Collision-induced absorption and emission

Last updated

In spectroscopy, collision-induced absorption and emission refers to spectral features generated by inelastic collisions of molecules in a gas. Such inelastic collisions (along with the absorption or emission of photons) may induce quantum transitions in the molecules, or the molecules may form transient supramolecular complexes with spectral features different from the underlying molecules. Collision-induced absorption and emission is particularly important in dense gases, such as hydrogen and helium clouds found in astronomical systems.

Contents

Collision-induced absorption and emission is distinguished from collisional broadening in spectroscopy in that collisional broadening comes from elastic collisions of molecules, whereas collision-induced absorption and emission is an inherently inelastic process.

Collision-induced spectra of gases

Ordinary spectroscopy is concerned with the spectra of single atoms or molecules. Here we outline the very different spectra of complexes consisting of two or more interacting atoms or molecules: the "interaction-induced" or "collision-induced" spectroscopy. [1] Both ordinary and collision-induced spectra may be observed in emission and absorption and require an electric or magnetic multipole moment - in most cases an electric dipole moment - to exist for an optical transition to take place from an initial to a final quantum state of a molecule or a molecular complex. (For brevity of expression we will use here the term "molecule" interchangeably for atoms as well as molecules). A complex of interacting molecules may consist of two or more molecules in a collisional encounter, or else of a weakly bound van der Waals molecule. On first sight, it may seem strange to treat optical transitions of a collisional complex, which may exist just momentarily, for the duration of a fly-by encounter (roughly 10−13 seconds), in much the same way as this was long done for molecules in ordinary spectroscopy. But even transient complexes of molecules may be viewed as a new, "supermolecular" system which is subject to the same spectroscopic rules as ordinary molecules. Ordinary molecules may be viewed as complexes of atoms that have new and possibly quite different spectroscopic properties than the individual atoms the molecule consists of, when the atoms are not bound together as a molecule (or are not "interacting"). Similarly, complexes of interacting molecules may (and usually do) acquire new optical properties, which often are absent in the non-interacting, well separated individual molecules.

Collision-induced absorption (CIA) and emission (CIE) spectra are well known in the microwave and infrared regions of the electromagnetic spectrum, but they occur in special cases also in the visible and near ultraviolet regions. [1] [2] Collision-induced spectra have been observed in nearly all dense gases, and also in many liquids and solids. [3] [4] CIA and CIE are due to the intermolecular interactions, which generate electric dipole moments. We note that an analogous collision-induced light scattering (CILS) or Raman process also exists, which is well studied and is in many ways completely analogous to CIA and CIE. CILS arises from interaction-induced polarizability increments of molecular complexes; the excess polarizability of a complex, relative the sum of polarizabilities of the noninteracting molecules. [5]

Interaction-induced dipoles

Molecules interact at close range through intermolecular forces (the "van der Waals forces"), which cause minute shifts of the electron density distributions (relative the distributions of electrons when the molecules are not interacting). Intermolecular forces are repulsive at near range, where electron exchange forces dominate the interaction, and attractive at somewhat greater separations, where the dispersion forces are active. (If separations are further increased, all intermolecular forces fall off rapidly and may be totally neglected.) Repulsion and attraction are due, respectively, to the small defects or excesses of electron densities of molecular complexes in the space between the interacting molecules, which often result in interaction-induced electric dipole moments that contribute some to interaction-induced emission and absorption intensities. The resulting dipoles are referred to as exchange force-induced dipole and dispersion force-induced dipoles, respectively.

