowband with the range of wavelengths transmission. The broadband filters have a wider range because the narrower transmission range causes a fainter image of sky objects, and since the work of these filters is revealing the details of nebulae from light polluted skies, it has a wider transmission for more brightness.[77] These filters are particularly designed for nebulae observing, are not useful with other deep sky objects. However, it can improve the contrast between the DSOs and the background sky, which may clarify the image.

Spacecraft

File:Galileo Solid- State Imaging.jpg
This is an image of the Solid-State Imaging camera of the Galileo spacecraft. Credit: NASA/JPL.

"[T]he Solid State Imaging (SSI) camera ... [at right of the Galileo spacecraft] uses a high-resolution, 800 x 800 charge-coupled device (CCD) array with a field of view of 0.46 degrees. Multi-spectral coverage is provided by an eight-position filter wheel on the camera, consisting of three broad-band filters: violet (404 nm), green (559 nm), and red (671 nm); four near-infrared filters: 727 nm, 756 nm, 889 nm, and 986 nm; and one clear filter (611 nm) with a very broad (440 nm) passband."[80]

Hypotheses

  1. At least one image has been taken in the green of each major astronomical object or source in the Solar System.

Acknowledgements

The content on this page was first contributed by: Henry A. Hoff.

Initial content for this page in some instances came from Wikiversity.

