Next Article in Journal
Hybrid Theory of Scattering and Its Applications
Next Article in Special Issue
Computation of Atomic Astrophysical Opacities
Previous Article in Journal
Energy Levels and Radiative Rates for Transitions in F-like Sc XIII and Ne-like Sc XII and Y XXX
Previous Article in Special Issue
Detailed Opacity Calculations for Astrophysical Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The ExoMol Atlas of Molecular Opacities

by
Jonathan Tennyson
* and
Sergei N. Yurchenko
Department of Physics and Astronomy, University College London, London WC1E 6BT, UK
*
Author to whom correspondence should be addressed.
Atoms 2018, 6(2), 26; https://doi.org/10.3390/atoms6020026
Submission received: 25 February 2018 / Revised: 3 May 2018 / Accepted: 4 May 2018 / Published: 10 May 2018
(This article belongs to the Special Issue Atomic and Molecular Opacity Data for Astrophysics)

Abstract

:
The ExoMol project is dedicated to providing molecular line lists for exoplanet and other hot atmospheres. The ExoMol procedure uses a mixture of ab initio calculations and available laboratory data. The actual line lists are generated using variational nuclear motion calculations. These line lists form the input for opacity models for cool stars and brown dwarfs as well as for radiative transport models involving exoplanets. This paper is a collection of molecular opacities for 52 molecules (130 isotopologues) at two reference temperatures, 300 K and 2000 K, using line lists from the ExoMol database. So far, ExoMol line lists have been generated for about 30 key molecular species. Other line lists are taken from external sources or from our work predating the ExoMol project. An overview of the line lists generated by ExoMol thus far is presented and used to evaluate further molecular data needs. Other line lists are also considered. The requirement for completeness within a line list is emphasized and needs for further line lists discussed.

1. Introduction

Molecular opacities dictate the atmospheric properties and evolution of a whole range of cool stars and all brown dwarfs. They are also important for the radiative transport properties of exoplanets. Even for a relatively simple molecule, the opacity function is generally complicated as it varies strongly with both wavelength and temperature. This is because molecules interact with light via a variety of different motions, namely rotational, vibrational and electronic, and the close-spacing of the energy levels leads to strong temperature dependence related to the thermal occupancy of the levels.
Molecular opacity functions, such as the ones provided by Sharp and Burrows [1], require very extensive input in the form of temperature-dependent spectroscopic data for atoms and molecules. At room temperature, the HITRAN database [2] provides comprehensive and validated spectroscopic parameters for molecules found in the Earth’s atmosphere and some other species. However, the complexity of molecular spectra increases rapidly with temperature and data from HITRAN and can only be used with extreme caution at higher temperatures [3]. The HITEMP database [4] is a HITRAN-style database designed to work at higher temperatures. However, HITEMP only contains spectroscopic line lists for five molecules and there are improved line lists available for each of these species, namely: H 2 O [5], CO 2 [6,7], CO [8], NO [9], and OH [10]. The opacity of all of these species are discussed further below.
The need for extensive, high temperature spectroscopic data on molecules, many of whom do not occur in the Earth’s atmosphere, has led to a number of systematic efforts to generate the molecular line lists required [7,11,12,13,14,15,16]. In particular, a number of groups have been progressively generating line lists for key molecules. These includes the TheoReTS [16] group (Reims/Tomsk), the NASA Ames team and our own ExoMol activity based in University College London, all of which used similar theoretical procedures discussed below, and an experimental initiative led by Bernath [14]. These activities have been significantly enhanced by the discovery of exoplanets and the requirement of extensive line lists to be used in exoplanet models and characterization [17,18].
As molecular line lists are extended, complicated and have a strong temperature dependence, it can be hard to understand a priori spectral regions where they show significant absorption. This work aims to summarize the absorption by the key molecular species required for cool stars, brown dwarfs and exoplanets. Many of the line lists used are taken from those computed by the Exomol project, but these are augmented by lists taken from other sources to give a comprehensive set, or atlas, of absorbing species.
In this work, we present a collection (atlas) of molecular opacities for 52 molecules (130 isotopologues) generated using line lists from the ExoMol database. The opacities are presented in the form of cross sections for two reference temperatures, 300 K and 2000 K using the Doppler (zero-pressure) broadening. The goal of this atlas is to present in a simple, visual form an overview of spectroscopic coverage, main spectroscopic signatures as well as temperature dependence of the molecular opacities relevant for atmospheric studies of hot exoplanets and stars.

2. Methodology

The general methodology used by the ExoMol project is very similar to that used by both the TheoReTS and NASA Ames group for producing molecular line lists. It has been well-described elsewhere [19,20] and below we will discuss only general considerations.
Our technique involves the following general steps: (1) computing accurate ab initio potential energy and dipole surface moment surfaces, (2) improving the potential energy surface (PES) using available high-resolution spectroscopic data and (3) generating a comprehensive line list using the improved PES, ab initio dipole moment surface (DMS) and an appropriate variational nuclear motion program [21,22,23]. Development of these program specifically for opacity calculations has been discussed by us elsewhere [24].
Performing such calculations is always a trade-off between completeness and accuracy. A complete line list should contain information on possible transitions at all wavelengths and at all temperatures. An accurate one should reproduce line positions and transition probabilities to the high standards of laboratory high resolution spectroscopy. In practice, for most systems, it is hard if not impossible to achieve both of these goals simultaneously so compromises have to be made. Experience shows that completeness is essential [25] while accuracy is something to strive for and, of course, is essential for high resolution spectroscopic studies [26].
Starting from an ab initio PES, there are three ways of improving the accuracy of the line positions. The first method is to fit the potential to measured transitions frequencies, empirical energies or a mixture of both. This is now a standard technique that is widely employed. It is capable of giving very good results [27] if in most cases not actual experimental accuracy. The second technique is to adjust the vibrational band origins during the calculation. This technique works only if the rotation–vibration basis used is a simple product between vibrational and rotational functions. A basis in this form is used by the polyatomic nuclear motion program TROVE [23] and the vibrational basis is contracted by performing an initial rotationless ( J = 0 ) calculations. At this stage, the band origins can be shifted to their empirical value [28]. Finally, the ExoMol data structure [29] presents a line list as a states file with energy levels and associated quantum numbers and a highly compact transitions file of Einstein A coefficients. This structure therefore allows computed energy levels to be replaced with empirical ones after the calculation is complete; indeed, this can be done some time after the line list is computed if new empirical data becomes available [30]. This allows many transition frequencies to be generated with experimental accuracy, including ones that have not actually been measured.
To provide lists of accurate empirical energy levels, Furtenbacher et al. [31,32] developed the measured active rotation vibration energy level procedure (MARVEL). Table 1 gives a summary of molecules of astronomical importance for which the MARVEL procedure has been applied. Overviews of this methodology can be found elsewhere [33,34].
The line lists generated by the methods discussed above form the input to opacity calculations. To judge the potential influence of each molecule on the opacity, one can generate cross sections as a function of wavelength and temperature. As several of the line lists considered contain many billions of lines (see Table 2), calculating cross sections can be computationally expensive. For this reason, we have developed a highly optimized program ExoCross [48], which generates wavelength-dependent cross sections as function of temperature and pressure. Here, we use ExoCross to systematically generate cross sections for the key species that are important for opacities of cool stars, brown dwarfs and exoplanets.
In generating these cross sections, we consider the main (parent) isotopologue of each species, which is taken as being in 100 % abundance. For simplicity, we use a Doppler profile on a wavenumber grid of 1 cm 1 . The cross sections have been generated using the methodology by Hill et al. [49]. The cross sections can be also obtained at higher resolutions (up to 0.01 cm 1 ) using the cross sections App at www.exomol.com.
Cross sections were generated at two standard temperatures of 300 K and 2000 K except when the molecule is expected to be entirely in the condensed phase at 300 K. For a few, key strongly bound species, we also consider the cross section at 5000 K, which is near the upper limit of the temperature where molecules make a contribution to opacities.