Other dipole induction mechanisms also exist in molecular (as opposed to monatomic) gases and in mixtures of gases, when molecular gases are present. Molecules have centers of positive charge (the nuclei), which are surrounded by a cloud of electrons. Molecules thus may be thought of being surrounded by various electric multipolar fields which will polarize any collisional partner momentarily in a fly-by encounter, generating the so-called multipole-induced dipoles. In diatomic molecules such as H2 and N2, the lowest-order multipole moment is the quadrupole, followed by a hexadecapole, etc., hence the quadrupole-induced, hexadecapole-induced,... dipoles. Especially the former is often the strongest, most significant of the induced dipoles contributing to CIA and CIE. Other induced dipole mechanisms exist. In collisional systems involving molecules of three or more atoms (CO2, CH4...), collisional frame distortion may be an important induction mechanism. [2] Collision-induced emission and absorption by simultaneous collisions of three or more particles generally do involve pairwise-additive dipole components, as well as important irreducible dipole contributions and their spectra. [6]

Historical sketch

Collision-induced absorption was first reported in compressed oxygen gas in 1949 by Harry Welsch and associates at frequencies of the fundamental band of the O2 molecule. [7] (Note that an unperturbed O2 molecule, like all other diatomic homonuclear molecules, is infrared inactive on account of the inversion symmetry and does thus not possess a "dipole allowed" rotovibrational spectrum at any frequency).

Collision-induced spectra

Molecular fly-by collisions take little time, something like 10−13 s. Optical transition of collisional complexes of molecules generate spectral "lines" that are very broad - roughly five orders of magnitude broader than the most familiar "ordinary" spectral lines (Heisenberg's uncertainty relation). [1] [2] The resulting spectral "lines" usually strongly overlap so that collision-induced spectral bands typically appear as continua (as opposed to the bands of often discernible lines of ordinary molecules).

Collision-induced spectra appear at the frequencies of the rotovibrational and electronic transition bands of the unperturbed molecules, and also at sums and differences of such transition frequencies: simultaneous transitions in two (or more) interacting molecules are well known to generate optical transitions of molecular complexes. [1]

Virial expansions of spectral intensities

Intensities of spectra of individual atoms or molecules typically vary linearly with the numerical gas density. However, if gas densities are sufficiently increased, quite generally contributions may also be observed that vary as density squared, cubed... These are the collision-induced spectra of two-body (and quite possibly three-body,...) collisional complexes. The collision-induced spectra have sometimes been separated from the continua of individual atoms and molecules, based on the characteristic density dependences. In other words, a virial expansion in terms of powers of the numerical gas density is often observable, just as this is widely known for the virial expansion of the equation of state of compressed gases. The first term of the expansion, which is linear in density, represents the ideal gas (or "ordinary) spectra where these exist. (This first term vanishes for the infrared inactive gases,) And the quadratic, cubic,... terms of the virial expansions arise from optical transitions of binary, ternary,... intermolecular complexes, which are (often unjustifyably) neglected in the ideal gas approximation of spectroscopy.

Spectra of van der Waals molecules

Two kinds of complexes of molecules exist: the collisional complexes discussed above, which are short lived. Besides, bound (i.e. relatively stable) complexes of two or more molecules exist, the so-called van der Waals molecules. These exist usually for much longer times than the collisional complexes and, under carefully chosen experimental conditions (low temperature, moderate gas density), their rotovibrational band spectra show "sharp" (or resolvable) lines (Heisenberg uncertainty principle), much like ordinary molecules. If the parent molecules are nonpolar, the same induced dipole mechanisms, which are discussed above, are responsible for the observable spectra of van der Waals molecules.

Figure 1 (to be included)

An example of CIA spectra

Figure 1 shows an example of a collision-induced absorption spectra of H2-He complexes at a variety of temperatures. The spectra were computed from the fundamental theory, using quantum chemical methods, and were shown to be in close agreement with laboratory measurements at temperatures, where such measurements exist (for temperatures around 300 K and lower). [8] The intensity scale of the figure is highly compressed. At the lowest temperature (300 K), a series of six striking maxima is seen, with deep minima between them. The broad maxima roughly coincide with the H2 vibrational bands. With increasing temperature, the minima become less striking and disappear at the highest temperature (curve at the top, for the temperature of 9000 K).