See also

References

  1. Malachite. WebExhibits. 2001. Retrieved 2007-12-08.
  2. The sRGB values are taken by converting the NCS color 2060-G using the "NCS Navigator" tool at the NCS website.
  3. 3.0 3.1 (Oxford English Dictionary, 2nd Edition, Clarendon Press, Oxford, 1989.) See also first definition in Webster's New World Dictionary of the American Language, The World Publishing Company, New York, 1962.
  4. Eva Heller, Psychologie de la couleur- effets et symboliques. pg. 87-104
  5. Glossary Term: Color wheel. Sanford Corp. 2005. Retrieved 2007-11-22.
  6. Kenneth L. Kelly (1943). "Color Designations for Lights". Journal of the Optical Society of America 33(11). 627–632.
  7. Anne Varichon (2000), Couleurs- pigments et teintures dans les mains des peoples. Pg. 210-211.
  8. Philip B. Gove, ed. (1963). Webster's Seventh New Collegiate Dictionary. Springfield, Massachusetts: G. & C. Merriam Company. p. 1221. |access-date= requires |url= (help)
  9. Stanley 1999, pp. 302–303
  10. 10.0 10.1 10.2 Holli Riebeek (January 3, 2009). Sediment off the Yucatan Peninsula. NASA. Retrieved 2012-07-22.
  11. S.F. Sánchez, J. Aceituno, U. Thiele, D. Pérez-Ramírez, J. Alves (2007). "The Night-Sky at the Calar Alto Observatory" (PDF). Publications of the Astronomical Society of the Pacific. 119 (860): 1186–1200. doi:10.1086/522378. Retrieved 2013-01-24. Unknown parameter |month= ignored (help)
  12. 12.0 12.1 12.2 12.3 12.4 12.5 12.6 12.7 A. S. Wilson, J. A. Braatz, T. M. Heckman, J. H. Krolik, and G. K. Miley (1993). "The Ionization Cones in the Seyfert Galaxy NGC 5728". The Astrophysical Journal Letters. 419 (12): L61–4. Bibcode:1993ApJ...419L..61W. doi:10.1086/187137. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)
  13. E. C. C. Baly (1898). "Helium in the Atmosphere". Nature. 58 (1510): 545. Bibcode:1898Natur..58Q.545B. doi:10.1038/058545a0. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)
  14. Barbara Mattson (September 5, 2006). Solution for Graphing Spectra Student Worksheet, Part II. NASA GSFC, Greenbelt, Maryland, USA: NASA's Imagine the Universe. Retrieved 2013-01-23.
  15. 15.0 15.1 15.2 15.3 15.4 15.5 K. J. McCarthy, A. Baciero, B. Zurro, and TJ-II Team (June 12, 2000). Impurity Behaviour Studies in the TJ-II Stellarator, In: 27th EPS Conference on Contr. Fusion and Plasma Phys. (PDF). 24B. Budapest: ECA. pp. 1244–7. Retrieved 2013-01-20.
  16. Anita L. Cochran, W. W. Jackson (2006). "What is the Parent of Cometary O (1S)?". Bulletin of the American Astronomical Society. 38: 516. Bibcode:2006DPS....38.2005C. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)
  17. D. R. Bates (1978). "Forbidden oxygen and nitrogen lines in the nightglow". Planetary and Space Science. 26 (10): 897–912. Bibcode:1978P&SS...26..897B. doi:10.1016/0032-0633(78)90073-9. Retrieved 2013-01-16. Unknown parameter |month= ignored (help)
  18. R. J. Thomas and R. A. Young (1981). "Measurement of atomic oxygen and related airglows in the lower thermosphere". Journal of Geophysical Research: Oceans. 86 (08): 7389–93. Bibcode:1981JGR....86.7389T. doi:10.1029/JC086iC08p07389. Retrieved 2013-01-16. Unknown parameter |month= ignored (help)
  19. 19.0 19.1 J. C. McLennan, G. M. Shrum (1925). "On the Origin of the Auroral Green Line 5577 Å, and other Spectra Associated with the Aurora Borealis" (PDF). Proceedings of the Royal Society A Mathematical, Physical & Engineering Sciences. 108 (747): 501–12. doi:10.1098/rspa.1925.0088. Retrieved 2013-01-24.
  20. Ds. Lauretta, L.P. Keller, S. Messenger (2005). "Supernova olivine from cometary dust". Science. 309 (5735): 737–741. Bibcode:2005Sci...309..737M. doi:10.1126/science.1109602. PMID 15994379.
  21. Spitzer sees crystal 'rain' in outer clouds of infant star, Whitney Clavin and Trent Perrotto, Physorg.com, May 27, 2011 . Accessed May 2011
  22. 22.0 22.1 P. Swings (1943). "Edlén's Identification of the Coronal Lines with Forbidden Lines of Fe X, XI, XIII, XIV, XV; Ni XII, XIII, XV, XVI; Ca XII, XIII, XV; a X, XIV". The Astrophysical Journal. 98 (07): 116–28. Bibcode:1943ApJ....98..116S. doi:10.1086/144550. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)