3. Results

Table 2 summarises the molecules for which line lists have been provided as part of the ExoMol project. Sources for line lists of other key species are given in Table 3. Data for all these species, including cross sections, are available on the ExoMol website.
Figure 1, Figure 2, Figure 3, Figure 4, Figure 5, Figure 6, Figure 7, Figure 8, Figure 9, Figure 10, Figure 11, Figure 12, Figure 13, Figure 14 and Figure 15 display temperature-dependent cross sections for each of these species. For all molecules except water, only results for the major (parent) isotopologue are given. As specified in Table 2, many ExoMol line lists also consider isotopically substituted species. In most cases, isotopic substitution leads to shifts in spectroscopic features that are observable at medium to high resolution but result in no fundamental change in structure of the spectrum. The exception is where this substitution leads to breaking of the symmetry, which can lead to new features due to changes in the way the Pauli principle applies, examples here include 12 C 12 C to 12 C 13 C, and changes in vibrational structure such as those encountered on moving from H 2 O to HDO.
For most species, absorption cross sections are plotted at two temperatures: 300 K and 2000 K. Exceptions are where the species is unlikely to have significant vapor pressure at 300 K, when this curve has been omitted, or when the species unlikely to survive at 2000 K, in which case an appropriate, lower temperature is used. The figures are grouped according to whether the line lists concerned cover the infrared only (wavenumbers below 12,000 cm 1 ), extend into the visible (wavenumbers below 20,000 cm 1 ) or cover the near ultraviolet (wavenumbers below 35,000 cm 1 ). Beyond that, the figures are grouped to contain roughly similar species. Each of the figures and species are considered in turn below.
H 2 O: the spectrum of water (see Figure 1) is important in a whole range of astronomical objects [14] including M-dwarfs [94] and exoplanets [95]. The ExoMol line lists for H 2 18 O and H 2 17 O complement the H 2 16 O BT2 line list [96], which has been widely used for astronomical studies including, for instance, as the foundation of the BT-Settl cool star model atmosphere [97]. A comprehensive line list for HDO is also available [93]. In fact, we have recently updated BT2 with a new H 2 16 O line list known as POKAZATEL [5]. Besides greatly improving the accuracy of the previous line list, it also extends the temperature range beyond 3000 K, which BT2 was designed for. POKAZATEL has been tested against laboratory spectra recorded in flames and was found to perform very well [98]. On the scale of the figure, BT2 and POKAZATEL are very similar for temperatures below 2000 K, although the detailed line positions given by POKAZATEL are considerably more accurate (see the laboratory study by Campargue et al. [99], for example). However, at high temperatures, POKAZATEL is much more complete. What is really noticeable is how flat the water opacity becomes at high temperatures.
Figure 2 shows cross sections for a number of other hydrides: methane, silane, phosphine, ammonia and ethylene.
CH 4 : Figure 2 shows the methane cross sections generated by the 10to10 line list, which contains almost 10 10 transitions. Other line lists available for methane are notably the TheoReTS line list(s) [100,101], and a comprehensive experimental line list from Hargreaves et al. [102,103]. The 10to10 line list has been used to identify spectroscopic features in T-dwarfs [104]. Model calculations have shown that reproducing the atmospheric opacity of these methane-rich brown dwarfs requires the explicit consideration of many billions of transitions [25]. As the temperature range of 10to10 is limited, we have recently extended it with a new 34to10 line list, which contains 34 × 10 10 transitions [105]. To make use of this line list tractable in radiative transport models, the weak lines are amalgamated into so-called super-lines [16]. The temperature coverage for the various methane line lists, which broadly agree with each other, is probably now adequate for astrophysical purposes. However, there is a need for better coverage at shorter wavelengths (< 0.85 μ m).
PH 3 : the phosphine cross sections (see Figure 2) show a very pronounced feature about 4.5 μ m, which displays only weak dependence on temperature. Phosphine is a well-known component of solar system gas giants [106], and it can be anticipated that this feature will at some point be used to identify PH 3 in exoplanets.
NH 3 : the spectrum of ammonia is given in Figure 2. Ammonia spectra are well known in brown dwarfs [104] and are thought to provide the key signature for coolest class of these species known as Y-dwarfs. Plotted is the BYTe line list, which is significantly less accurate than most of the other line lists considered here and is really only complete for wavenumbers below 10,000 cm 1 and T < 1500 K. A new ExoMol line list that should help resolve these issues is currently under construction. A line list for 15 NH 3 is also available [107].
C 2 H 4 : the final spectrum given in Figure 2 is of ethylene. Only one temperature is plotted since, despite containing 60 billions line, the line list is only complete up to 700 K. However, it is likely that ethylene will decompose at higher temperatures. Although not shown here, the main features of the ethylene spectrum show an unusually small sensitivity to temperature [75]. Rey et al. [108] have also computed a far-infrared ethylene line list.
Figure 3 shows cross sections for a variety of polyatomic oxides plus hydrogen cyanide.
H 2 O 2 : hydrogen peroxide cross sections are given in Figure 3. It should be noted that even at room temperature the line list for H 2 O 2 given by the current release of HITRAN [2] is not complete, as the strong mid-infrared absorptions are missing. This is because of the difficulty of analyzing the observed spectra in this region.
HCN: the line list shown in Figure 3 is unusual for two reasons. First, it contains transition sets for two isomers, HCN and HNC, which are often considered to be distinct molecules and second it is not new; rather it is a reworking of earlier line lists [110,111] based on extensive compilations of empirical energy levels due to Mellau [112,113,114]. A similar empirically-informed line list is available for H 13 CN [115]. The updated Harris line list was used to tentatively identify HCN in super-Earth exoplanet 55 Cancri e [116]. However, the 1.54 μ m feature used for this detection is close to a similar acetylene feature. As discussed below, there is still no good line list for hot acetylene. The opacity of HCN is particularly important for reliable models of cool carbon stars where its inclusion leads to fundamental changes in the stellar structure [117]. The ratio between HCN and HNC can provide a potential thermometer the atmospheres of cool stars [118].
SO 2 : sulphur dioxide cross sections are given in Figure 3. SO 2 is thought to be an important component of oxygen-rich exoplanet atmospheres [119,120] and to have been important in both early-Earth [121] and early-Mars [122].
SO 3 : sulphur trioxide cross sections are given in Figure 3. SO 3 can be hard to detect because it usually appears in conjunction with SO 2 , which absorbs strongly in the same region. Interestingly, our cross sections suggest that it may be easier to distinguish SO 3 from SO 2 at higher temperatures as, while the spectrum SO 3 retains much of it structure at higher temperatures, the spectrum of SO 2 becomes increasingly flattened. The line list is only complete up to 800 K and this has been used as the upper temperature.
H 2 CO: Figure 3 shows temperature dependent cross sections of formaldehyde.
CO 2 : carbon dioxide cross sections given in Figure 3 are obtained from the Ames-2016 (4000 K) line list of Huang et al. [7,123], for the main isotopologue ( 12 C 16 O 2 ) and covers rotational excitations up to J = 220. The line list is based on an empirical potential energy surface and accurate variational nuclear motion calculations. The Ames-2016 database also contains line lists for 12 other isotopologues of CO 2 .
Figure 4 shows cross sections of HNO 3 , CH 3 Cl and C 2 H 2 .
HNO 3 : nitric acid, whose spectrum is shown in Figure 4, is not likely to be an important source of opacity. However, its spectroscopic signature can be clearly observed in the Earth’s atmosphere from space [3,124,125] and thus HNO 3 could be a possible biomarker on exoplanets. Like H 2 O 2 , the HITRAN data for HNO 3 is incomplete for frequencies above about 1000 cm 1 .
CH 3 Cl: methyl chloride is also a potential biosignature in the search for life outside of the solar system [126]. ExoMol’s mew line list for methyl chloride covers the two major isomers, CH 3 35 Cl and CH 3 37 Cl.
C 2 H 2 : acetylene cross sections given in Figure 4 are generated using the ASD-1000 database [90]. ASD-1000 was generated using an Effective Hamiltonian and contains about 34,000,000 lines. This would appear to be too few lines for the line list to be complete at elevated temperatures as normally, hot line lists for tetratomics contain billions of transitions in order to be complete for high temperatures. A variational acetylene line list is currently under construction as part of the ExoMol project.
H 2 S: Figure 5 shows that hydrogen sulphide has a rather unusual spectrum. Transitions involving the vibrational fundamentals are anomalously weak [127] for this system, which shifts the peak of the vibration–rotation spectrum to shorter wavelengths than is usual: the figure shows that the strongest vibration band is an overtone lying in the 2.5 μ m region.
Figure 6 gives cross sections for NS as well as the alkaline earth monohydrides MgH and CaH. Together with BeH, MgH and CaH were the first set of species considered by ExoMol. These ExoMol line lists only contain pure rotation and rotation–vibration transitions within the X 2 Σ + electronic ground states for each of these species.
NS: NS is one of the first ten diatomic molecules to be detected in space [128,129].
MgH: the A 2 Π –X 2 Σ + band is also important for magnesium monohydride as, in particular, its use of major importance for establishing isotopic abundances of Mg in stars [130,131,132,133,134]. Weck et al. [13] constructed a purely ab initio line list, which covers this band but is of limited accuracy. More recently, experimentally-driven studies [135,136] have sought to remedy this problem.
CaH: A 2 Π –X 2 Σ + band is also important in cool stars and brown dwarfs [137].
Figure 7 shows cross sections covering electronic spectra of three diatomic monohydrides, BeH, AlH and CH.
BeH: Darby-Lewis et al. [138] have recently developed a new model for beryllium monohydride, which both improves (marginally) on the earlier Yadin line list [51] and, more significantly, includes a treatment of the A 2 Π –X 2 Σ + band. This band has been observed in sunspot [139], but the motivation for extending the line list was actually fusion plasmas where the use of Be walls has led to the observation of A–X emissions in the plasma [140]. Darby-Lewis et al. demonstrate that their new line list gives excellent reproduction of emission spectra of both BeH and BeD, as well as predictions for the spectrum of BeT, which is important for fusion studies.
AlH: it is known to be present in sunspots through lines in its A 1 Π –X 1 Σ + electronic band which lie in the blue [141]; AlH was also recently detected around Mira-variable o Ceti by Kaminski et al. [142]. The ExoMol line list for AlH contains these electronic transitions as well as ground state lines.
CH: an extensive, empirical line list is available for CH due to Masseron et al. [83]. This line list covers transitions within the X 2 Π electronic ground state as well as rovibronic transitions within the A 2 Δ –X 2 Π , B 2 Σ –X 2 Π and C 2 Σ + –X 2 Π bands. Masseron et al. [83] also provide a line list for 13 CH.
Figure 8 shows temperature dependent cross sections for HCl, HS, CrH, and OH.
HCl: the line list used to generate the HCl cross sections given in Figure 8 was generated alongside line lists for other hydrogen halides, namely HF, HBr and HI, using semi-empirical methods by Li. et al. [86,143].
Figure 9 shows cross sections for NaCl, KCl, PO, PN and CS.
NaCl and KCl: cross sections are shown in Figure 9. These species have large transition dipoles and hence strong transitions, and they are thought to be of importance for studies of hot super-Earths [20].
PO: phosphorus monoxide has been detected in a number of locations in space: red Supergiant Star VY Canis Majoris [144], in the wind of the oxygen-rich AGB star IK Tauri [145], and in star-forming regions [146,147]. The ExoMol line list covers transitions within the X 2 Π ground electronic state.
PN: it is an important molecule used to probe different regions of the interstellar medium (ISM). The ExoMol line list for PN covers just pure rotation–vibration transitions.
CS: cross sections in Figure 9 illustrate the temperature dependence of the CS spectra as generated using the comprehensive vibration–rotation ExoMol line lists for eight isotopologues of carbon monosulphide (CS) ( 12 C 32 S, 12 C 33 S, 12 C 34 S, 12 C 36 S, 13 C 32 S, 13 C 33 S, 13 C 34 S, 13 C 36 S) in their ground electronic states.
Figure 10 shows cross sections for the important species CO plus CN, CP and CaO.
CO: cross sections for carbon monoxide are given in Figure 10. CO is ubiquitous throughout the spectra of cool stars [148] and brown dwarfs, and can clearly be seen in exoplanets [149,150]. CO is a very strongly bound species so the molecule survives at relatively high temperatures. As can be seen from the figure, its broadband spectrum changes from comprising a series of sharp bands to becoming quasi-continuous as the temperature is raised. Unlike the other species shown in this figure, the CO cross sections through the visible are simply given by rotation–vibration transitions. We note that Li et al. [8] improved their intensities by refining their CO ab initio dipole moment curve by fitting to experiment. This is not the usual practice [151], but Li et al. found that they could not get satisfactory results using a purely ab initio approach.
CN and CP: cross sections are shown in Figure 10. These two open shell or radical molecules have the same electronic structure and their spectra are dominated by the A 2 Π –X 2 Σ + and B 2 Σ + –X 2 Σ + bands, which for CN lie in the red and violet regions of the visible. For the heavier CP radical, these bands are shifted to longer wavelengths.
CaO: cross sections for calcium oxide are given in Figure 10. CaO is unlikely to exist in the gas phase in dense atmospheres at 300 K, this spectrum is shown for the consistency. CaO has yet to be detected in space, but it is thought of as a possible component of super-Earth exoplanet atmospheres. CaO possesses transitions with notably large cross sections, which should facilitate its detection.
Figure 11 shows cross sections for NO and PS.
NO and PS: are both radicals with X 2 Π electronic ground states. NO is an important species in the Earth’s atmosphere and is likely to be so in oxygen-rich exoplanets. The current ExoMol line list [9] supersedes the one provided by HITEMP [4]. For NO, the transitions considered in the figure are all within the X 2 Π ground state manifold, while, for PS, the B 2 Π –X 2 Π electronic band is also considered. The equivalent electronic transition for NO, which is actually designated the A 2 Π – X 2 Π band, lies well into the ultra violet at about 230 nm [152].
Figure 12 shows cross sections for LiH, ScH, FeH and NH.
LiH cross sections are shown in Figure 12. They are based on ab initio calculations by Coppola et al. [80]. LiH is a four electron system so purely theoretical procedures should give reliable results. The spectrum of LiH is important for possible applications of the lithium test in brown dwarfs [153]. We note that very recently Bittner and Bernath [154] have constructed line lists for both LiF and LiCl.
ScH cross sections are shown in Figure 12. These were also based on an essentially ab initio line list, which contains an empirical adjustment for the band positions. Unlike LiH, ScH is a many-electron system with an open d-shell; such systems present considerable challenges from an ab initio perspective [155,156]; the provision of high accuracy line list for ScH will probably require further experimental input.
FeH cross sections, shown in Figure 12, are conversely based on experimental studies. FeH is an important stellar species, which has been observed in M and L dwarfs [87,157,158] as well as sunspots [159,160,161]. There is a strong need to use these laboratory studies to construct a rigorous theoretical model for the system, which can be used to generate line lists over an extended range of temperatures and wavelengths; however, ab initio calculations on the system remain a challenge [162].
NH: an empirical line list for NH from the Bernath group [82] covering rotation–vibrational transitions within the X 3 Σ ground electronic state.
Figure 13 shows cross sections for SiH, NaH and TiH.
SiH: is an accurate line list covering rotation–vibration transitions within the ground X 2 Π electronic state as well as transitions to the low-lying A 2 Δ and a 4 Σ states.
NaH cross sections are shown in Figure 13. NaH remains undetected in stellar spectra but is thought to be an important opacity source in M-dwarfs [164]. However, as our cross sections show the blue A 2 Σ + –X 2 Σ + band is extended and gives a rather flat band shape. This will make the clear detection of NaH in a stellar atmosphere difficult from its main electronic band.
TiH cross sections shown in Figure 13 represent an empirical line list for TiH from Burrows et al. [88].
Figure 13 shows cross sections for SiO, AlO, VO and TiO.
SiO cross sections are shown in Figure 14. SiO is a strongly bound molecule whose infrared spectrum is well-known in stars [165] and sunspots [166,167]. Its electronic spectrum has also been observed in sunspots [168]. At present, the ExoMol project has only provided a long-wavelength line list for SiO based on its temperature-dependent vibration–rotation and rotation spectrum. Work is in progress providing a companion line list covering its electronic bands starting from ab initio potential energy curves recently calculated by Bauschlicher Jr. et al. [169]. In order to illustrate the UV spectra of SiO, Figure 14 also shows cross sections from Kurucz’z database [170].
AlO: the spectrum of aluminium monoxide is illustrated in Figure 14. This spectrum is actually important for plasma diagnostics [171,172]. Rather remarkably, the only previous AlO line list [173], which was prepared for plasma studies, only provided relative as opposed to absolute line intensities.
VO cross sections are shown in Figure 14. The VO spectrum is heavily overlapped by that of TiO and the two molecules often occur together. VO spectral signatures are important for classifying late M dwarfs [174,175], and it is one of the dominant species in the spectra of young hot brown dwarfs [175,176,177]. VO is thought to be present in the atmosphere of hot Jupiter exoplanets [178].
TiO cross sections are shown in Figure 14. TiO is a well-known and important molecule in the atmosphere of M-dwarf stars [179]. Detections in the atmospheres of exoplanets are also beginning to be reported [180,181,182]. The figure is based on the line list of Schwenke et al. [89]; there is also a line list due to Plez [12], which has been updated in a recent release of the VALD database [183]. However, there are known problems with the available spectroscopic data on TiO (see Hoeijmakers et al. [26], for example). An ExoMol TiO line list is currently nearing completion [184]; this line list will make use of the MARVEL study on TiO [43], which leads to more accurate transition frequencies than are provided by the currently available line lists.
Figure 15 gives cross sections for the open shell diatomic species C 2 and of the important molecular ion H 3 + .
C 2 has a long spectroscopic history [185,186]. In astronomical objects, its spectrum has been observed via at least six distinct electronic bands, with other band systems required to explain observed populations. An empirical line list for the Swan system was provided by Brooke et al. [187]; the figure shows cross sections generated using the newly constructed ExoMol line list [77], which considers transitions between the eight lowest electronic states in the system. Given the many perturbations between the electronic states that are hard to model and that there are astronomically observed bands not covered by this eight-state model, further work on C 2 will undoubtedly be needed.
H 3 + : the spectrum of H 3 + given in Figure 15 is based on ab initio calculations, which have proved themselves to be very accurate for this molecular ion [188]. H 3 + is an unusual species in that it has no (allowed [189]) rotational spectrum and no known electronic spectrum leaving only rotation–vibration transitions. These are well-known in the ionospheres of solar system gas giant planets [190] but have yet to be observed in hot Jupiter exoplanets [191], where they might be thought to be prominent. However, the cooling provided by H 3 + infrared emissions appears to be the key for the stability of atmospheres of giant extrasolar planets close to their host star [192]. This has made H 3 + cooling functions of great importance [193,194]. Similarly H 3 + has yet to observed in stellar or brown dwarf spectra, but H 3 + is the main source of electrons in objects such as cool white dwarfs [195]. The partition function of H 3 + [196] thus controls the amount of H present, which, in turn, dominates the opacity. A line list for H 2 D + [197] is also available.