A similar picture is to be expected for the CIA spectra of pure hydrogen gas (i.e. without admixed gases) and, in fact for the CIA spectra of many other gases. The main difference, say if nitrogen CIA spectra are considered instead of those of hydrogen gas, would be a much closer spacing, if not a total overlapping, of the diverse CIA bands which appear roughly at the frequencies of the vibrational bands of the N2 molecule.

Significance

The significance of CIA for astrophysics was recognized early-on, especially where dense atmospheres of mixtures of molecular hydrogen and helium gas exist. [9]

Planets

Herzberg pointed out direct evidence of H2 molecules in the atmospheres of the outer planets. [10] [11] The atmospheres of the inner planets and of Saturn's big moon Titan also show significant CIA in the infrared due to concentrations of nitrogen, oxygen, carbon dioxide and other molecular gases. [12] [13] [14] However, the total CIA contribution of Earth's major gases, N2 and O2, to the atmosphere's natural greenhouse effect is relatively minor except near the poles. [15] Extrasolar planets have been discovered with hot atmospheres (a thousand kelvin or more) which otherwise resemble Jupiter's atmosphere (mixtures of mostly H2 and He) where relatively strong CIA exists. [16]

Cool white dwarf stars

Stars that burn hydrogen are called main sequence (MS) stars - these are by far the most common objects in the night sky. When the hydrogen fuel is exhausted and temperatures begin to fall, the object undergoes various transformations and a white dwarf star is eventually born, the ember of the expired MS star. Temperatures of a new-born white dwarf may be in the hundreds of thousand kelvin, but if the mass of the white dwarf is less than just a few solar masses, burning of 4He to 12C and 16O is not possible and the star will slowly cool down forever. The coolest white dwarfs observed have temperatures of roughly 4000 K, which must mean that the universe is not old enough so that lower temperature stars cannot be found. The emission spectra of "cool" white dwarfs does not at all look like a Planck blackbody spectrum. [17] Instead, nearly the whole infrared is attenuated or missing altogether from the star's emission, owing to CIA in the hydrogen-helium atmospheres surrounding their cores. [18] [19] The impact of CIA on the observed spectral energy distribution is well understood and accurately modeled for most cool white dwarfs. [20] For white dwarfs with a mix H/He atmosphere, the intensity of the H2-He CIA can be used to infer the hydrogen abundance at the white dwarf photosphere. [21] However, predicting CIA in the atmospheres of the coolest white dwarfs is more challenging, [22] in part because of the formation of many-body collisional complexes. [23]

Other cool stars

The atmospheres of low metallicity cool stars are composed primarily of hydrogen and helium. Collision-induced absorption by H2-H2 and H2-He transient complexes will be a more or less important opacity source of their atmospheres. For example, CIA in the H2 fundamental band, which falls on top of an opacity window between H2O/CH4 or H2O/CO (depending on the temperature), plays an important role in shaping brown dwarf spectra. [24] [25] [26] Higher gravity brown dwarf stars often show even stronger CIA, owing to the density squared dependence of CIA intensities, when other "ordinary" opacity sources are linearly dependent on density. CIA is also important in low-metallicity brown dwarfs, since "low metallicity" means reduced CNO (and other) elemental abundances compared to H2 and He, and thus stronger CIA compared to H2O, CO, and CH4 absorption. CIA absorption of H2-X collisional complexes is thus an important diagnostic of high-gravity and low-metallicity brown dwarfs. [27] [28] All of this is also true of the M dwarfs, but to a lesser extent. M dwarf atmospheres are hotter so that some increased portion of the H2 molecules is in the dissociated state, which weakens CIA by H2--X complexes. The significance of CIA for cool astronomical objects was long suspected or known to some degree. [29] [30]

First stars

Attempts to model the formation of the "first" star from the pure hydrogen and helium gas clouds below about 10,000 K show that the heat generated in the gravitational contraction phase must be somehow radiatively released for further cooling to be possible. This is no problem as long as temperatures are still high enough so that free electrons exist: electrons are efficient emitters when interacting with neutrals (bremsstrahlung). However, at the lower temperatures in neutral gases, the recombination of hydrogen atoms to H2 molecules is a process that generates enormous amounts of heat that must somehow be radiated away in CIE processes; if CIE were non-existing, molecule formation could not take place and temperatures could not fall further. Only CIE processes permit further cooling, so that molecular hydrogen will accumulate. A dense, cool environment will thus develop so that a gravitational collapse and star formation can actually proceed. [31] [32]