  23. 23.0 23.1 23.2 P. G. Nutting (1906). "Line Structure. I.". The Astrophysical Journal. 23 (1): 64–78. Bibcode:1906ApJ....23...64N. doi:10.1086/141302. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)
  24. 24.0 24.1 24.2 24.3 24.4 24.5 24.6 G. Stellmacher and E. Wiehr (1991). "Geometric line elevation in solar limb faculae". Astronomy and Astrophysics. 248 (1): 227–31. Bibcode:1991A&A...248..227S. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)
  25. M. Rybansky, V. Rusin, M. Minarovjech and P. Gaspar (1994). "Coronal index of solar activity: Years 1939-1963". Solar Physics. 152 (1): 153–9. Bibcode:1994SoPh..152..153R. doi:10.1007/BF01473198. |access-date= requires |url= (help)
  26. John G. Phillips, Sumner P. Davis, Bo Lindgren, and Walter J. Balfour (1987). "The near-infrared spectrum of the FeH molecule". The Astrophysical Journal Supplement Series. 65 (12): 721–78. Bibcode:1987ApJS...65..721P. doi:10.1086/191241. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)
  27. DE Fawzy, NH Youssef, O. Engvold (1998). "Identification of FeH molecular lines in the spectrum of a sunspot umbra". Astronomy and Astrophysics Supplement. 129 (5): 435–43. Bibcode:1998A&AS..129..435F. doi:10.1051/aas:1998196. Retrieved 2012-02-18. Unknown parameter |month= ignored (help)
  28. 28.0 28.1 28.2 28.3 28.4 M. C. Festou and P. D. Feldman (1981). "The Forbidden Oxygen Lines in Comets". Astronomy & Astrophysics. 103 (1): 154–9. Bibcode:1981A&A...103..154F. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)
  29. Jean E. Sansonetti and W. C. Martin, and S. L. Young (December 9, 2011). Handbook of Basic Atomic Spectroscopic Data. 100 Bureau Drive, Gaithersburg, Maryland USA: Physical Measurement Laboratory, NIST. Retrieved 2013-01-24.
  30. Dispersive refraction by webexhibits.org.
  31. Green and red rims by Andy Young.
  32. 32.0 32.1 32.2 G. A. Chapman & T. E. McGuire (1977). "The wavelength dependence of the facular excess brightness". The Astrophysical Journal. 217 (10): 657–60. Bibcode:1977ApJ...217..657C. doi:10.1086/155611. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)
  33. 33.0 33.1 33.2 H. A. Hill & R. T. Stebbins (1975). "The intrinsic visual oblateness of the sun". The Astrophysical Journal. 200 (9): 471–7, 477–83. Bibcode:1975ApJ...200..471H. doi:10.1086/153813. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)
  34. 34.0 34.1 R. Schwenn, B. Inhester, S. P. Plunkett, A. Epple, B. Podlipnik, D. K. Bedford, C. J. Eyles, G. M. Simnett, S. J. Tappin and M. V. Bout; et al. (1997). "First View of the Extended Green-Line Emission Corona At Solar Activity Minimum Using the Lasco-C1 Coronagraph on Soho" (PDF). Solar Physics. 175 (2): 667–84. doi:10.1023/A:1004948913883. Retrieved 2012-02-17. Unknown parameter |month= ignored (help)
  35. Y. C. Unruh, S. K. Solanki, & M. Fligge (1999). "The spectral dependence of facular contrast and solar irradiance variations". Astronomy and Astrophysics. 345 (5): 635–42. Bibcode:1999A&A...345..635U. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)
  36. T. G. Slanger (2005). "Practicality of Using Oxygen Atom Emissions to Evaluate the Habitability of Extra-Solar Planets". American Geophysical Union, Fall Meeting 2005. Bibcode:2005AGUFMSA53B1174S. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)
  37. T. G. Slanger, D. L. Huestis, P. C. Cosby, N. J. Chanover, T. A. Bida, (2006). "The Venus nightglow: Ground-based observations and chemical mechanisms". Icarus. 182 (1): 1–9. Bibcode:2006Icar..182....1S. doi:10.1016/j.icarus.2005.12.007. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)
  38. Tom G. Slanger, Nancy J. Chanover, Brian D. Sharpee, Thomas A. Bida (2012). "O/O2 emissions in the Venus nightglow". Icarus. 217 (2): 845–8. doi:10.1016/j.icarus.2011.03.031. Retrieved 2013-01-20. Unknown parameter |month= ignored (help)
  39. Dan Burbank (December 21, 2011). Love and Joy for the New Year. International Space Station: NASA's Earth Observatory. Retrieved 2012-07-22.