4. Conclusions

Compiling molecular opacity functions requires a large range of spectroscopic data on a large range of molecules. As discussed in this article, for many key species, there are now extensive line lists available that can be used to compute temperature-dependent opacity functions. This is a process of constant improvement and immediate improvements are indeed identified for several species discussed above. Here, we consider what species may be missing from the current compilations.
In their validation of the well-used BT–Settl model [97] atmospheres for M-dwarfs, Rajpurohit et al. [164] identified only AlH, NaH and CaOH as key species for which data was missing. In response to this, the ExoMol project has provided spectroscopic line lists for AlH [76] and NaH [59]; this leaves CaOH, which has a strong band in the visible at 557 nm, as the one key missing species.
The chemistry of oxygen-rich M-dwarf stars is somewhat simpler than that of carbon-rich stars. There are a number of carbon-containing species for which reliable line lists are not available, notably acetylene (HCCH) and C 3 . Acetylene spectra are well-studied experimentally [44,198,199] and recently Lyulin and Perevalov [90] produced an effective-Hamiltonian based, empirical 12 C 2 H 2 line list. Work on an ExoMol acetylene line list is nearing completion. Jørgensen et al. [200] produced an early, purely ab initio line list for C 3 . They demonstrated the importance of this species for models of cool carbon stars, but their line list cannot be considered reliable by modern standards where ab initio methods have improved considerably and tuning models to experimental data is now routine. The construction of an improved C 3 line list is underway as part of the ExoMol project. Both ExoMol and TheoReTS are considering line lists for higher hydrocarbons that are needed for studies of exoplanets.
Transition metal containing diatomics provide strong sources of opacities since these open shell species often display strong electronic transitions in the near infrared and visible i.e., near the peak of the stellar flux for a cool star. A few of these species, notably TiO, VO, FeH, ScH and TiH, are considered above but there are many more possible candidates. These include CrH [201], MnH [202] and MgO [78] for which ExoMol has near complete line lists, and species such NiH, ZrO, YO, FeO. Carbides, nitrides and even sulfide species may also need to be considered in due course. Performing accurate ab initio electronic structure calculation on transition metal containing molecules remains very difficult [155,156].
Recent observations have identified a whole new class of exoplanets with masses somewhat larger than the Earth’s and orbits close to their host stars. These hot rocky super-Earths, lava planets or magma planets as they are variously known as are just beginning to be characterized [116]. Even though at this point the molecular composition of their atmosphere is essentially unknown, study of these bodies will create demands for spectroscopic data and opacities for a whole range of new molecules. We recently reviewed the data needs for hot super-Earth exoplanets [20] and refer any interested reader to that article.
Finally, line broadening must be mentioned. Molecular transitions undergo broadening due to temperature and pressure effects. While temperature-induced Doppler broadening is straightforward to model, pressure broadening is more difficult because its importance is, at least in principle, different for every transition. Broadening parameters for key collision broadeners such as H 2 and He at elevated temperatures are largely missing. Approximate methods are being used to fill this void for key molecules such as water [182,203,204]. We have implemented a systematic and complete if approximate broadening “diet” [205] as part of the ExoMol data base [50]. This allows the generation of both temperature- and pressure-dependent cross sections using our program ExoCross [48]. However, there remains much work to be done in both improving temperature-dependent models for line broadening and providing the appropriate broadening parameters.

Author Contributions

Both authors contributed equally.