Database

Because of the great importance of many types of CIA spectra in planetary and astrophysical research, a well known spectroscopy database (HITRAN) has been expanded to include a number of CIA spectra in various frequency bands and for a variety of temperatures. [33]

Related Research Articles

An intermolecular force (IMF) is the force that mediates interaction between molecules, including the electromagnetic forces of attraction or repulsion which act between atoms and other types of neighbouring particles, e.g. atoms or ions. Intermolecular forces are weak relative to intramolecular forces – the forces which hold a molecule together. For example, the covalent bond, involving sharing electron pairs between atoms, is much stronger than the forces present between neighboring molecules. Both sets of forces are essential parts of force fields frequently used in molecular mechanics.

<span class="mw-page-title-main">Spectroscopy</span> Study involving matter and electromagnetic radiation

Spectroscopy is the field of study that measures and interprets electromagnetic spectra. In narrower contexts, spectroscopy is the precise study of color as generalized from visible light to all bands of the electromagnetic spectrum.

<span class="mw-page-title-main">Interstellar medium</span> Matter and radiation in the space between the star systems in a galaxy

In astronomy, the interstellar medium (ISM) is the matter and radiation that exists in the space between the star systems in a galaxy. This matter includes gas in ionic, atomic, and molecular form, as well as dust and cosmic rays. It fills interstellar space and blends smoothly into the surrounding intergalactic space. The energy that occupies the same volume, in the form of electromagnetic radiation, is the interstellar radiation field. Although the density of atoms in the ISM is usually far below that in the best laboratory vacuums, the mean free path between collisions is short compared to typical interstellar lengths, so on these scales the ISM behaves as a gas (more precisely, as a plasma: it is everywhere at least slightly ionized), responding to pressure forces, and not as a collection of non-interacting particles.

<span class="mw-page-title-main">Astrochemistry</span> Study of molecules in the Universe and their reactions

Astrochemistry is the study of the abundance and reactions of molecules in the universe, and their interaction with radiation. The discipline is an overlap of astronomy and chemistry. The word "astrochemistry" may be applied to both the Solar System and the interstellar medium. The study of the abundance of elements and isotope ratios in Solar System objects, such as meteorites, is also called cosmochemistry, while the study of interstellar atoms and molecules and their interaction with radiation is sometimes called molecular astrophysics. The formation, atomic and chemical composition, evolution and fate of molecular gas clouds is of special interest, because it is from these clouds that solar systems form.

<span class="mw-page-title-main">Absorption spectroscopy</span> Spectroscopic techniques that measure the absorption of radiation

Absorption spectroscopy is spectroscopy that involves techniques that measure the absorption of electromagnetic radiation, as a function of frequency or wavelength, due to its interaction with a sample. The sample absorbs energy, i.e., photons, from the radiating field. The intensity of the absorption varies as a function of frequency, and this variation is the absorption spectrum. Absorption spectroscopy is performed across the electromagnetic spectrum.

<span class="mw-page-title-main">Rotational spectroscopy</span> Spectroscopy of quantized rotational states of gases

Rotational spectroscopy is concerned with the measurement of the energies of transitions between quantized rotational states of molecules in the gas phase. The rotational spectrum of polar molecules can be measured in absorption or emission by microwave spectroscopy or by far infrared spectroscopy. The rotational spectra of non-polar molecules cannot be observed by those methods, but can be observed and measured by Raman spectroscopy. Rotational spectroscopy is sometimes referred to as pure rotational spectroscopy to distinguish it from rotational-vibrational spectroscopy where changes in rotational energy occur together with changes in vibrational energy, and also from ro-vibronic spectroscopy where rotational, vibrational and electronic energy changes occur simultaneously.

<span class="mw-page-title-main">Hydroxyl radical</span> Neutral form of the hydroxide ion (OH−)

The hydroxyl radical, HO, is the neutral form of the hydroxide ion (HO). Hydroxyl radicals are highly reactive and consequently short-lived; however, they form an important part of radical chemistry. Most notably hydroxyl radicals are produced from the decomposition of hydroperoxides (ROOH) or, in atmospheric chemistry, by the reaction of excited atomic oxygen with water. It is also an important radical formed in radiation chemistry, since it leads to the formation of hydrogen peroxide and oxygen, which can enhance corrosion and SCC in coolant systems subjected to radioactive environments. Hydroxyl radicals are also produced during UV-light dissociation of H2O2 (suggested in 1879) and likely in Fenton chemistry, where trace amounts of reduced transition metals catalyze peroxide-mediated oxidations of organic compounds.

<span class="mw-page-title-main">Einstein coefficients</span> Quantities describing probability of absorption or emission of light

In atomic, molecular, and optical physics, the Einstein coefficients are quantities describing the probability of absorption or emission of a photon by an atom or molecule. The Einstein A coefficients are related to the rate of spontaneous emission of light, and the Einstein B coefficients are related to the absorption and stimulated emission of light. Throughout this article, "light" refers to any electromagnetic radiation, not necessarily in the visible spectrum.

<span class="mw-page-title-main">Trihydrogen cation</span> Polyatomic ion (H₃, charge +1)

The trihydrogen cation or protonated molecular hydrogen is a cation with formula H+
3
, consisting of three hydrogen nuclei (protons) sharing two electrons.

<span class="mw-page-title-main">HITRAN</span>

HITRAN molecular spectroscopic database is a compilation of spectroscopic parameters used to simulate and analyze the transmission and emission of light in gaseous media, with an emphasis on planetary atmospheres. The knowledge of spectroscopic parameters for transitions between energy levels in molecules is essential for interpreting and modeling the interaction of radiation (light) within different media.

<span class="mw-page-title-main">Helium hydride ion</span> Chemical compound

The helium hydride ion, hydridohelium(1+) ion, or helonium is a cation (positively charged ion) with chemical formula HeH+. It consists of a helium atom bonded to a hydrogen atom, with one electron removed. It can also be viewed as protonated helium. It is the lightest heteronuclear ion, and is believed to be the first compound formed in the Universe after the Big Bang.

<span class="mw-page-title-main">Van der Waals molecule</span> Weakly bound complex of atoms or molecules

A Van der Waals molecule is a weakly bound complex of atoms or molecules held together by intermolecular attractions such as Van der Waals forces or by hydrogen bonds. The name originated in the beginning of the 1970s when stable molecular clusters were regularly observed in molecular beam microwave spectroscopy.

<span class="mw-page-title-main">Imidogen</span> Inorganic radical with the chemical formula NH

Imidogen is an inorganic compound with the chemical formula NH. Like other simple radicals, it is highly reactive and consequently short-lived except as a dilute gas. Its behavior depends on its spin multiplicity.

<span class="mw-page-title-main">Chromium(I) hydride</span> Chemical compound

Chromium(I) hydride, systematically named chromium hydride, is an inorganic compound with the chemical formula (CrH)
n
. It occurs naturally in some kinds of stars where it has been detected by its spectrum. However, molecular chromium(I) hydride with the formula CrH has been isolated in solid gas matrices. The molecular hydride is very reactive. As such the compound is not well characterised, although many of its properties have been calculated via computational chemistry.

<span class="mw-page-title-main">Iron(I) hydride</span> Chemical compound

Iron(I) hydride, systematically named iron hydride and poly(hydridoiron) is a solid inorganic compound with the chemical formula (FeH)
n
(also written ([FeH])
n
or FeH). It is both thermodynamically and kinetically unstable toward decomposition at ambient temperature, and as such, little is known about its bulk properties.