  40. Ultraviolet Waves.
  41. Earth Sciences and Image Analysis Laboratory, Johnson Space Center (November 11, 2001). Green Aurora Seen from the Space Station. International Space Station: NASA. Retrieved 2012-07-22.
  42. 42.0 42.1 Holli Riebeek (May 5, 2009). Coral Reefs in the Persian Gulf. NASA Earth Observatory. Retrieved 2012-07-22.
  43. 43.0 43.1 Adam Voiland (April 19, 2012). Spring Paints the Piedmont Green. NASA's Aqua satellite: NASA's Earth Observatory. Retrieved 2012-07-22.
  44. 44.0 44.1 44.2 44.3 Michon Scott (February 28, 2002). Hydrogen Sulfide Emissions off of Africa. NASA Earth Observatory. Retrieved 2012-07-22.
  45. 45.0 45.1 Rebecca Lindsey and Norman Kuring (March 8, 2010). Phytoplankton Bloom in the Arabian Sea. NASA’s Aqua satellite: NASA Earth Observatory. Retrieved 2012-07-22.
  46. 46.0 46.1 46.2 46.3 Michael Carlowicz (December 16, 2011). Tremors, Eruption at El Hierro Subsiding?. NASA Earth Observatory. Retrieved 2012-07-22.
  47. W. J. Baggaley (1976). "The Excitation of the Oxygen Metastable OI(1S) State in Meteors". Bulletin of the Astronomical Institutes of Czechoslovakia. 27 (3): 173–81. Bibcode:1976BAICz..27..173B. |access-date= requires |url= (help)
  48. Trnka, M.; Houzar, S. (2002). "Moldavites: a review PDF" (PDF). Bulletin of the Czech Geological Survey. 77 (4): 283–302.
  49. Serefiddin, F.; Herzog, G. F.; Koeberl, C. (2007). "Beryllium-10 concentrations of tektites from the Ivory Coast and from Central Europe: Evidence for near-surface residence of precursor materials" (PDF). Geochimica et Cosmochimica Acta. 71 (6): 1574–82.
  50. Internet Encyclopedia of Science Accessed April 2010
  51. Reach, W. T. (1997). "The structured zodiacal light: IRAS, COBE, and ISO observations", page 1 (in Introduction)
  52. John T. Jefferies (1975). "University of Hawaii, Institute for Astronomy, Honolulu, Hawaii 96822". Bulletin of the Astronomical Society. 7 (01): 62–9. Bibcode:1975BAAS....7...62J. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)
  53. 53.0 53.1 53.2 Danielle Briot, Jean Schneider, and Luc arnold (October 2003). M. Fridlund, T. Henning, ed. The terrestrial vegetation observed in the Earthshine spectrum: a test for the detectability of vegetation on extrasolar planets, In: Proceedings of the Conference on Towards Other Earths: DARWIN/TPF and the Search for Extrasolar Terrestrial Planets. Noordwijk, Netherlands: European Space agency. pp. 375–8. Bibcode:2003ESASP.539..375B. ISBN 92-9092-849-2. |access-date= requires |url= (help)
  54. 54.0 54.1 Yvette Smith (December 8, 2009). Earth's Moon. NASA. Retrieved 2012-07-22.
  55. Guillaume Gronoff, Cyril Simon Wedlund, Christopher J. Mertens, Mathieu Barthélemy, Robert J. Lillis, Olivier Witasse (2012). "Computing uncertainties in ionosphere-airglow models: II. The Martian airglow". Journal of Geophysical Research. 117 (A5). Bibcode:2012JGRA..11705309G. doi:10.1029/2011JA017308. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)
  56. Tom Slanger, Brian Sharpee, Dusan Pejakovic, Richard Gattinger, Edward J. Llewellyn, Ian McDade, David Siskind, Kenneth Minschwaner (2010). "The 557.7 and 297.2 nm lines of O(1S) in the atmospheres of the terrestrial planets". 38th COSPAR Scientific Assembly: 9. Bibcode:2010cosp...38.1393S. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)
  57. Sue Lavoie (February 1, 1996). PIA00136: Asteroid Ida and Its Moon. Pasadena, California USA: NASA/JPL. Retrieved 2013-01-25.
  58. 58.0 58.1 Sue Lavoie (October 13, 1998). PIA01637: Io's Aurorae. Pasadena, California: NASA and the Jet Propulsion Laboratory, California Institute of Technology. Retrieved 2012-07-22.
  59. 59.0 59.1 59.2 59.3 Sue Lavoie (November 28, 2005). PIA07760: Spray Above Enceladus. Pasadena, California: NASA and the Jet Propulsion Laboratory, California Institute of Technology. Retrieved 2012-07-22.
  60. 60.0 60.1 Sue Lavoie (December 26, 2005). PIA07660: A Moon with Two Dark Sides. NASA and the Jet Propulsion Laboratory, California Institute of Technology. Retrieved 2012-07-22.