Acknowledgments

We thank the members of the ExoMol team for their work on this project. The ExoMol project was funded by the ERC Advanced Investigator Project 267219 and is currently supported by UK STFC under grant ST/M001334/1.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sharp, C.M.; Burrows, A. Atomic and Molecular Opacities for Brown Dwarf and Giant Planet Atmospheres. Astrophys. J. Suppl. 2007, 168, 140. [Google Scholar] [CrossRef]
  2. Gordon, I.E.; Rothman, L.S.; Hill, C.; Kochanov, R.V.; Tan, Y.; Bernath, P.F.; Birk, M.; Boudon, V.; Campargue, A.; Chance, K.V.; et al. The HITRAN 2016 molecular spectroscopic database. J. Quant. Spectrosc. Radiat. Transf. 2017, 203, 3–69. [Google Scholar] [CrossRef]
  3. Schreier, F.; Staedt, S.; Hedelt, P.; Godolt, M. Transmission Spectroscopy with the ACE-FTS Infrared Spectral Atlas of Earth: A Model Validation and Feasibility Study. Mol. Astrophys. 2018, 11, 1–22. [Google Scholar] [CrossRef]
  4. Rothman, L.S.; Gordon, I.E.; Barber, R.J.; Dothe, H.; Gamache, R.R.; Goldman, A.; Perevalov, V.I.; Tashkun, S.A.; Tennyson, J. HITEMP, the High-Temperature Molecular Spectroscopic Database. J. Quant. Spectrosc. Radiat. Transf. 2010, 111, 2139–2150. [Google Scholar] [CrossRef]
  5. Polyansky, O.L.; Kyuberis, A.A.; Zobov, N.F.; Tennyson, J.; Yurchenko, S.N.; Lodi, L. ExoMol molecular line lists XXX: A complete high-accuracy line list for water. Mon. Not. R. Astron. Soc. 2018, in press. [Google Scholar]
  6. Tashkun, S.A.; Perevalov, V.I. CDSD-4000: High-resolution, high-temperature carbon dioxide spectroscopic databank. J. Quant. Spectrosc. Radiat. Transf. 2011, 112, 1403–1410. [Google Scholar] [CrossRef]
  7. Huang, X.; Schwenke, D.W.; Freedman, R.S.; Lee, T.J. Ames-2016 Line Lists for 13 Isotopologues of CO2: Updates, Consistency, and Remaining Issues. J. Quant. Spectrosc. Radiat. Transf. 2017, 203, 224–241. [Google Scholar] [CrossRef]
  8. Li, G.; Gordon, I.E.; Rothman, L.S.; Tan, Y.; Hu, S.M.; Kassi, S.; Campargue, A.; Medvedev, E.S. Rovibrational line lists for nine isotopologues of the CO molecule in the X 1Σ+ ground electronic state. Astrophys. J. Suppl. 2015, 216, 15. [Google Scholar] [CrossRef]
  9. Wong, A.; Yurchenko, S.N.; Bernath, P.; Mueller, H.S.P.; McConkey, S.; Tennyson, J. ExoMol Line List XXI: Nitric Oxide (NO). Mon. Not. R. Astron. Soc. 2017, 470, 882–897. [Google Scholar] [CrossRef]
  10. Brooke, J.S.A.; Bernath, P.F.; Western, C.M.; Sneden, C.; Afşar, M.; Li, G.; Gordon, I.E. Line strengths of rovibrational and rotational transitions in the ground state of {OH}. J. Quant. Spectrosc. Radiat. Transf. 2016, 138, 142–157. [Google Scholar] [CrossRef]
  11. Jørgensen, U.G.; Larsson, M.; Iwamae, A.; Yu, B. Line intensities for CH and their application to stellar atmospheres. Astron. Astrophys. 1996, 315, 204. [Google Scholar]
  12. Plez, B. A new TiO line list. Astron. Astrophys. 1998, 337, 495–500. [Google Scholar]
  13. Weck, P.F.; Schweitzer, A.; Stancil, P.C.; Hauschildt, P.H.; Kirby, K. The molecular line opacity of MgH in cool stellar atmospheres. Astrophys. J. 2003, 582, 1059–1065. [Google Scholar] [CrossRef]
  14. Bernath, P.F. Molecular astronomy of cool stars and sub-stellar objects. Int. Rev. Phys. Chem. 2009, 28, 681–709. [Google Scholar] [CrossRef]
  15. Tennyson, J.; Yurchenko, S.N. ExoMol: Molecular line lists for exoplanet and other atmospheres. Mon. Not. R. Astron. Soc. 2012, 425, 21–33. [Google Scholar] [CrossRef]
  16. Rey, M.; Nikitin, A.V.; Babikov, Y.L.; Tyuterev, V.G. TheoReTS—An information system for theoretical spectra based on variational predictions from molecular potential energy and dipole moment surfaces. J. Mol. Spectrosc. 2016, 327, 138–158. [Google Scholar] [CrossRef]
  17. Tinetti, G.; Vidal-Madjar, A.; Liang, M.C.; Beaulieu, J.P.; Yung, Y.; Carey, S.; Barber, R.J.; Tennyson, J.; Ribas, I.; Allard, N.; et al. Water vapour in the atmosphere of a transiting extrasolar planet. Nature 2007, 448, 169–171. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Tinetti, G.; Encrenaz, T.; Coustenis, A. Spectroscopy of planetary atmospheres in our Galaxy. Astron. Astrophys. Rev. 2013, 21, 1–65. [Google Scholar] [CrossRef]
  19. Tennyson, J. Accurate variational calculations for line lists to model the vibration rotation spectra of hot astrophysical atmospheres. WIREs Comput. Mol. Sci. 2012, 2, 698–715. [Google Scholar] [CrossRef]
  20. Tennyson, J.; Yurchenko, S.N. Laboratory spectra of hot molecules: Data needs for hot super-Earth exoplanets. Mol. Astrophys. 2017, 8, 1–18. [Google Scholar] [CrossRef]
  21. Yurchenko, S.N.; Lodi, L.; Tennyson, J.; Stolyarov, A.V. Duo: A general program for calculating spectra of diatomic molecules. Comput. Phys. Commun. 2016, 202, 262–275. [Google Scholar] [CrossRef]
  22. Tennyson, J.; Kostin, M.A.; Barletta, P.; Harris, G.J.; Polyansky, O.L.; Ramanlal, J.; Zobov, N.F. DVR3D: A program suite for the calculation of rotation–vibration spectra of triatomic molecules. Comput. Phys. Commun. 2004, 163, 85–116. [Google Scholar] [CrossRef]
  23. Yurchenko, S.N.; Thiel, W.; Jensen, P. Theoretical ROVibrational Energies (TROVE): A robust numerical approach to the calculation of rovibrational energies for polyatomic molecules. J. Mol. Spectrosc. 2007, 245, 126–140. [Google Scholar] [CrossRef]
  24. Tennyson, J.; Yurchenko, S.N. The ExoMol project: Software for computing molecular line lists. Int. J. Quantum Chem. 2017, 117, 92–103. [Google Scholar] [CrossRef]
  25. Yurchenko, S.N.; Tennyson, J.; Bailey, J.; Hollis, M.D.J.; Tinetti, G. Spectrum of hot methane in astronomical objects using a comprehensive computed line list. Proc. Natl. Acad. Sci. USA 2014, 111, 9379–9383. [Google Scholar] [CrossRef] [PubMed]
  26. Hoeijmakers, H.J.; de Kok, R.J.; Snellen, I.A.G.; Brogi, M.; Birkby, J.L.; Schwarz, H. A search for TiO in the optical high-resolution transmission spectrum of HD 209458b: Hindrance due to inaccuracies in the line database. Astron. Astrophys. 2015, 575, A20. [Google Scholar] [CrossRef]
  27. Partridge, H.; Schwenke, D.W. The determination of an accurate isotope dependent potential energy surface for water from extensive ab initio calculations and experimental data. J. Chem. Phys. 1997, 106, 4618–4639. [Google Scholar] [CrossRef]
  28. Yurchenko, S.N.; Barber, R.J.; Tennyson, J.; Thiel, W.; Jensen, P. Towards efficient refinement of molecular potential energy surfaces: Ammonia as a case study. J. Mol. Spectrosc. 2011, 268, 123–129. [Google Scholar] [CrossRef]
  29. Tennyson, J.; Hill, C.; Yurchenko, S.N. Data structures for ExoMol: Molecular line lists for exoplanet and other atmospheres. AIP Conf. Proc. 2013, 1545, 186–195. [Google Scholar]
  30. Barber, R.J.; Strange, J.K.; Hill, C.; Polyansky, O.L.; Mellau, G.C.; Yurchenko, S.N.; Tennyson, J. ExoMol line lists—III. An improved hot rotation–vibration line list for HCN and HNC. Mon. Not. R. Astron. Soc. 2014, 437, 1828–1835. [Google Scholar] [CrossRef]
  31. Furtenbacher, T.; Császár, A.G.; Tennyson, J. MARVEL: Measured active rotational-vibrational energy levels. J. Mol. Spectrosc. 2007, 245, 115–125. [Google Scholar] [CrossRef]
  32. Furtenbacher, T.; Császár, A.G. MARVEL: Measured active rotational-vibrational energy levels. II. Algorithmic improvements. J. Quant. Spectrosc. Radiat. Transf. 2012, 113, 929–935. [Google Scholar] [CrossRef]
  33. Császár, A.G.; Czakó, G.; Furtenbacher, T.; Mátyus, E. An active database approach to complete rotational–vibrational spectra of small molecules. Annu. Rep. Comput. Chem. 2007, 3, 155–176. [Google Scholar]
  34. Tennyson, J.; Bernath, P.F.; Brown, L.R.; Campargue, A.; Császár, A.G.; Daumont, L.; Gamache, R.R.; Hodges, J.T.; Naumenko, O.V.; Polyansky, O.L.; et al. A Database of Water Transitions from Experiment and Theory (IUPAC Technical Report). Pure Appl. Chem. 2014, 86, 71–83. [Google Scholar] [CrossRef]
  35. Tennyson, J.; Bernath, P.F.; Brown, L.R.; Campargue, A.; Carleer, M.R.; Császár, A.G.; Gamache, R.R.; Hodges, J.T.; Jenouvrier, A.; Naumenko, O.V.; et al. IUPAC critical Evaluation of the Rotational-Vibrational Spectra of Water Vapor. Part I. Energy Levels and Transition Wavenumbers for H217O and H218O. J. Quant. Spectrosc. Radiat. Transf. 2009, 110, 573–596. [Google Scholar] [CrossRef]
  36. Tennyson, J.; Bernath, P.F.; Brown, L.R.; Campargue, A.; Carleer, M.R.; Császár, A.G.; Daumont, L.; Gamache, R.R.; Hodges, J.T.; Naumenko, O.V.; et al. IUPAC critical Evaluation of the Rotational-Vibrational Spectra of Water Vapor. Part II. Energy Levels and Transition Wavenumbers for HD16O, HD17O, and HD18O. J. Quant. Spectrosc. Radiat. Transf. 2010, 111, 2160–2184. [Google Scholar] [CrossRef]
  37. Tennyson, J.; Bernath, P.F.; Brown, L.R.; Campargue, A.; Carleer, M.R.; Császár, A.G.; Daumont, L.; Gamache, R.R.; Hodges, J.T.; Naumenko, O.V.; et al. IUPAC critical evaluation of the rotational-vibrational spectra of water vapor. Part III. Energy levels and transition wavenumbers for H216O. J. Quant. Spectrosc. Radiat. Transf. 2013, 117, 29–80. [Google Scholar] [CrossRef]
  38. Tennyson, J.; Bernath, P.F.; Brown, L.R.; Campargue, A.; Császár, A.G.; Daumont, L.; Gamache, R.R.; Hodges, J.T.; Naumenko, O.V.; Polyansky, O.L.; et al. IUPAC critical evaluation of the rotational-vibrational spectra of water vapor. Part IV. Energy levels and transition wavenumbers for D216O, D217O and D218O. J. Quant. Spectrosc. Radiat. Transf. 2014, 142, 93–108. [Google Scholar] [CrossRef]
  39. Furtenbacher, T.; Szidarovszky, T.; Mátyus, E.; Fábri, C.; Császár, A.G. Analysis of the Rotational–Vibrational States of the Molecular Ion H 3 + . J. Chem. Theory Comput. 2013, 9, 5471–5478. [Google Scholar] [CrossRef] [PubMed]
  40. Furtenbacher, T.; Szidarovszky, T.; Fábri, C.; Császár, A.G. MARVEL analysis of the rotational–vibrational states of the molecular ions H2D+ and D2H+. Phys. Chem. Chem. Phys. 2013, 15, 10181–10193. [Google Scholar] [CrossRef] [PubMed]
  41. Al Derzi, A.R.; Furtenbacher, T.; Yurchenko, S.N.; Tennyson, J.; Császár, A.G. MARVEL analysis of the measured high-resolution spectra of 14NH3. J. Quant. Spectrosc. Radiat. Transf. 2015, 161, 117–130. [Google Scholar] [CrossRef]
  42. Furtenbacher, T.; Szabó, I.; Császár, A.G.; Bernath, P.F.; Yurchenko, S.N.; Tennyson, J. Experimental Energy Levels and Partition Function of the 12C2 Molecule. Astrophys. J. Suppl. 2016, 224, 44. [Google Scholar] [CrossRef]
  43. McKemmish, L.K.; Masseron, T.; Sheppard, S.; Sandeman, E.; Schofield, Z.; Furtenbacher, T.; Császár, A.G.; Tennyson, J.; Sousa-Silva, C. MARVEL analysis of the measured high-resolution spectra of 48Ti16O. Astrophys. J. Suppl. 2017, 228, 15. [Google Scholar] [CrossRef]
  44. Chubb, K.L.; Joseph, M.; Franklin, J.; Choudhury, N.; Furtenbacher, T.; Császár, A.G.; Gaspard, G.; Oguoko, P.; Kelly, A.; Yurchenko, S.N.; et al. MARVEL analysis of the measured high-resolution spectra of C2H2. J. Quant. Spectrosc. Radiat. Transf. 2018, 204, 42–55. [Google Scholar] [CrossRef]
  45. Tóbiás, R.; Furtenbacher, T.; Császár, A.G.; Naumenko, O.V.; Tennyson, J.; Flaud, J.M.; Kumard, P.; Poirier, B. Critical Evaluation of Measured Rotational-Vibrational Transitions ofFour Sulphur Isotopologues of S16O2. J. Quant. Spectrosc. Radiat. Transf. 2018, 208, 152–163. [Google Scholar] [CrossRef]