Stellar molecules are molecules that exist or form in stars. Such formations can take place when the temperature is low enough for molecules to form – typically around 6,000 K or cooler. Otherwise the stellar matter is restricted to atoms and ions in the forms of gas or – at very high temperatures – plasma.

<span class="mw-page-title-main">Calcium monohydride</span> Chemical compound

Calcium monohydride is a molecule composed of calcium and hydrogen with formula CaH. It can be found in stars as a gas formed when calcium atoms are present with hydrogen atoms.

<span class="mw-page-title-main">Magnesium monohydride</span> Chemical compound

Magnesium monohydride is a molecular gas with formula MgH that exists at high temperatures, such as the atmospheres of the Sun and stars. It was originally known as magnesium hydride, although that name is now more commonly used when referring to the similar chemical magnesium dihydride.

<span class="mw-page-title-main">Argonium</span> Chemical compound

Argonium (also called the argon hydride cation, the hydridoargon(1+) ion, or protonated argon; chemical formula ArH+) is a cation combining a proton and an argon atom. It can be made in an electric discharge, and was the first noble gas molecular ion to be found in interstellar space.

<span class="mw-page-title-main">Diargon</span> Chemical compound

Diargon or the argon dimer is a molecule containing two argon atoms. Normally, this is only very weakly bound together by van der Waals forces. However, in an excited state, or ionised state, the two atoms can be more tightly bound together, with significant spectral features. At cryogenic temperatures, argon gas can have a few percent of diargon molecules.

References

  1. 1 2 3 4 Frommhold, Lothar (2006) [1993]. Collision-induced Absorption in Gases. Cambridge (NY): Cambridge University Press.
  2. 1 2 3 Abel, Martin; Frommhold, Lothar (2013) [1991]. "Collision-induced spectra and current astronomical research". Canadian Journal of Physics. 91 (11): 857–869. Bibcode:2013CaJPh..91..857A. doi:10.1139/cjp-2012-0532.
  3. Hunt, J. L.; Poll, J. D. (1986). A second bibliography on collision induced absorption. Vol. 59. Department of Physics, University of Guelph. pp. 163–164, Publication 1/86.{{cite book}}: |work= ignored (help)CS1 maint: location missing publisher (link)
  4. G. Birnbaum, ed. (1985). Phenomena Induced by Intermolecular Interactions. New York: Plenum Press.
  5. Borysow, Aleksandra; Frommhold, Lothar (1989). "Collision-induced light scattering: A bibliography". Advances in Chemical Physics. Vol. 75. pp. 439–505. doi:10.1002/9780470141243.ch7. ISBN   9780470141243.
  6. Moraldi, Massimo; Frommhold, Lothar (1996). "Dipole moments induced in three interacting molecules". Journal of Molecular Liquids. 70 (2–3): 143–158. doi:10.1016/0167-7322(96)00964-6.
  7. M. F. Crawford; H. L. Welsh; J. L. Locke (1949). "Infrared absorption of oxygen and nitrogen induced by intermolecular forces". Phys. Rev. 75 (10): 1607. Bibcode:1949PhRv...75.1607C. doi:10.1103/PhysRev.75.1607.
  8. Abel, Martin; Frommhold, Lothar; Li, Xiaoping; Hunt, Katharine L. C. (2011). "Computation of collision-induced absorption by dense deuterium-helium gas mixtures". Journal of Chemical Physics. 134 (7): 076101:1–076101:2. doi:10.1063/1.3556876. PMID   21341876.
  9. H. L. Welsh (1972). "3". In A. D. Buckingham; D. A. Ramsay (eds.). Pressure induced absorption spectra of hydrogen. Vol. III: Spectroscopy. Butterworths, London: MTP Internat. pp. 33–71.{{cite book}}: |work= ignored (help)
  10. Herzberg, G. (1952). "The atmospheres of the planets". Journal of the Royal Astronomical Society of Canada. 45: 100 a.
  11. Herzberg, G. (1952). "Spectroscopic evidence of molecular hydrogen in the atmospheres of Uranus and Neptune". The Astrophysical Journal. 115: 337. Bibcode:1952ApJ...115..337H. doi:10.1086/145552.
  12. A. A. Vigasin; Z. Slanina, eds. (1998). Molecular Complexes in Earth's, Planetary, Cometary, and Interstellar Atmospheres. Singapore: World Sci.
  13. C. Camy-Peyret; A. A. Vigasin, eds. (2003). Weakly Interacting Molecular Pairs: Unconventional Absorbers of Radiation in the Atmosphere. Vol. 27. Dordrecht. Kluwer. NATO Science Series, Earth and Environmental Sciences.
  14. A. Coustenis; F. W. Taylor (2008). Titan: Exploring an Earth-like world. World Scientific.
  15. Höpfner, M.; Milz, M.; Buehler, S.; Orphall, J.; Stiller, G. (24 May 2012). "The natural greenhouse effect of atmospheric oxygen (O2) and nitrogen (N2)". Geophysical Research Letters. 39 (L10706). Bibcode:2012GeoRL..3910706H. doi:10.1029/2012GL051409. ISSN   1944-8007. S2CID   128823108.
  16. S. Seager (2010). Exoplanet Atmospheres: Physical Processes. Series in Astrophysics. Princeton University Press.
  17. S. T. Hodgkin; B. R. Oppenheimer; N. C. Hambly; R. F. Jameson; S. J. Smart; I. A. Steele (2000). "Infrared spectrum of an extremely cool white-dwarf star". Nature. 403 (6765): 57–59. Bibcode:2000Natur.403...57H. doi:10.1038/47431. PMID   10638748. S2CID   4424397.
  18. H. L. Shipman (1977). "Masses, radii and model atmospheres for cool white-dwarf stars". Astrophys. J. 213: 138–144. Bibcode:1977ApJ...213..138S. doi:10.1086/155138.
  19. Saumon, Didier; Jacobson, S. B. (1999). "Pure hydrogen model atmospheres for very cool white dwarfs". The Astrophysical Journal. 511 (2): L107–110. arXiv: astro-ph/9812107 . Bibcode:1999ApJ...511L.107S. doi:10.1086/311851. S2CID   16199375.
  20. Bergeron, P.; Saumon, Didier; Wesemael, F. (April 1995). "New model atmospheres for very cool white dwarfs with mixed H/He and pure He compositions". The Astrophysical Journal. 443: 764. Bibcode:1995ApJ...443..764B. doi:10.1086/175566.
  21. Kilic, Mukremin; Leggett, S. K.; Tremblay, P.-E.; Hippel, Ted von; Bergeron, P.; Harris, Hugh C.; Munn, Jeffrey A.; Williams, Kurtis A.; Gates, Evalyn; Farihi, J. (2010). "A Detailed Model Atmosphere Analysis of Cool White Dwarfs in the Sloan Digital Sky Survey". The Astrophysical Journal Supplement Series. 190 (1): 77. arXiv: 1007.2859 . Bibcode:2010ApJS..190...77K. doi:10.1088/0067-0049/190/1/77. ISSN   0067-0049. S2CID   4571557.
  22. Agüeros, M. A.; Canton, Paul; Andrews, Jeff J.; Bergeron, P.; Kilic, Mukremin; Thorstensen, John R.; Curd, B.; Gianninas, A. (1 June 2015). "Ultracool white dwarfs and the age of the Galactic disc". Monthly Notices of the Royal Astronomical Society. 449 (4): 3966–3980. arXiv: 1503.03065 . doi:10.1093/mnras/stv545. ISSN   0035-8711. S2CID   119290935.
  23. Blouin, S.; Kowalski, P. M.; Dufour, P. (2017). "Pressure Distortion of the H2-He Collision-induced Absorption at the Photosphere of Cool White Dwarf Stars". The Astrophysical Journal. 848 (1): 36. arXiv: 1709.01394 . Bibcode:2017ApJ...848...36B. doi: 10.3847/1538-4357/aa8ad6 . ISSN   0004-637X. S2CID   118930159.
  24. Burrows, Adam; Hubbard, William B.; Lunine, Jonathan I.; Liebert, James (2001). "The theory of brown dwarfs and extrasolar giant planets". Rev. Mod. Phys. 73 (3): 719–765. arXiv: astro-ph/9706080 . Bibcode:2001RvMP...73..719B. doi:10.1103/revmodphys.73.719. S2CID   204927572.
  25. Saumon, Didier; Bergeron, P.; Lunine, Jonathan I.; Hubbard, W. B.; Burrows, Adam (1994). "Cool zero-metallicity stellar atmospheres". The Astrophysical Journal. 424: 333. Bibcode:1994ApJ...424..333S. doi: 10.1086/173892 .
  26. Saumon, Didier; Marley, Mark S.; Abel, Martin; Frommhold, Lothar; Freedman, Richard S. (2012). "New H2 collision-induced absorption and NH3 opacity and the spectra of the coolest brown dwarfs". The Astrophysical Journal. 750 (1): 74. arXiv: 1202.6293 . Bibcode:2012ApJ...750...74S. doi:10.1088/0004-637X/750/1/74. S2CID   11605094.
  27. Burgasser, Adam J.; Kirkpatrik, J. Davy; Burrows, Adam; Liebert, James; Reid, I. Neill; Gizis, John E.; McGovern, Mark R.; Prato, Lisa; McLean, Ian S. (2003). "The First Substellar Subdwarf? Discovery of a Metal-poor L Dwarf with Halo Kinematics". The Astrophysical Journal. 592 (2): 1186–1192. arXiv: astro-ph/0304174 . Bibcode:2003ApJ...592.1186B. doi:10.1086/375813. S2CID   11895472.
  28. Burgasser, Adam J.; Burrows, Adam; Kirkpatrik, J. Davy (2006). "A method for determining the physical properties of the coldest known brown dwarfs". The Astrophysical Journal. 639 (2): 1095–1113. arXiv: astro-ph/0510707 . Bibcode:2006ApJ...639.1095B. CiteSeerX   10.1.1.983.294 . doi:10.1086/499344. S2CID   9291848.
  29. B. M. S. Hansen; E. S. Phinney (1998). "Stellar forensics cooling curves". Mon. Not. R. Astron. Soc. 294 (4): 557–568. arXiv: astro-ph/9708273 . Bibcode:1998MNRAS.294..557H. doi: 10.1111/j.1365-8711.1998.01232.x .
  30. J. L. Linsky (1969). "On the pressure-induced opacity of molecular hydrogen in late-type stars". The Astrophysical Journal. 156: 989. Bibcode:1969ApJ...156..989L. doi:10.1086/150030.
  31. P. Lenzuni; D. F. Chernoff; E. Salpeter (1991). "Rosseland and Planck mean opacities of a zero-metallicity gas". Astrophys. J. 76: 759. Bibcode:1991ApJS...76..759L. doi:10.1086/191580.
  32. Th. H. Greif; V. Bromm; P. C. Clark; S. C. O. Glover; R. J. Smith; R. S. Klessen; N. Yoshida; V. Springel. (2012). "Formation and evolution of primordial protostellar systems". Mon. Not. R. Astron. Soc. 424 (1): 399–415. arXiv: 1202.5552 . Bibcode:2012MNRAS.424..399G. doi:10.1111/j.1365-2966.2012.21212.x.
  33. Richard, C.; Gordon, I. E.; Rothman, L. S.; Abel, Martin; Frommhold, Lothar; Gustafsson, M.; Hartmann, J. M.; Hermans, C.; Lafferty, W. J.; Orton, G.; Smith, K. M.; Tran, H. (2012). "New section of the HITRAN database: Collision-induced absorption (cia)". Journal of Quantitative Spectroscopy and Radiative Transfer. 113 (11): 1276–1285. Bibcode:2012JQSRT.113.1276R. doi:10.1016/j.jqsrt.2011.11.004.