  61. 61.0 61.1 61.2 Tilmann Denk, Carolyn Porco, Joe Mason (December 10, 2009). Color Dichotomy on Iapetus. NASA/JPL/Space Science Institute. Retrieved 2012-07-22.
  62. 62.0 62.1 62.2 Tony Greicius (February 3, 2012). Cassini Unlocking Saturn's Secrets. NASA. Retrieved 2012-08-09.
  63. 63.0 63.1 Heidi Hammel (July 31, 1997). Clouds on Uranus. Boston, Massachusetts: NASA. Retrieved 2012-07-21.
  64. 64.0 64.1 64.2 64.3 I. Bertini, N. Thomas, and C. Barbieri (2007). "Modeling of the light scattering properties of cometary dust using fractal aggregates". Astronomy & Astrophysics. 461 (1): 351–64. Bibcode:2007A&A...461..351B. doi:10.1051/0004-6361:20065461. Retrieved 2011-12-08. Unknown parameter |month= ignored (help); Unknown parameter |pdf= ignored (help)
  65. Anita L. Cochran, William D. Cochran (2001). "Observations of O (1S) and O (1D) in Spectra of C/1999 S4 (LINEAR)" (PDF). Icarus. 154 (2): 381–90. arXiv:astro-ph/0108065. Bibcode:2001Icar..154..381C. doi:10.1006/icar.2001.6718. Retrieved 2013-01-16. Unknown parameter |month= ignored (help)
  66. Rnt20 (8 January 2007). "Stardust (spacecraft)". San Francisco, California: Wikimedia Foundation, Inc. Retrieved 5 July 2019.
  67. News Release Number: STScI-2001-34 (December 19, 2001). Wallpaper: The Ghost-Head Nebula (NGC 2080). NASA and the Hubble Space Telescope. Retrieved 2012-07-21.
  68. The Colour of Stars, In: Australia Telescope, Outreach and Education. Commonwealth Scientific and Industrial Research Organisation. December 21, 2004. Retrieved 2012-01-16.
  69. 69.0 69.1 Pierre Kervella and Frederic Thevenin (March 15, 2003). A Family Portrait of the Alpha Centauri System. ESO. Retrieved 2008-06-06.
  70. Research Consortium on Nearby Stars (17 September 2007). The One Hundred Nearest Star Systems. Georgia State University. Retrieved 2007-11-06.
  71. M. Bazot; et al. (2007). "Asteroseismology of α Centauri A. Evidence of rotational splitting". Astronomy and Astrophysics. 470 (1): 295–302. arXiv:0706.1682. Bibcode:2007A&A...470..295B. doi:10.1051/0004-6361:20065694.
  72. C. A. Hummel; et al. (1994). "Very high precision orbit of Capella by long baseline interferometry". The Astronomical Journal. 107 (5): 1859–67. Bibcode:1994AJ....107.1859H. doi:10.1086/116995. Unknown parameter |month= ignored (help)
  73. Berdyugina 2.4 RS CVn stars
  74. 74.0 74.1 74.2 David N. Talbott (1980). The Saturn Myth. Garden City, New York, USA: Knopf Doubleday & Company, Inc. p. 419. ISBN 0-385-11376-5. Retrieved 2013-01-03.
  75. 75.0 75.1 75.2 John Krist and Richard Hook (June 2004). The Tiny Tim User’s Guide, Version 6.3. http://tinytim.stsci.edu/static/tinytim.pdf: Space Telescope Science Institute. Retrieved 2013-01-22.
  76. 76.0 76.1 astronomy for everyone: the use of filters. Retrieved 2010-11-22.
  77. 77.0 77.1 77.2 filters - popular and hot telescope filters. Lumicon international. Retrieved 2010-11-22.
  78. UHC filters. Retrieved 2010-11-22.
  79. Meade series 4000 Broadband Nebular filters. Meade Instruments. Retrieved 2010-11-23.
  80. David R. Williams (June 13, 2008). Galileo (1990, 1992). Greenbelt, MD 20771 USA: NASA Goddard Space Flight Center. Retrieved 2013-01-25.

Further reading

  • Alexander Kramida, Yuri Ralchenko, and Joseph Reader (September 24, 2012). NIST Atomic Spectra Database, Version 5. 100 Bureau Drive, Gaithersburg, Maryland USA: National Institutes of Standards and Technology (NIST). Retrieved 2013-01-20.
  • Jean E. Sansonetti and W. C. Martin, and S. L. Young (December 9, 2011). Handbook of Basic Atomic Spectroscopic Data. 100 Bureau Drive, Gaithersburg, Maryland USA: Physical Measurement Laboratory, NIST. Retrieved 2013-01-24.
  • G. Stellmacher and E. Wiehr (1991). "Geometric line elevation in solar limb faculae". Astronomy and Astrophysics. 248 (1): 227–31. Bibcode:1991A&A...248..227S. Unknown parameter |month= ignored (help); |access-date= requires |url= (help)

External links

{{Radiation astronomy resources}}{{Geology resources}}Template:Principles of radiation astronomyTemplate:Sisterlinks