  46. Chubb, K.L.; Naumenko, O.V.; Keely, S.; Bartolotto, S.; MacDonald, S.; Mukhtar, M.; Grachov, A.; White, J.; Coleman, E.; Hu, S.M.; et al. MARVEL analysis of the measured high-resolution rovibrational spectra of H2S. J. Quant. Spectrosc. Radiat. Transf. 2018. submitted. [Google Scholar]
  47. McKemmish, L.K.; Goodhew, K.; Sheppard, S.; Bennet, A.; Martin, A.; Singh, A.; Sturgeon, C.; Godden, R.; Furtenbacher, T.; Császár, A.G.; et al. MARVEL analysis of the measured high-resolution spectra of 90Zr16O. Astrophys. J. Suppl. 2018. to be submitted. [Google Scholar]
  48. Yurchenko, S.N.; Al-Refaie, A.F.; Tennyson, J. ExoCross: A general program for generating spectra from molecular line lists. Astron. Astrophys. 2018. [Google Scholar] [CrossRef]
  49. Hill, C.; Yurchenko, S.N.; Tennyson, J. Temperature-dependent molecular absorption cross sections for exoplanets and other atmospheres. Icarus 2013, 226, 1673–1677. [Google Scholar] [CrossRef]
  50. Tennyson, J.; Yurchenko, S.N.; Al-Refaie, A.F.; Barton, E.J.; Chubb, K.L.; Coles, P.A.; Diamantopoulou, S.; Gorman, M.N.; Hill, C.; Lam, A.Z.; et al. The ExoMol database: Molecular line lists for exoplanet and other hot atmospheres. J. Mol. Spectrosc. 2016, 327, 73–94. [Google Scholar] [CrossRef]
  51. Yadin, B.; Vaness, T.; Conti, P.; Hill, C.; Yurchenko, S.N.; Tennyson, J. ExoMol Molecular linelists: I The rovibrational spectrum of BeH, MgH and CaH the X 2Σ+ state. Mon. Not. R. Astron. Soc. 2012, 425, 34–43. [Google Scholar] [CrossRef]
  52. Barton, E.J.; Yurchenko, S.N.; Tennyson, J. ExoMol Molecular linelists—II. The ro-vibrational spectrum of SiO. Mon. Not. R. Astron. Soc. 2013, 434, 1469–1475. [Google Scholar] [CrossRef]
  53. Yurchenko, S.N.; Tennyson, J. ExoMol line lists IV: The rotation–vibration spectrum of methane up to 1500 K. Mon. Not. R. Astron. Soc. 2014, 440, 1649–1661. [Google Scholar] [CrossRef]
  54. Barton, E.J.; Chiu, C.; Golpayegani, S.; Yurchenko, S.N.; Tennyson, J.; Frohman, D.J.; Bernath, P.F. ExoMol Molecular linelists—V. The ro-vibrational spectra of NaCl and KCl. Mon. Not. R. Astron. Soc. 2014, 442, 1821–1829. [Google Scholar] [CrossRef]
  55. Yorke, L.; Yurchenko, S.N.; Lodi, L.; Tennyson, J. ExoMol line lists VI: A high temperature line list for Phosphorus Nitride. Mon. Not. R. Astron. Soc. 2014, 445, 1383–1391. [Google Scholar] [CrossRef]
  56. Sousa-Silva, C.; Al-Refaie, A.F.; Tennyson, J.; Yurchenko, S.N. ExoMol line lists—VII. The rotation–vibration spectrum of phosphine up to 1500 K. Mon. Not. R. Astron. Soc. 2015, 446, 2337–2347. [Google Scholar] [CrossRef]
  57. Al-Refaie, A.F.; Yurchenko, S.N.; Yachmenev, A.; Tennyson, J. ExoMol line lists—VIII: A variationally computed line list for hot formaldehyde. Mon. Not. R. Astron. Soc. 2015, 448, 1704–1714. [Google Scholar] [CrossRef]
  58. Patrascu, A.T.; Tennyson, J.; Yurchenko, S.N. ExoMol molecular linelists: VII: The spectrum of AlO. Mon. Not. R. Astron. Soc. 2015, 449, 3613–3619. [Google Scholar] [CrossRef]
  59. Rivlin, T.; Lodi, L.; Yurchenko, S.N.; Tennyson, J.; Le Roy, R.J. ExoMol line lists X: The spectrum of sodium hydride. Mon. Not. R. Astron. Soc. 2015, 451, 5153–5157. [Google Scholar] [CrossRef]
  60. Pavlyuchko, A.I.; Yurchenko, S.N.; Tennyson, J. ExoMol line lists XI: A Hot Line List for nitric acid. Mon. Not. R. Astron. Soc. 2015, 452, 1702–1706. [Google Scholar] [CrossRef]
  61. Paulose, G.; Barton, E.J.; Yurchenko, S.N.; Tennyson, J. ExoMol Molecular linelists—XII. Line lists for eight isotopologues of CS. Mon. Not. R. Astron. Soc. 2015, 454, 1931–1939. [Google Scholar] [CrossRef]
  62. Yurchenko, S.N.; Blissett, A.; Asari, U.; Vasilios, M.; Hill, C.; Tennyson, J. ExoMol Molecular linelists—XIII. The spectrum of CaO. Mon. Not. R. Astron. Soc. 2016, 456, 4524–4532. [Google Scholar] [CrossRef]
  63. Underwood, D.S.; Tennyson, J.; Yurchenko, S.N.; Huang, X.; Schwenke, D.W.; Lee, T.J.; Clausen, S.; Fateev, A. ExoMol line lists XIV: A line list for hot SO2. Mon. Not. R. Astron. Soc. 2016, 459, 3890–3899. [Google Scholar] [CrossRef]
  64. Al-Refaie, A.F.; Polyansky, O.L.; Ovsyannikov, I.R.; Tennyson, J.; Yurchenko, S.N. ExoMol line lists XV: A hot line-list for hydrogen peroxide. Mon. Not. R. Astron. Soc. 2016, 461, 1012–1022. [Google Scholar] [CrossRef]
  65. Azzam, A.A.A.; Yurchenko, S.N.; Tennyson, J.; Naumenko, O.V. ExoMol line lists XVI: A Hot Line List for H2S. Mon. Not. R. Astron. Soc. 2016, 460, 4063–4074. [Google Scholar] [CrossRef]
  66. Underwood, D.S.; Tennyson, J.; Yurchenko, S.N.; Clausen, S.; Fateev, A. ExoMol line lists XVII: A line list for hot SO3. Mon. Not. R. Astron. Soc. 2016, 462, 4300–4313. [Google Scholar] [CrossRef]
  67. McKemmish, L.K.; Yurchenko, S.N.; Tennyson, J. ExoMol Molecular linelists—XVIII. The spectrum of Vanadium Oxide. Mon. Not. R. Astron. Soc. 2016, 463, 771–793. [Google Scholar] [CrossRef]
  68. Polyansky, O.L.; Kyuberis, A.A.; Lodi, L.; Tennyson, J.; Ovsyannikov, R.I.; Zobov, N. ExoMol molecular line lists XIX: High accuracy computed line lists for H217O and H218O. Mon. Not. R. Astron. Soc. 2017, 466, 1363–1371. [Google Scholar] [CrossRef]
  69. Mizus, I.I.; Alijah, A.; Zobov, N.F.; Kyuberis, A.A.; Yurchenko, S.N.; Tennyson, J.; Polyansky, O.L. ExoMol molecular line lists XX: A comprehensive line list for H 3 + . Mon. Not. R. Astron. Soc. 2017, 468, 1717–1725. [Google Scholar] [CrossRef]
  70. Owens, A.; Yurchenko, S.N.; Yachmenev, A.; Thiel, W.; Tennyson, J. ExoMol molecular line lists XXII. The rotation–vibration spectrum of silane up to 1200 K. Mon. Not. R. Astron. Soc. 2017, 471, 5025–5032. [Google Scholar] [CrossRef]
  71. Prajapat, L.; Jagoda, P.; Lodi, L.; Gorman, M.N.; Yurchenko, S.N.; Tennyson, J. ExoMol molecular line lists XXIII. Spectra of PO and PS. Mon. Not. R. Astron. Soc. 2017, 472, 3648–3658. [Google Scholar] [CrossRef]
  72. Yurchenko, S.N.; Sinden, F.; Lodi, L.; Hill, C.; Gorman, M.N.; Tennyson, J. ExoMol Molecular linelists—XXIV: A new hot line list for silicon monohydride, SiH. Mon. Not. R. Astron. Soc. 2018, 473, 5324–5333. [Google Scholar] [CrossRef]
  73. Upadhyay, A.; Conway, E.K.; Tennyson, J.; Yurchenko, S.N. ExoMol Molecular linelists—XXV: A hot line list for silicon sulphide, SiS. Mon. Not. R. Astron. Soc. 2018. [Google Scholar] [CrossRef]
  74. Yurchenko, S.N.; Bond, W.; Gorman, M.N.; Lodi, L.; McKemmish, L.K.; Nunn, W.; Shah, R.; Tennyson, J. ExoMol Molecular linelists—XXVI: Spectra of SH and NS. Mon. Not. R. Astron. Soc. 2018. [Google Scholar] [CrossRef]
  75. Mant, B.P.; Yachmenev, A.; Tennyson, J.; Yurchenko, S.N. ExoMol molecular line lists—XXVII: Spectra of C2H4. Mon. Not. R. Astron. Soc. 2018, in press. [Google Scholar]
  76. Yurchenko, S.N.; Williams, H.; Leyland, P.C.; Lodi, L.; Tennyson, J. ExoMol line lists XXVIII: The rovibronic spectrum of AlH. Mon. Not. R. Astron. Soc. 2018, in press. [Google Scholar] [CrossRef]
  77. Yurchenko, S.N.; Szabo, I.; Pyatenko, E.; Tennyson, J. ExoMol Molecular line lists XXXI: The spectrum of C2. Mon. Not. R. Astron. Soc. 2018, in press. [Google Scholar] [CrossRef]
  78. Li, H.Y.; Patrascu, A.; Tennyson, J.; Yurchenko, S.N. ExoMol molecular line lists XXXII: The rovibronic spectrum of MgO. Mon. Not. R. Astron. Soc. 2018. to be submitted. [Google Scholar]
  79. Yurchenko, S.N.; Barber, R.J.; Tennyson, J. A variationally computed hot line list for NH3. Mon. Not. R. Astron. Soc. 2011, 413, 1828–1834. [Google Scholar] [CrossRef]
  80. Coppola, C.M.; Lodi, L.; Tennyson, J. Radiative cooling functions for primordial molecules. Mon. Not. R. Astron. Soc. 2011, 415, 487–493. [Google Scholar] [CrossRef]
  81. Lodi, L.; Yurchenko, S.N.; Tennyson, J. The calculated rovibronic spectrum of scandium hydride, ScH. Mol. Phys. 2015, 113, 1559–1575. [Google Scholar] [CrossRef]
  82. Brooke, J.S.A.; Bernath, P.F.; Western, C.M.; van Hemert, M.C.; Groenenboom, G.C. Line strengths of rovibrational and rotational transitions within the X 3Σ ground state of NH. J. Chem. Phys. 2014, 141, 054310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Masseron, T.; Plez, B.; Van Eck, S.; Colin, R.; Daoutidis, I.; Godefroid, M.; Coheur, P.F.; Bernath, P.; Jorissen, A.; Christlieb, N. CH in stellar atmospheres: An extensive linelist. Astron. Astrophys. 2014, 571, A47. [Google Scholar] [CrossRef]
  84. Brooke, J.S.A.; Ram, R.S.; Western, C.M.; Li, G.; Schwenke, D.W.; Bernath, P.F. Einstein A Coefficients and Oscillator Strengths for the A2Π–X2Σ+ (Red) and B2Σ+X2Σ+ (Violet) Systems and Rovibrational Transitions in the X2Σ+ State of CN. Astrophys. J. Suppl. 2014, 210, 23. [Google Scholar] [CrossRef]
  85. Ram, R.S.; Brooke, J.S.A.; Western, C.M.; Bernath, P.F. Einstein A-values and oscillator strengths of the A 2Π–X 2Σ+ system of CP. J. Quant. Spectrosc. Radiat. Transf. 2014, 138, 107–115. [Google Scholar] [CrossRef]
  86. Li, G.; Gordon, I.E.; Hajigeorgiou, P.G.; Coxon, J.A.; Rothman, L.S. Reference spectroscopic data for hydrogen halides, Part II: The line lists. J. Quant. Spectrosc. Radiat. Transf. 2013, 130, 284–295. [Google Scholar] [CrossRef]
  87. Wende, S.; Reiners, A.; Seifahrt, A.; Bernath, P.F. CRIRES spectroscopy and empirical line-by-line identification of FeH molecular absorption in an M dwarf. Astron. Astrophys. 2010, 523, A58. [Google Scholar] [CrossRef]
  88. Burrows, A.; Dulick, M.; Bauschlicher, C.W.; Bernath, P.F.; Ram, R.S.; Sharp, C.M.; Milsom, J.A. Spectroscopic constants, abundances, and opacities of the TiH molecule. Astrophys. J. 2005, 624, 988–1002. [Google Scholar] [CrossRef]
  89. Schwenke, D.W. Opacity of TiO from a coupled electronic state calculation parametrized by ab initio and experimental data. Faraday Discuss. 1998, 109, 321–334. [Google Scholar] [CrossRef]
  90. Lyulin, O.M.; Perevalov, V.I. ASD-1000: High-resolution, high-temperature acetylene spectroscopic databank. J. Quant. Spectrosc. Radiat. Transf. 2017, 201, 94–103. [Google Scholar] [CrossRef]
  91. Burrows, A.; Ram, R.S.; Bernath, P.; Sharp, C.M.; Milsom, J.A. New CrH opacities for the study of L and brown dwarf atmospheres. Astrophys. J. 2002, 577, 986–992. [Google Scholar] [CrossRef]
  92. Owens, A.; Yachmenev, A.; Küpper, J.; Yurchenko, S.N.; Thiele, W. The rotation–vibration spectrum of methyl fluoride from first principles. Phys. Chem. Chem. Phys. 2018, in press. [Google Scholar]
  93. Voronin, B.A.; Tennyson, J.; Tolchenov, R.N.; Lugovskoy, A.A.; Yurchenko, S.N. A high accuracy computed line list for the HDO molecule. Mon. Not. R. Astron. Soc. 2010, 402, 492–496. [Google Scholar] [CrossRef]
  94. Allard, F.; Hauschildt, P.H.; Miller, S.; Tennyson, J. The influence of H2O line blanketing on the spectra of cool dwarf stars. Astrophys. J. 1994, 426, L39–L41. [Google Scholar] [CrossRef]
  95. Tinetti, G.; Tennyson, J.; Griffiths, C.A.; Waldmann, I. Water in Exoplanets. Philos. Trans. R. Soc. Lond. A 2012, 370, 2749–2764. [Google Scholar] [CrossRef] [PubMed]
  96. Barber, R.J.; Tennyson, J.; Harris, G.J.; Tolchenov, R.N. A high accuracy computed water line list. Mon. Not. R. Astron. Soc. 2006, 368, 1087–1094. [Google Scholar] [CrossRef]
  97. Allard, F. The BT-Settl Model Atmospheres for Stars, Brown Dwarfs and Planets. In IAU Symposium; Booth, M., Matthews, B.C., Graham, J.R., Eds.; International Astronomical Union: Pairs, France, 2014; Volume 299, pp. 271–272. [Google Scholar]
  98. Rutkowski, L.; Foltynowicz, A.; Johansson, A.C.; Khodabakhsh, A.; Schmidt, F.M.; Kyuberis, A.A.; Zobov, N.F.; Polyansky, O.L.; Yurchenko, S.N.; Tennyson, J. An experimental water line list at 1950 K in the 6250–6670 cm−1 region. J. Quant. Spectrosc. Radiat. Transf. 2018, 205, 213–219. [Google Scholar] [CrossRef]
  99. Campargue, A.; Mikhailenko, S.N.; Vasilchenko, S.; Reynaud, C.; Beguier, S.; Cermak, P.; Mondelain, D.; Kassi, S.; Romanini, D. The absorption spectrum of water vapor in the 2.2 μm transparency window: High sensitivity measurements and spectroscopic database. J. Quant. Spectrosc. Radiat. Transf. 2017, 189, 407–416. [Google Scholar] [CrossRef]
  100. Rey, M.; Nikitin, A.V.; Tyuterev, V.G. Theoreticak hot methane line list up T=2000 K for astrophysical applications. Astrophys. J. 2014, 789, 2. [Google Scholar] [CrossRef]
  101. Rey, M.; Nikitin, A.V.; Tyuterev, V.G. Accurate Theoretical Methane Line Lists in the Infrared up to 3000 K and Quasi-continuum Absorption/Emission Modeling for Astrophysical Applications. Astrophys. J. 2017, 847, 105. [Google Scholar] [CrossRef]
  102. Hargreaves, R.J.; Beale, C.A.; Michaux, L.; Irfan, M.; Bernath, P.F. Hot methane line lists for exoplanet and brown dwarf atmospheres. Astrophys. J. 2012, 757, 46. [Google Scholar] [CrossRef]
  103. Hargreaves, R.J.; Bernath, P.F.; Bailey, J.; Dulick, M. Empirical Line Lists and Absorption Cross Sections for Methane at High Temperatures. Astrophys. J. 2015, 813, 12. [Google Scholar] [CrossRef]
  104. Canty, J.I.; Lucas, P.W.; Tennyson, J.; Yurchenko, S.N.; Leggett, S.K.; Tinney, C.G.; Jones, H.R.A.; Burningham, B.; Pinfield, D.J.; Smart, R.L. Methane and Ammonia in the near-infrared spectra of late T dwarfs. Mon. Not. R. Astron. Soc. 2015, 450, 454–480. [Google Scholar] [CrossRef]
  105. Yurchenko, S.N.; Amundsen, D.S.; Tennyson, J.; Waldmann, I.P. A hybrid line list for CH4 and hot methane continuum. Astron. Astrophys. 2017, 605, A95. [Google Scholar] [CrossRef]
  106. Fletcher, L.N.; Orton, G.S.; Teanby, N.A.; Irwin, P.G.J. Phosphine on Jupiter and Saturn from Cassini/CIRS. Icarus 2009, 202, 543–564. [Google Scholar] [CrossRef]
  107. Yurchenko, S.N. A theoretical room-temperature line list for 15NH3. J. Quant. Spectrosc. Radiat. Transf. 2015, 152, 28–36. [Google Scholar] [CrossRef]
  108. Rey, M.; Delahaye, T.; Nikitin, A.V.; Tyuterev, V.G. First theoretical global line lists of ethylene (12 C2H4 ) spectra for the temperature range 50,700 K in the far-infrared for quantification of absorption and emission in planetary atmospheres. Astron. Astrophys. 2016, 594, A47. [Google Scholar] [CrossRef]
  109. Al-Refaie, A.F.; Ovsyannikov, R.I.; Polyansky, O.L.; Yurchenko, S.N.; Tennyson, J. A variationally calculated room temperature line-list for H2O2. J. Mol. Spectrosc. 2015, 318, 84–90. [Google Scholar] [CrossRef]
  110. Harris, G.J.; Polyansky, O.L.; Tennyson, J. Opacity data for HCN and HNC from a new ab initio linelist. Astrophys. J. 2002, 578, 657–663. [Google Scholar] [CrossRef]
  111. Harris, G.J.; Tennyson, J.; Kaminsky, B.M.; Pavlenko, Y.V.; Jones, H.R.A. Improved HCN/HNC linelist, model atmospheres synthetic spectra for WZ Cas. Mon. Not. R. Astron. Soc. 2006, 367, 400–406. [Google Scholar] [CrossRef]
  112. Mellau, G.C. Highly excited rovibrational states of HNC. J. Mol. Spectrosc. 2011, 269, 77–85. [Google Scholar] [CrossRef]
  113. Mellau, G.C. Complete experimental rovibrational eigenenergies of HCN up to 6880 cm−1 above the ground state. J. Chem. Phys. 2011, 134, 234303. [Google Scholar] [CrossRef] [PubMed]
  114. Mellau, G.C. Rovibrational eigenenergy structure of the [H,C,N] molecular system. J. Chem. Phys. 2011, 134, 194302. [Google Scholar] [CrossRef] [PubMed]
  115. Harris, G.J.; Larner, F.C.; Tennyson, J.; Kaminsky, B.M.; Pavlenko, Y.V.; Jones, H.R.A. A H13CN/HN13C linelist, model atmospheres and synthetic spectra for carbon stars. Mon. Not. R. Astron. Soc. 2008, 390, 143–148. [Google Scholar] [CrossRef] [Green Version]
  116. Tsiaras, A.; Rocchetto, M.; Waldmann, I.P.; Tinetti, G.; Varley, R.; Morello, G.; Barton, E.J.; Yurchenko, S.N.; Tennyson, J. Detection of an atmosphere around the super-Earth 55 Cancri e. Astrophys. J. 2016, 820, 99. [Google Scholar] [CrossRef]
  117. Eriksson, K.; Gustafsson, B.; Jørgensen, U.G.; Nordlund, A. Effects of HCN molecules in carbon star atmospheres. Astron. Astrophys. 1984, 132, 37–44. [Google Scholar]
  118. Barber, R.J.; Harris, G.J.; Tennyson, J. Temperature dependent partition functions and equilibrium constant for HCN and HNC. J. Chem. Phys. 2002, 117, 11239–11243. [Google Scholar] [CrossRef]
  119. Kaltenegger, L.; Henning, W.G.; Sasselov, D.D. Detecting volcanism on extrasolar planets. Astron. J. 2010, 140, 1370. [Google Scholar] [CrossRef]
  120. Hu, R.; Seager, S.; Bains, W. Photochemistry in terrestrial exoplanet atmospheres. II. H2S and SO2 photochemistry in anoxic atmospheres. Astrophys. J. 2013, 769, 6. [Google Scholar] [CrossRef]
  121. Whitehill, A.R.; Xie, C.; Hu, X.; Xie, D.; Guo, H.; Ono, S. Vibronic origin of sulfur mass-independent isotope effect in photoexcitation of SO2 and the implications to the early earth’s atmosphere. Proc. Natl. Acad. Sci. USA 2013, 110, 17697–17702. [Google Scholar] [CrossRef] [PubMed]
  122. Johnson, S.S.; Pavlov, A.A.; Mischna, M.A. Fate of SO2 in the ancient Martian atmosphere: Implications for transient greenhouse warming. J. Geophys. Res. 2009, 114, E11011. [Google Scholar] [CrossRef]
  123. Huang, X.; Gamache, R.R.; Freedman, R.S.; Schwenke, D.W.; Lee, T.J. Reliable infrared line lists for 13 CO2 isotopologues up to E =18,000 cm−1 and 1500 K, with line shape parameters. J. Quant. Spectrosc. Radiat. Transf. 2014, 147, 134–144. [Google Scholar]
  124. Błęcka, M.I.; De Mazière, M. Detection of nitric acid and nitric oxides in the terrestrial atmosphere in the middle-infrared spectral region. Ann. Geophys. 1997, 14, 1103–1110. [Google Scholar] [CrossRef]
  125. Cooper, M.; Martin, R.V.; Sauvage, B.; Boone, C.D.; Walker, K.A.; Bernath, P.F.; McLinden, C.A.; Degenstein, D.A.; Volz-Thomas, A.; Wespes, C. Evaluation of ACE-FTS and OSIRIS Satellite retrievals of ozone and nitric acid in the tropical upper troposphere: Application to ozone production efficiency. J. Geophys. Res. 2011, 116, D12306. [Google Scholar] [CrossRef]
  126. Seager, S.; Bains, W.; Hu, R. A Biomass-based Model to Estimate the Plausibility of Exoplanet Biosignature Gases. Astrophys. J. 2013, 775, 104. [Google Scholar] [CrossRef]
  127. Brown, L.R.; Crisp, J.A.; Crisp, D.; Naumenko, O.V.; Smirnov, M.A.; Sinitsa, L.N.; Perrin, A. The Absorption Spectrum of H2S Between 2150 and 4260 cm−1: Analysis of the Positions and Intensities in the First (2ν2, ν1, and ν3) and Second (3ν2, ν1/ν2, and ν2/ν3) Triad Regions. J. Mol. Spectrosc. 1998, 188, 148–174. [Google Scholar] [CrossRef] [PubMed]
  128. Somerville, W.B. Interstellar radio spectrum lines. Rep. Prog. Phys. 1977, 40, 483–565. [Google Scholar]
  129. Lovas, F.J.; Johnson, D.R.; Snyder, L.E. Recommended rest frequencies for observed interstellar molecular transitions. Astrophys. J. Suppl. 1979, 41, 451–480. [Google Scholar] [CrossRef]
  130. Tomkin, J.; Lambert, D.L. Isotopic adbundances of magnesium 5 G-dwarfs and K-dwarfs. Astrophys. J. 1980, 235, 925–938. [Google Scholar] [CrossRef]
  131. McWilliam, A.; Lambert, D.L. Isotopic magnesium abundances in stars. Mon. Not. R. Astron. Soc. 1988, 230, 573–585. [Google Scholar] [CrossRef]
  132. Gay, P.L.; Lambert, D.L. The isotopic abundances of magnesium in stars. Astrophys. J. 2000, 533, 260–270. [Google Scholar] [CrossRef]
  133. Yong, D.; Lambert, D.L.; Ivans, I.I. Magnesium isotopic abundance ratios in cool stars. Astrophys. J. 2003, 599, 1357–1371. [Google Scholar] [CrossRef]
  134. Hinkle, K.H.; Wallace, L.; Ram, R.S.; Bernath, P.F.; Sneden, C.; Lucatello, S. The magnesium isotopologues of MgH in the A 2Π – X 2Σ+ system. Astrophys. J. Suppl. 2013, 207, 26. [Google Scholar] [CrossRef]
  135. GharibNezhad, E.; Shayesteh, A.; Bernath, P.F. Einstein A coefficients for rovibronic lines of the A 2Π→ X 2Σ+ and B’ 2Σ+→ X 2Σ+ transitions of MgH. Mon. Not. R. Astron. Soc. 2013, 432, 2043–2047. [Google Scholar] [CrossRef]
  136. Henderson, R.D.E.; Shayesteh, A.; Tao, J.; Haugen, C.C.; Bernath, P.F.; Le Roy, R.J. Accurate Analytic Potential and Born-Oppenheimer Breakdown Functions for MgH and MgD from a Direct-Potential-Fit Data Analysis. J. Phys. Chem. A 2013, 117, 13373–13387. [Google Scholar] [CrossRef] [PubMed]
  137. Reiners, A.; Homeier, D.; Hauschildt, P.H.; Allard, F. A high resolution spectral atlas of brown dwarfs. Astron. Astrophys. 2007, 473, 245–255. [Google Scholar] [CrossRef]
  138. Darby-Lewis, D.; Tennyson, J.; Lawson, K.D.; Yurchenko, S.N.; Stamp, M.F.; Shaw, A.; Brezinsek, S.; JET Contributor. Synthetic spectra of BeH, BeD and BeT for emission modelling in JET plasmas. J. Phys. B At. Mol. Opt. Phys. 2018. submitted. [Google Scholar]
  139. Shanmugavel, R.; Bagare, S.P.; Rajamanickam, N.; Balachandra Kumar, K. Identification of Beryllium Hydride Isotopomer Lines in Sunspot Umbral Spectra. Serb. Astron. J. 2008, 176, 51–58. [Google Scholar] [CrossRef]
  140. Duxbury, G.; Stamp, M.F.; Summers, H.P. Observations and modelling of diatomic molecular spectra from JET. Plasma Phys. Control. Fusion 1998, 40, 361–370. [Google Scholar] [CrossRef]
  141. Wallace, L.; Hinkle, K.; Livingston, W. An Atlas of Sunspot Umbral Spectra in the Visible from 15,000 to 25,000 cm−1 (3920 to 6664 Å); Technical Report Tech. Rep. 00-001; National Solar Observatory: Tucson, AZ, USA, 2000. [Google Scholar]
  142. Kaminski, T.; Wong, K.T.; Schmidt, M.R.; Mueller, H.S.P.; Gottlieb, C.A.; Cherchneff, I.; Menten, K.M.; Keller, D.; Bruenken, S.; Winters, J.M.; et al. An observational study of dust nucleation in Mira (o Ceti) I. Variable features of AlO and other Al- bearing species. Astron. Astrophys. 2016, 592, A42. [Google Scholar] [CrossRef]
  143. Li, G.; Gordon, I.E.; Bernath, P.F.; Rothman, L.S. Direct fit of experimental ro-vibrational intensities to the dipole moment function: Application to {HCl}. J. Quant. Spectrosc. Radiat. Transf. 2011, 112, 1543–1550. [Google Scholar] [CrossRef]
  144. Tenenbaum, E.D.; Woolf, N.J.; Ziurys, L.M. Identification of phosphorus monoxide (X 2Π) in VY Canis Majoris: Detection of the first P-O bond in space. Astrophys. J. 2007, 666, L29–L32. [Google Scholar] [CrossRef]
  145. De Beck, E.; Kaminski, T.; Patel, N.A.; Young, K.H.; Gottlieb, C.A.; Menten, K.M.; Decin, L. PO and PN in the wind of the oxygen-rich AGB star IK Tauri. Astron. Astrophys. 2013, 558, 9. [Google Scholar] [CrossRef]
  146. Rivilla, V.M.; Fontani, F.; Beltrán, M.T.; Vasyunin, A.; Caselli, P.; Martín-Pintado, J.; Cesaroni, R. The First Detections of the Key Prebiotic Molecule PO in Star-forming Regions. Astrophys. J. 2016, 826, 161. [Google Scholar] [CrossRef]
  147. Lefloch, B.; Vastel, C.; Viti, S.; Jimenez-Serra, I.; Codella, C.; Podio, L.; Ceccarelli, C.; Mendoza, E.; Lepine, J.R.D.; Bachiller, R. Phosphorus-bearing molecules in solar-type star-forming regions: First PO detection. Mon. Not. R. Astron. Soc. 2016, 462, 3937. [Google Scholar] [CrossRef]
  148. Jones, H.R.A.; Pavlenko, Y.; Viti, S.; Barber, R.J.; Yakovina, L.; Pinfold, D.; Tennyson, J. Carbon monoxide in low-mass dwarf stars. Mon. Not. R. Astron. Soc. 2005, 358, 105–112. [Google Scholar] [CrossRef]
  149. De Kok, R.J.; Brogi, M.; Snellen, I.A.G.; Birkby, J.; Albrecht, S.; de Mooij, E.J.W. Detection of carbon monoxide in the high-resolution day-side spectrum of the exoplanet HD 189,733b. Astron. Astrophys. 2013, 554, A82. [Google Scholar] [CrossRef]
  150. Brogi, M.; de Kok, R.J.; Birkby, J.L.; Schwarz, H.; Snellen, I.A.G. Carbon monoxide and water vapor in the atmosphere of the non-transiting exoplanet HD 179,949 b. Astron. Astrophys. 2014, 565, A124. [Google Scholar] [CrossRef]
  151. Tennyson, J. Vibration-rotation transition intensities from first principles. J. Mol. Spectrosc. 2014, 298, 1–6. [Google Scholar] [CrossRef]
  152. Danielak, J.; Domin, U.; Kepa, R.; Rytel, M.; Zachwieja, M. Reinvestigation of the emission gamma band system (A 2Σ+ – X 2Π) of the NO molecule. J. Mol. Spectrosc. 1997, 181, 394–402. [Google Scholar] [CrossRef]
  153. Pavlenko, Y.V. A “lithium test” and modeling of lithium lines in late-type M dwarfs: Teide1. Astron. Rep. 1997, 41, 537–542. [Google Scholar]
  154. Bittner, D.M.; Bernath, P.F. Line Lists for LiF and LiCl in the X 1Σ+ Ground State. Astrophys. J. Suppl. 2018, 235, 8. [Google Scholar] [CrossRef]
  155. McKemmish, L.K.; Yurchenko, S.N.; Tennyson, J. Ab initio calculations to support accurate modelling of the rovibronic spectroscopy calculations of vanadium monoxide (VO). Mol. Phys. 2016, 114, 3232–3248. [Google Scholar] [CrossRef]
  156. Tennyson, J.; Lodi, L.; McKemmish, L.K.; Yurchenko, S.N. The ab initio calculation of spectra of open shell diatomic molecules. J. Phys. B At. Mol. Opt. Phys. 2016, 49, 102001. [Google Scholar] [CrossRef]
  157. Cushing, M.C.; Rayner, J.T.; Davis, S.P.; Vacca, W.D. FeH absorption in the near-infrared spectra of late M and L dwarfs. Astrophys. J. 2003, 582, 1066–1072. [Google Scholar] [CrossRef]
  158. Hargreaves, R.J.; Hinkle, K.H.; Bauschlicher, C.W., Jr.; Wende, S.; Seifahrt, A.; Bernath, P.F. High-resolution 1.6 μm spectra of FeH in M and L dwarfs. Astron. J. 2010, 140, 919–924. [Google Scholar] [CrossRef]
  159. Wallace, L.; Hinkle, K. Detection of the 1.6 μm E 4Π–A 4Π FeH system in sunspot and cool star spectra. Astrophys. J. 2001, 559, 424–427. [Google Scholar] [CrossRef]
  160. Ramos, A.A.; Bueno, J.T.; Collados, M. Detection of polarization from the E 4Π–A 4Π system of FeH in sunspot spectra. Astrophys. J. 2004, 603, L125–L128. [Google Scholar] [CrossRef]
  161. Harrison, J.J.; Brown, J.M. Measurement of the magnetic properies of FeH in its X 4Δ and F 4Δ states from sunspot spectra. Astrophys. J. 2008, 686, 1426–1431. [Google Scholar] [CrossRef]
  162. DeYonker, N.J.; Allen, W.D. Taming the low-lying electronic states of FeH. J. Chem. Phys. 2012, 137, 234303. [Google Scholar] [CrossRef] [PubMed]
  163. Dulick, M.; Bauschlicher, C.W.; Burrows, A.; Sharp, C.M.; Ram, R.S.; Bernath, P. Line intensities and molecular opacities of the FeH F 4Δ–X 4Δ transition. Astrophys. J. 2003, 594, 651–663. [Google Scholar] [CrossRef]
  164. Rajpurohit, A.S.; Reyle, C.; Allard, F.; Homeier, D.; Schultheis, M.; Bessell, M.S.; Robin, A.C. The effective temperature scale of M dwarfs. Astron. Astrophys. 2013, 556, A15. [Google Scholar] [CrossRef]
  165. Ardila, D.R.; Van Dyk, S.D.; Makowiecki, W.; Stauffer, J.; Song, I.; Rho, J.; Fajardo-Acosta, S.; Hoard, D.W.; Wachter, S. The Spitzer atlas of stellar spectra (SASS). Astrophys. J. Suppl. 2010, 191, 301–339. [Google Scholar] [CrossRef]
  166. Campbell, J.M.; Klapstein, D.; Dulick, M.; Bernath, P.F. Infrared-absorption and emission-spectra of SiO. Astrophys. J. Suppl. 1995, 101, 237–254. [Google Scholar] [CrossRef]
  167. Wallace, L.; Hinkle, K. Medium-resolution stellar spectra in the L band from 2400 to 3000 cm−1 (3.3 to 4.2 microns). Astron. J. 2002, 124, 3393–3399. [Google Scholar] [CrossRef]
  168. Joshi, G.C.; Punetha, L.M.; Pande, M.C. (A–X) system of SiO in sunspots. Sol. Phys. 1979, 62, 77–82. [Google Scholar] [CrossRef]
  169. Bauschlicher, C.W., Jr. The low-lying electronic states of SiO. Chem. Phys. Lett. 2016, 658, 76–79. [Google Scholar] [CrossRef]
  170. Kurucz, R.L. Including all the lines. Can. J. Phys. 2011, 89, 417–428. [Google Scholar] [CrossRef]
  171. Bai, X.; Motto-Ros, V.; Lei, W.; Zheng, L.; Yu, J. Experimental determination of the temperature range of AlO molecular emission in laser-induced aluminum plasma in air. Spectra Chim. Acta B 2014, 99, 193–200. [Google Scholar] [CrossRef]
  172. Surmick, D.M.; Parigger, C.G. Aluminum Monoxide Emission Measurements in a Laser-Induced Plasma. Appl. Spectrosc. 2014, 68, 992–996. [Google Scholar] [CrossRef] [PubMed]
  173. Parigger, C.G.; Hornkohl, J.O. Computation of AlO B 2Σ+→ X 2Σ+ emission spectra. Spectra Chim. Acta A 2011, 81, 404–411. [Google Scholar] [CrossRef] [PubMed]
  174. Kirkpatrick, J.D.; Kelly, D.M.; Rieke, G.H.; Liebert, J.; Allard, F.; Wehrse, R. M dwarf spectra from 0.6 to 1.5 micron-A spectral sequence, model atmosphere fitting, and the temperature scale. Astrophys. J. 1993, 402, 643–654. [Google Scholar] [CrossRef]
  175. McGovern, M.R.; Kirkpatrick, J.D.; McLean, I.S.; Burgasser, A.J.; Prato, L.; Lowrance, P.J. Identifying Young Brown Dwarfs Using Gravity-Sensitive Spectral Features. Astrophys. J. 2004, 600, 1020. [Google Scholar] [CrossRef]
  176. Kirkpatrick, J.D.; Barman, T.S.; Burgasser, A.J.; McGovern, M.R.; McLean, I.S.; Tinney, C.G.; Lowrance, P.J. Discovery of a Very Young Field L Dwarf, 2MASS J01415823–4633574. Astrophys. J. 2006, 639, 1120. [Google Scholar] [CrossRef]
  177. Peterson, D.E.; Megeath, S.T.; Luhman, K.L.; Pipher, J.L.; Stauffer, J.R.; Barrado y Navascués, D.; Wilson, J.C.; Skrutskie, M.F.; Nelson, M.J.; Smith, J.D. New Young Brown Dwarfs in the Orion Molecular Cloud 2/3 Region. Astrophys. J. 2008, 685, 313. [Google Scholar] [CrossRef]
  178. Desert, J.M.; Vidal-Madjar, A.; des Etangs, A.L.; Sing, D.; Ehrenreich, D.; Hebrard, G.; Ferlet, R. TiO and VO broad band absorption features in the optical spectrum of the atmosphere of the hot-Jupiter HD 209458b. Astron. Astrophys. 2008, 492, 585–592. [Google Scholar] [CrossRef] [Green Version]
  179. Allard, F.; Hauschildt, P.H.; Schwenke, D. TiO and H2O absorption lines in cool stellar atmospheres. Astrophys. J. 2000, 540, 1005–1015. [Google Scholar] [CrossRef]
  180. Sedaghati, E.; Boffin, H.M.J.; MacDonald, R.J.; Gandhi, S.; Madhusudhan, N.; Gibson, N.P.; Oshagh, M.; Claret, A.; Rauer, H. Detection of titanium oxide in the atmosphere of a hot Jupiter. Nature 2017, 549, 238. [Google Scholar] [CrossRef] [PubMed]
  181. Nugroho, S.K.; Kawahara, H.; Masuda, K.; Hirano, T.; Kotani, T.; Tajitsu, A. High-resolution Spectroscopic Detection of TiO and a Stratosphere in the Day-side of WASP-33b. Astrophys. J. 2017, 154, 221. [Google Scholar] [CrossRef]
  182. Tsiaras, A.; Waldmann, I.P.; Zingales, T.; Rocchetto, M.; Morello, G.; Damiano, M.; Karpouzas, K.; Tinetti, G.; McKemmish, L.K.; Tennyson, J.; et al. A Population Study of Gaseous Exoplanets. Astron. J. 2018, 155, 156. [Google Scholar] [CrossRef]
  183. Ryabchikova, T.; Piskunov, N.; Kurucz, R.L.; Stempels, H.C.; Heiter, U.; Pakhomov, Y.; Barklem, P.S. A major upgrade of the VALD database. Phys. Scr. 2015, 90, 054005. [Google Scholar] [CrossRef]
  184. McKemmish, L.K.; Masseron, T.; Yurchenko, S.N.; Tennyson, J. ExoMol Molecular linelists—XXXIV. The spectrum of Titanium Oxide. Mon. Not. R. Astron. Soc. 2018. to be submitted. [Google Scholar]
  185. Wollaston, W.H. A Method of Examining Refractive and Dispersive Powers, by Prismatic Reflection. Philos. Trans. R. Soc. Lond. 1802, 92, 365–380. [Google Scholar] [CrossRef]
  186. Swan, W. On the Prismatic Spectra of the Flames of Compounds of Carbon and Hydrogen. Trans. R. Soc. Edinb. 1857, 21, 411. [Google Scholar] [CrossRef]
  187. Brooke, J.S.; Bernath, P.F.; Schmidt, T.W.; Bacskay, G.B. Line strengths and updated molecular constants for the C2 Swan system. J. Quant. Spectrosc. Radiat. Transf. 2013, 124, 11–20. [Google Scholar] [CrossRef]
  188. Pavanello, M.; Adamowicz, L.; Alijah, A.; Zobov, N.F.; Mizus, I.I.; Polyansky, O.L.; Tennyson, J.; Szidarovszky, T.; Császár, A.G.; Berg, M.; et al. Precision measurements and computations of transition energies in rotationally cold triatomic hydrogen ions up to the mid-visible spectral range. Phys. Rev. Lett. 2012, 108, 023002. [Google Scholar] [CrossRef] [PubMed]
  189. Miller, S.; Tennyson, J. Calculated rotational and ro-vibrational transitions in the spectrum of H 3 + . Astrophys. J. 1988, 335, 486–490. [Google Scholar] [CrossRef]
  190. Miller, S.; Achilleos, N.; Ballester, G.E.; Geballe, T.R.; Joseph, R.D.; Prange, R.; Rego, D.; Stallard, T.; Tennyson, J.; Trafton, L.M.; et al. The role of H 3 + in planetary atmospheres. Philos. Trans. R. Soc. Lond. A 2000, 358, 2485–2502. [Google Scholar] [CrossRef]
  191. Shkolnik, E.; Gaidos, E.; Moskovitz, N. No Detectable H 3 + Emission from the Atmospheres of Hot Jupiters. Astrophys. J. 2006, 132, 1267. [Google Scholar]
  192. Koskinen, T.T.; Aylward, A.D.; Miller, S. A stability limit for the atmospheres of giant extrasolar planets. Nature 2007, 450, 845–848. [Google Scholar] [CrossRef] [PubMed]
  193. Miller, S.; Stallard, T.; Melin, H.; Tennyson, J. H 3 + cooling in planetary atmospheres. Faraday Discuss. 2010, 147, 283–291. [Google Scholar] [CrossRef] [PubMed]
  194. Miller, S.; Stallard, T.; Tennyson, J.; Melin, H. Cooling by H 3 + emission. J. Phys. Chem. A 2013, 117, 9770–9777. [Google Scholar] [CrossRef] [PubMed]
  195. Bergeron, P.; Ruiz, M.T.; Leggett, S.K. The chemical evolution of cool white dwarfs and the age of the local galactic disk. Astrophys. J. Suppl. 1997, 108, 339–387. [Google Scholar] [CrossRef]
  196. Neale, L.; Tennyson, J. A high temperature partition function for H 3 + . Astrophys. J. 1995, 454, L169–L173. [Google Scholar] [CrossRef]
  197. Sochi, T.; Tennyson, J. A computed line list for the H2D+ molecular ion. Mon. Not. R. Astron. Soc. 2010, 405, 2345–2350. [Google Scholar] [CrossRef]
  198. Herman, M.; Campargue, A.; El Idrissi, M.I.; Vander Auwera, J. Vibrational spectroscopic database on acetylene, X1Σg+ (12C2H2, 12C2D2 and 13C2H2). J. Phys. Chem. Ref. Data 2003, 32, 921–1361. [Google Scholar] [CrossRef]
  199. Herman, M. The acetylene ground state saga. Mol. Phys. 2007, 105, 2217–2241. [Google Scholar] [CrossRef]
  200. Jørgensen, U.G.; Almlöf, J.; Siegbahn, P.E.M. Complete active space self-consistent field calculations of the vibrational band strengths for C3. Astrophys. J. 1989, 343, 554. [Google Scholar] [CrossRef]
  201. Gorman, M.; Yurchenko, S.N.; Tennyson, J. ExoMol Molecular linelists—XXXIII. The spectrum of Chromium Hydride. Mon. Not. R. Astron. Soc. 2018. to be submitted. [Google Scholar]
  202. Gorman, M.; Yurchenko, S.N.; Tennyson, J. ExoMol Molecular linelists—XXXV. The spectrum of manganese hydride. Mon. Not. R. Astron. Soc. 2018. to be submitted. [Google Scholar]
  203. Faure, A.; Wiesenfeld, L.; Tennyson, J.; Drouin, B.J. Pressure broadening of water and carbon monoxide transitions by molecular hydrogen at high temperatures. J. Quant. Spectrosc. Radiat. Transf. 2013, 116, 79–86. [Google Scholar] [CrossRef]
  204. Renaud, C.L.; Cleghorn, K.; Hartmann, L.; Vispoel, B.; Gamache, R.R. Line shape parameters for the H2O–H2 collision system for application to exoplanet and planetary atmospheres. Icarus 2018, 306, 275–284. [Google Scholar] [CrossRef]
  205. Barton, E.J.; Hill, C.; Czurylo, M.; Li, H.Y.; Hyslop, A.; Yurchenko, S.N.; Tennyson, J. The ExoMol diet of line-by-line pressure-broadening parameters. J. Quant. Spectrosc. Radiat. Transf. 2017, 203, 490–495. [Google Scholar] [CrossRef]
Figure 1. Cross sections for H 2 16 O from the POKAZATEL line list [5], H 2 17 O from the HotWat78 line list [68] and HDO from the VTT line list [93]. All cross sections are for 100% abundance.
Figure 1. Cross sections for H 2 16 O from the POKAZATEL line list [5], H 2 17 O from the HotWat78 line list [68] and HDO from the VTT line list [93]. All cross sections are for 100% abundance.
Atoms 06 00026 g001
Figure 2. Cross sections generated using ExoMol line lists for methane, 10to10 line list [53], silane [72], phosphine [56], ammonia [79] and ethylene [75].
Figure 2. Cross sections generated using ExoMol line lists for methane, 10to10 line list [53], silane [72], phosphine [56], ammonia [79] and ethylene [75].
Atoms 06 00026 g002
Figure 3. Cross sections for polyatomic oxides and HCN. Line lists are from ExoMol for hydrogen peroxide [109], hydrogen cyanide [30], sulfur dioxide [63], nitric aid [60] and sulfur trioxide [66]. The carbon dioxide data is taken from Ames-2016 [7].
Figure 3. Cross sections for polyatomic oxides and HCN. Line lists are from ExoMol for hydrogen peroxide [109], hydrogen cyanide [30], sulfur dioxide [63], nitric aid [60] and sulfur trioxide [66]. The carbon dioxide data is taken from Ames-2016 [7].
Atoms 06 00026 g003
Figure 4. Cross sections obtained from ExoMol line lists for HNO 3 [60], CH 3 Cl [75], and C 2 H 2 [90].
Figure 4. Cross sections obtained from ExoMol line lists for HNO 3 [60], CH 3 Cl [75], and C 2 H 2 [90].
Atoms 06 00026 g004
Figure 5. Cross sections for hydrogen sulfide generated using the ExoMol line list for H 2 S [65].
Figure 5. Cross sections for hydrogen sulfide generated using the ExoMol line list for H 2 S [65].
Atoms 06 00026 g005
Figure 6. Cross sections for alkaline earth monohydrides MgH and CaH from the Yadin ExoMol line lists [51] and NS from the SNaSH line list [74].
Figure 6. Cross sections for alkaline earth monohydrides MgH and CaH from the Yadin ExoMol line lists [51] and NS from the SNaSH line list [74].
Atoms 06 00026 g006
Figure 7. Cross sections for alkaline earth monohydrides and CH. BeH uses the updated ExoMol line list of Darby-Lewis et al. [138], AlH is the new ExoMol line list [76] and CH is the empirical work of Masseron et al. [83]; the CH line list is only defined for T > 1000 K.
Figure 7. Cross sections for alkaline earth monohydrides and CH. BeH uses the updated ExoMol line list of Darby-Lewis et al. [138], AlH is the new ExoMol line list [76] and CH is the empirical work of Masseron et al. [83]; the CH line list is only defined for T > 1000 K.
Atoms 06 00026 g007
Figure 8. Cross sections for monohydrides: an empirical list due to Li. et al. [86] for HCl, an ExoMol line list for mercapto radical SH [74], chromium hydride [91]. The OH data are taken from HITEMP [4].
Figure 8. Cross sections for monohydrides: an empirical list due to Li. et al. [86] for HCl, an ExoMol line list for mercapto radical SH [74], chromium hydride [91]. The OH data are taken from HITEMP [4].
Atoms 06 00026 g008
Figure 9. Cross sections generated from ExoMol line lists for sodium chloride [54], potassium chloride [54], phosphorous monoxide [71], carbon monosulfide [61] and phosphorous nitride [55].
Figure 9. Cross sections generated from ExoMol line lists for sodium chloride [54], potassium chloride [54], phosphorous monoxide [71], carbon monosulfide [61] and phosphorous nitride [55].
Atoms 06 00026 g009
Figure 10. Cross sections for carbon monoxide, cyanide, carbon phosphide and calcium oxide. The CN [84] and CP [85] cross sections are based on empirical line lists from the Bernath group. The CaO data are taken from an ExoMol line list [62]. The CO [8] line list is based on an empirical dipole moment function.
Figure 10. Cross sections for carbon monoxide, cyanide, carbon phosphide and calcium oxide. The CN [84] and CP [85] cross sections are based on empirical line lists from the Bernath group. The CaO data are taken from an ExoMol line list [62]. The CO [8] line list is based on an empirical dipole moment function.
Atoms 06 00026 g010
Figure 11. Cross sections generated from ExoMol line lists for nitric oxide [9] and phosphorous monosulfide [71].
Figure 11. Cross sections generated from ExoMol line lists for nitric oxide [9] and phosphorous monosulfide [71].
Atoms 06 00026 g011
Figure 12. Cross sections for metal hydrides and NH. Line lists for lithium hydride [80] and scandium hydride [81] are theoretical while those for FeH and NH are derived from the experiments of the Bernath group [82,87,158,163].
Figure 12. Cross sections for metal hydrides and NH. Line lists for lithium hydride [80] and scandium hydride [81] are theoretical while those for FeH and NH are derived from the experiments of the Bernath group [82,87,158,163].
Atoms 06 00026 g012
Figure 13. Cross sections for hydride species based on ExoMol line lists for sodium hydride [59] and silicon monohydride [72], and the empirical titanium monohydride line list of Burrows et al. [88].
Figure 13. Cross sections for hydride species based on ExoMol line lists for sodium hydride [59] and silicon monohydride [72], and the empirical titanium monohydride line list of Burrows et al. [88].
Atoms 06 00026 g013
Figure 14. Cross sections for metal oxides generated using ExoMol line lists for silicon monoxide [52], aluminium monoxide [58] and vanadium monoxide [67]. The titanium monoxide cross sections are based on the computed line list due to Schwenke [89]. Also shown are short-wavelength silicon monoxide cross sections generated using line data from the database due to Kurucz [170].
Figure 14. Cross sections for metal oxides generated using ExoMol line lists for silicon monoxide [52], aluminium monoxide [58] and vanadium monoxide [67]. The titanium monoxide cross sections are based on the computed line list due to Schwenke [89]. Also shown are short-wavelength silicon monoxide cross sections generated using line data from the database due to Kurucz [170].
Atoms 06 00026 g014
Figure 15. Cross sections based on ExoMol line lists for carbon dimer [77] and H 3 + molecular ion [69].
Figure 15. Cross sections based on ExoMol line lists for carbon dimer [77] and H 3 + molecular ion [69].
Atoms 06 00026 g015
Table 1. Summary of MARVEL analyses performed for astronomically important molecules.
Table 1. Summary of MARVEL analyses performed for astronomically important molecules.
Molecule N iso N trans N E Reference(s)
H 2 O9182,15618,486[35,36,37,38]
H 3 + 31410911[39,40]
NH 3 128,5304961[41]
C 2 123,3435699[42]
TiO148,59010,564[43]
HCCH137,81311,213[44]
SO 2 340,26915,130[45]
H 2 S139,2677651[46]
ZrO121,1958329[47]
N iso : Number of isotopologues considered; N trans : Number of transitions validated for the main isotopologue; N E : Number of energy levels given for the main isotopologue.
Table 2. Datasets created by the ExoMol project and included in the ExoMol database [50].
Table 2. Datasets created by the ExoMol project and included in the ExoMol database [50].
PaperMolecule N iso T max N elec N lines DSNameReference
IBeH12000116,400Yadin[51]
IMgH32000110,354Yadin[51]
ICaH12000115,278Yadin[51]
IISiO590001254,675EJBT[52]
IIIHCN/HNC140001399,000,000Harris[30]
IVCH 4 1150019,819,605,160YT10to10[53]
VNaCl230001702,271Barton[54]
VKCl4300011,326,765Barton[54]
VIPN250001142,512YYLT[55]
VIIPH 3 11500116,803,703,395SAlTY[56]
VIIIH 2 CO11500110,000,000,000AYTY[57]
IXAlO4800034,945,580ATP[58]
XNaH27000279,898Rivlin[59]
XIHNO 3 150016,722,136,109AlJS[60]
XIICS830001548,312JnK[61]
XIIICaO15000521,279,299VBATHY[62]
XIVSO 2 1200011,300,000,000ExoAmes[63]
XVH 2 O 2 11250120,000,000,000APTY[64]
XIVH 2 S120001115,530,3730AYT2[65]
XVSO 3 1800121,000,000,000UYT2[66]
XVIVO1200013277,131,624VOMYT[67]
XIXH 2 17 , 18 O230001519,461,789HotWat78[68]
XXH 3 + 13000111,500,000,000MiZATeP[69]
XXINO6500022,281,042NOName[9]
XXIISiH 4 11200162,690,449,078OY2T[70]
XXIIIPO1500012,096,289POPS[71]
XXIIIPS15000330,394,544POPS[71]
XXIVSiH4500031,724,841SiGHTLY[72]
XXVSiS125000191,715UCTY[73]
XXVIHS650001219,463SNaSH[74]
XXVINS6500013,479,067SNaSH[74]
XXVIIC 2 H 4 1700149,841,085,051MaYTY[75]
XXVIIIAlH35000340,000AlHambra[76]
XXIXCH 3 Cl212001166,279,593,333OYT[75]
XXXH 2 16 O1500011,500,000,000Pokazatel[5]
XXXIC 2 3500086,080,9208State[77]
XXXIIMgO35000422,579,054LiPTY[78]
N iso : Number of isotopologues considered; T max : Maximum temperature for which the line list is complete; N elec : Number of electronic states considered; N lines : Number of lines, value is for the main isotopologue; DSName: Name of line list and of data set in ExoMol database [50].
Table 3. Other molecular line lists considered. These data can also be obtained from the ExoMol website.
Table 3. Other molecular line lists considered. These data can also be obtained from the ExoMol website.
Molecule N iso T max N elec N lines DSNameReferenceMethodology
Line PositionsIntensities
NH 3 2150011,138,323,351BYTe[79]empirical a ab initio a
LiH112,000118,982CLT[80]ab initioab initio
ScH1500061,152,827LYT[81]tuned ab initioab initio
NH1 110,41414BrBeWe[82]empiricalab initio
CH26000454,08614MaPlVa[83]empiricalab initio
CO990001752,97615LiGoRo[8]empiricalemp./ab initio b
OH16000145,00016BrBeWe[10]empiricalab initio
CN1 1195,12014BrRaWe[84]empiricalab initio
CP1 128,73514RaBrWe[85]empiricalab initio
HCl1 1258811LiGoBe[86]empiricalab initio
FeH1 293,04010WEReSe[87]empiricalab initio
TiH1 3181,08005BuDuBa[88]empiricalab initio
CO 2 1310001149,587,373Ames-2016[7]empiricalab initio
TiO1 1345,000,000Schwenke[89]tuned ab initioab initio
C 2 H 2 11000133,890,981ASD-1000[90]effect. Hamilt. c effect. dipole c
CrH1 213,82402BuRaBe[91]empiricalab initio
CH 3 F140011,391,882,159OYKYT[92]ab initioab initio
N iso : Number of isotopologues considered; T max : Maximum temperature for which the line list is complete (where stated); N elec : Number of electronic states considered; N lines : Number of lines: value is for the main isotopologue; DSName: Name of line list and of data set in ExoMol database [50]; a ExoMol methodology, see text; b A mixture of empirical and ab initio values of the dipole moment function; c Empirical, based on Effective Hamiltonian and Dipole moment expansions.

Share and Cite

MDPI and ACS Style

Tennyson, J.; Yurchenko, S.N. The ExoMol Atlas of Molecular Opacities. Atoms 2018, 6, 26. https://doi.org/10.3390/atoms6020026

AMA Style

Tennyson J, Yurchenko SN. The ExoMol Atlas of Molecular Opacities. Atoms. 2018; 6(2):26. https://doi.org/10.3390/atoms6020026

Chicago/Turabian Style

Tennyson, Jonathan, and Sergei N. Yurchenko. 2018. "The ExoMol Atlas of Molecular Opacities" Atoms 6, no. 2: 26. https://doi.org/10.3390/atoms6020026

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop