H2O2 induced greenhouse warming on oxidized early Mars
Abstract
The existence of liquid water within an oxidized environment on early Mars has been inferred by the Mn-rich rocks found during recent explorations on Mars. The oxidized atmosphere implied by the Mn-rich rocks would basically be comprised of CO2 and H2O without any reduced greenhouse gases such as H2 and CH4. So far, however, it has been thought that early Mars could not have been warm enough to sustain water in liquid form without the presence of reduced greenhouse gases. Here, we propose that H2O2 could have been the gas responsible for warming the surface of the oxidized early Mars. Our one-dimensional atmospheric model shows that only 1 ppm of H2O2 is enough to warm the planetary surface because of its strong absorption at far-infrared wavelengths, in which the surface temperature could have reached over 273 K for a CO2 atmosphere with a pressure of 3 bar. A wet and oxidized atmosphere is expected to maintain sufficient quantities of H2O2 gas in its upper atmosphere due to its rapid photochemical production in slow condensation conditions. Our results demonstrate that a warm and wet environment could have been maintained on an oxidized early Mars, thereby suggesting that there may be connections between its ancient atmospheric redox state and possible aqueous environment.
1 Introduction
One of the most intriguing and debatable problems in planetary science is elucidating how an early Martian surface environment could have been warm enough to sustain liquid water (Wordsworth, 2016; Ramirez & Craddock, 2018). Climate models have shown that a CO2–H2O atmosphere alone could not have kept early Mars warm enough to sustain liquid water globally even if any amount of atmospheric pressure is assumed (e.g., Kasting, 1991). This suggests that the other greenhouse components could have played a key role in the early Martian atmosphere. Current theoretical models suggest that the warming of early Mars was caused by a CO2–H2O atmosphere combined with additional greenhouse substances; clouds (e.g., Forget & Pierrehumbert, 1997; Wordsworth et al., 2013); reducing gases such as H2, CH4, and NH3 (e.g., Ramirez et al., 2014; Ramirez, 2017; Wordsworth et al., 2017; Sagan & Mullen, 1972; Kasting et al., 1992); and/or volcanic gases such as H2S and SO2 (e.g., Postawko & Kuhn, 1986; Johnson et al., 2008; Tian et al., 2010).
Recently, NASA’s Curiosity rover discovered a high abundance of Mn in sedimentary rocks (Lanza et al., 2016). During the era in which the observed Mn-oxide-rich rocks at Gale crater would have precipitated out, Mars may have had both liquid water on its surface and a highly oxidized atmosphere (Lanza et al., 2016; Noda et al., 2019). Furthermore, the existence of rocks with a high concentration of manganese at Endeavour crater (Arvidson et al., 2016) suggests that such an oxidized and wet surface environment was a global phenomenon at that time. These findings suggest that the early Martian surface had once experienced a wet and warm environment, but with the absence of reduced gas species that would have enhanced the greenhouse effect of a CO2–H2O dominated Martian atmosphere to allow the existence of liquid water. One might consider SO2 as a candidate greenhouse gas in an oxidized atmosphere (e.g., Johnson et al., 2008), but its presence seems unlikely during this era because Mn and S are not correlated in the rocks found at Gale crater (Lanza et al., 2016).
In an attempt to address this uncertainty, in this study we investigate the greenhouse effect due to hydrogen peroxide (H2O2) gas in the early Martian atmosphere. Previously, it had been proposed that H2O2 gas was responsible for oxidizing the early Martian surface (e.g., Zahnle et al., 2008). Although the idea that H2O2 was one of the greenhouse gases responsible for the warming of early Mars has widely been ignored, H2O2 does absorb radiation at wavenumbers near 500 cm-1, where the blackbody radiation at a temperature of K has peak intensity and CO2 has an absorption window, as shown in Fig. 1 (see also Figure 4 in Wordsworth, 2016). Also, the absorption cross section of H2O2 is larger than those of known greenhouse gases such as SO2, NH3, CH4 and OCS in a wavenumber range from 250 cm-1 to 450 cm-1, as shown in Fig. 2.
The remainder of this paper is organized as follows. In Section 2, we describe our atmospheric model and numerical setup. In Section 3 we show the surface temperature as a function of H2O2 abundance under the conditions that may have been present on early Mars. We discuss the photochemical production and condensation of H2O2 in a warm and wet early Martian atmosphere and the possible warming scenario of H2O2 in an oxidized early Martian environment in Section 4. Finally we summarize our results in Section 5.


2 Atmospheric Model
We set up a vertical, one-dimensional CO2-dominant atmospheric model and determine the surface temperature required to achieve balance between the absorbed solar radiation and the outgoing planetary radiation with approximated radiative–convective equilibrium temperature-pressure profiles and given compositions. The numerical scheme is based on the same line-by-line calculations used in the calculations of the surface temperature warmed by H2O (Schaefer et al., 2016) and CO2–H2–CH4 atmospheres (Wordsworth et al., 2017).
The atmosphere in hydrostatic equilibrium is vertically divided into 100 layers from the ground to the top of the atmosphere ( bar for the model atmosphere). The surface pressure ranges from 0.01–3 bar. The vertical grid of the atmosphere is set so that the logarithms of pressure are evenly spaced. Following previous models (Ramirez et al., 2014; Ramirez, 2017), we set the modeled atmosphere to one composed of 95% CO2, fully-saturated H2O, fully-saturated or different mixing ratios of H2O2, and 5 % N2. For the saturated H2O2 amount, we calculate the vapor pressure using the Clausius-Clapeyron equation. Then, we use the thermal properties of H2O2 (Foley & Giguére, 1951a) and its saturation vapor pressure of bar at the melting point (272.69 K) as a reference pressure (Manatt & Manatt, 2004). In the other cases, the molar fraction of H2O2 is assumed to be vertically constant and its value was in the range from 10 ppb to 10 ppm. Additionally, the abundance of H2O is determined by the saturation vapor pressure of water (Eqs. 11 and 12 in Kasting et al., 1984).
The atmospheric temperature profile is assumed to be that of a moist adiabat of H2O and CO2 from the surface to the tropopause and an isothermal stratosphere above the tropopause. Using the heat capacity, , given by the Shomate equation222http://old.vscht.cz/fch/cz/pomucky/fchab/Shomate.html, the vapor amount and latent heat of H2O (Eq.12 in Kasting et al., 1984) and the gravity of Mars, , the moist adiabat of H2O is given by Eq. (2.48) in Andrews (2000). Decreasing gravity with altitude is included in this model. The moist adiabat of CO2 is adopted where the moist adiabat of H2O is colder than the saturation vapor pressure of CO2 using Eqs. A5 and A6 in Kasting (1991). We assume that the stratospheric temperature is 155 K ( K), which is based on the results of Kasting (1991), who uses 167 K as the stratospheric temperature for current solar heating and scales it for different solar heating rates by assuming that the stratospheric temperature is proportional to the skin temperature.
Using the atmospheric structure described above, we calculate the outgoing planetary radiation based on a line-by-line radiative transfer calculation. The outgoing planetary radiation is given by
(1) | |||||
where is the surface temperature, is the cosine of the zenith angle, is the Planck function at wavenumber, , and is the total optical depth of the atmosphere. In hydrostatic equilibrium, the optical depth is given by
(2) |
where is atmospheric pressure, is the mean mass of the atmospheric gas particles, and and are the molar fraction and absorption cross section of an absorber , respectively. Following Ramirez (2017) and Kopparapu et al. (2013), the line absorption cross section profile of CO2 is assumed to be a sub-Lorentzian (Perrin & Hartmann, 1989) truncated at 500 cm-1 from the line center, while that of H2O is assumed to be a Voigt profile truncated at 25 cm-1 from the line center. The line profile of H2O2 is also assumed to be a Voigt profile truncated at 25 cm-1 from the line center. For the line absorption of each gas species, the line data given by HITRAN2012 (Rothman et al., 2013) and the line profile calculation code EXOCROSS (Yurchenko et al., 2018) are used in this model. Additionally, the collision-induced absorption of CO2–CO2 (Gruszka & Borysow, 1997; Baranov et al., 2004) is considered. In practice, to save memory and CPU time, we have prepared a numerical table in which the absorption cross sections are given as functions of temperature, , and log10 . The table was created using values of and 350 K, and log10 (/bar) and 1. We evaluate the integral shown in Eq. (1) over a wavenumber range from 1 cm-1 to 10000 cm-1 with a resolution of 1 cm-1. The numerical integration of Eq. (1) with respect to the zenith angle is performed using the exponential integral calculation code presented by Press et al. (1996), while the other integrals are evaluated using trapezoidal integration.
We iteratively determine the surface temperature at which the outgoing planetary radiation balances the absorbed solar radiation. The absorbed solar radiation is given by , where is the planetary albedo and is the solar flux. The solar flux is assumed to be W/m2, and we use the planetary albedo of a wet CO2(95%)-N2(5%) atmosphere not warmed by any additional greenhouse mechanism (Ramirez et al., 2014, private communication). Note that our assumed planetary albedo underestimates the surface temperature in a warm atmosphere with enhanced saturated–H2O content more than the atmosphere not warmed by H2O2. This is because the absorption of solar radiation by H2O decreases the planetary albedo (Kasting, 1988), and H2O2 might work in the same way. While the Rayleigh scattering cross section per a H2O2 molecule is comparable with that of CO2 based on its electric dipole polarizability (Maroulis, 1992), the abundance of H2O2 in our model is too low (up to 10 ppm) to increase the planetary albedo. Also, though there is no public absorption data of H2O2 in the optical regime , its absorption in the optical is likely to not be very strong (see also the MPI–Mainz UV/VIS Spectral Atlas333http://satellite.mpic.de/spectral_atlas/index.html; Keller-Rudek et al., 2013).
2.1 MODEL VALIDATION


We have performed two benchmark tests of our simulation code, and we have confirmed that our model reproduce the numerical solutions for the dry and wet CO2-rich atmospheres of early Mars shown in Ramirez et al. (2014).
We first compare our line-by-line model against a well-tested line-by-line model, SMART (Meadows & Crisp, 1996), for a dry, 2-bar CO2(95%)-N2(5%) atmosphere. With the same temperature profile shown in Figure S1 of Ramirez et al. (2014), we calculated the outgoing planetary radiation using our model. Fig. 3 shows a comparison between the SMART result and that from our model. Our model spectra agree well with the SMART spectra, although there is some difference in the wavenumber region from 800 cm-1 to 1200 cm-1, which is likely due to the different absorption data used in both studies. The total flux of our model is 87.2 W/m2, which agrees well with that found by SMART (88.4 W/m2). Note that the calculated fluxes differ by at most 0.05 %, even if we double the resolution of the wavenumber or the number of vertical layers.
Next, we compare the surface temperatures of a wet, 2-bar CO2(95%)-N2(5%) atmosphere with that calculated by the one-dimensional radiative convective model Ramirez et al. (2014). Our results agrees well with those of Ramirez et al. (2014), as shown in Fig. 4, where the differences in the calculated surface temperatures are no more than 4 K. Because the results of our models agree to within 2 % of the previous studies, we have confirmed that our model is consistent with these models.
3 Results
() Outgoing planetary radiation

() Atmospheric structure


Fig. 5 (a) shows the outgoing planetary radiation for a fixed surface pressure and temperature of 2 bar and 273 K, respectively. When the atmosphere consists of H2O and CO2 (cyan), there are atmospheric windows at wavenumbers below 500 cm-1 and around 1000 cm-1, which are consistent with the results of previous climate models (e.g., Wordsworth, 2016; Ramirez, 2017). The addition of H2O2 reduces the planetary radiation at wavenumbers below 500 cm-1 due to its strong far–IR absorption (blue, olive, green). Although H2O2 effectively absorbs photons with wavenumbers around 1200 cm-1(Fig. 1), this only slightly affects the outgoing planetary radiation because CO2 also absorbs photons at the same wavenumbers.
The outgoing planetary radiation is 87.6 W/m2 for an H2O2 free atmosphere (cyan), which decreases drastically when H2O2 is added. For vertically constant molar fractions of 1 ppm (olive) and 10 ppm (green) of H2O2, the outgoing planetary radiation is 68.8 W/m2 and 56.5 W/m2, respectively. If the abundance of H2O2 can be constrained by the saturation vapor pressure, the planetary radiation is 84.2 W/m2 (blue), and the greenhouse effect of H2O2 is not remarkable. This is because the abundance of saturated H2O2 is too low in the low pressure region to absorb photons effectively (Fig. 5 (b)).
Next, Fig.6 shows the surface temperature as a function of surface pressure. The differences in surface temperatures between the atmospheres without H2O2 (black) and with saturated H2O2 (blue) are at most 4 K. However, in the case of abundant H2O2, the planetary surface is warm enough to sustain liquid water (Fig.6). In particular, for the 2 bar atmosphere with added 1 ppm (olive) or 10 ppm (green) of H2O2, the surface temperature increases by about 40 K or 65 K from that of the H2O2-free case ( 230 K), respectively. Our results show that a concentration of only 1 ppm level of H2O2 is sufficient to effectively cut off the outgoing planetary radiation and warm the planetary surface to temperatures above 273 K.
4 Discussion
4.1 H2O2 in a wet and oxidized atmosphere
H2O2 is much more abundant in a wet and oxidized atmosphere, though the concentration of H2O2 in the current dry Martian atmosphere is about 10 ppb (Encrenaz et al., 2004). Although chemical models suggest that the concentration of H2O2 reaches at most 0.1 ppm in dry atmospheres (Parkinson & Hunten, 1972; Gao et al., 2015), a wet and oxidized atmosphere which is suitable for the formation of H2O2 would contain H2O2 in a concentration higher than 0.1 ppm. This is because H2O2 is produced through the chemical reactions of HOx gas species such as H, OH and HO2 which originate from H2O. Also, the abundance of H2O2 would be higher in an oxidized atmosphere because such an atmosphere inhibits the regeneration of H2O from HOx and enhances the production of H2O2.
In a wet and oxidized Martian atmosphere, the photolysis of H2O2 is considered to be an effective pathway to regenerate CO2 (Yung & Demore, 1999). This regeneration is necessary because CO2 is destroyed by far-UV irradiation (227.5 nm) from the Sun via;
(R1) |
Indeed CO2 regeneration is required to maintain the CO2 atmosphere over geological timescales. In a wet atmosphere, H2O2 can be sufficiently produced as an intermediate product through the following catalytic cycle:
(R2) | ||||
(R3) | ||||
(R4) | ||||
(R5) | ||||
(S1) |
Meanwhile, although a thick and dry CO2-rich atmosphere is unstable (Zahnle et al., 2008), in a wet and oxidized atmosphere of early Mars, CO2 could have been stabilized by S1 (=R2+R3+R4+R5) even if the atmosphere was thick.
We estimate the H2O2 abundance in a warm/wet and oxidized CO2 atmosphere by assuming that S1 is the cycle most responsible for the regeneration of CO2 against loses due to R1. We assume that the bulk CO2 abundance in the atmosphere is in balance between its photo-dissociation flux (R1), and twice the H2O2 photo-dissociation flux (R4) that produces OH for oxidizing CO. Also, it is assumed that there is no optical shielding effect for photons with wavelengths longer than the shielded wavelength, , but there is complete shielding for all other UV photons to H2O2, for simplicity. Then, the vertical column density of H2O2, , can be written as;
(3) |
where is solar photon flux, and and are the photo-dissociation cross section and the threshold wavelength for a photon to effectively dissociate H2O2, respectively. Owing to the low bonding energy of H2O2 ( 50 kcal/mol 570 nm; Bach et al. 1996), the photo-dissociation is caused not only by UV but also by visible light photons. Therefore, H2O2 is not completely shielded from stellar irradiation by H2O, O2 and CO2 (Yung & Demore, 1999). On the other hand, a developed O3 layer may shield solar photons with wavelengths 300 nm, as displayed on Earth today. Here we use nm and 300 nm as fiducial values of a shielded wavelength.
The dissociation cross section of H2O2 has been measured only for photon wavelengths in the range 410 nm (Kahan et al., 2012) because of the technical problem of measuring small absorption cross sections. Hence, we use nm as a fiducial value of the threshold wavelength. Note that, according to Kahan et al. (2012), the photolysis of H2O2 mainly occurs at photon wavelengths shorter than 350 nm. Therefore, inputing nm, the measured cross section with 410 nm (Lin et al., 1978; Kahan et al., 2012) and the solar spectral irradiance at 4 Ga developed by combining the observed spectrum from the Sun with those of solar-type stars at different ages (Claire et al., 2012) in Eq. (3), we find cm-2. When we substitute nm into Eq. (3), we find cm-2. These values change only 10 % if the solar spectral irradiance at 3.5 Ga is used instead.
The column densities estimated here are significantly larger than the current typical value of cm-2, which corresponds to 10 ppb at 6 mbar, in the present-day Martian atmosphere. These large column densities produce optical depth over wavenumbers –500 cm-1 of –0.6 for nm and –4 for nm, assuming a far-IR absorption cross section of H2O2, – cm2, which is shown in Fig. 1. Thus, if the other gases such as O3 sufficiently can reduce the photolysis of H2O2, then the amount of H2O2 in the atmosphere would be large enough to warm the planetary surface. Note that, the column density of H2O2 estimated by Eq. (3) is just a typical value when S1 is the cycle most responsible for the stabilization of CO2, while this value could be increased if the self-shielding effect was taken into account in Eq. (3). This is because we impose the restriction that only the OH produced by the photolysis of H2O2 is used to oxidize CO via R5, but all other reactions which produce and remove OH are ignored. Also, the other process potentially affecting the concentration of H2O2 is discussed in Sec 4.3.
4.2 Condensation of H2O2
It is likely that H2O2 in a warm and wet atmosphere of early Mars is super-saturated because the timescale for condensation is likely longer than that for photochemical production. As described later, a timescale for condensation would be much longer than that that governing the production and photo-dissociation of H2O2, which was shown by Nair et al. (1994), who used a photochemical model, to be of order several hours.
The condensation time can be estimated by assuming that H2O2 condenses as soon as it collides with condensation nuclei, namely;
where and are the size and concentration of the condensation nuclei, respectively, is the atmospheric mass density and is the thermal velocity of the gas. The timescale for condensation is longer at higher altitudes because the nuclei concentration decreases with increasing altitude. Note that, the condensation timescale is underestimated in an atmospheric region with a mean free path smaller than the size of the nuclei (i.e., a dense region) because the diffusive motion of the gas around the nuclei delays the timescale (see Lohmann et al., 2016, for the diffusive case).

Achieving sufficient warming is possible even if H2O2 condenses at lower altitudes due to the subsequent shorter condensation times. Fig. 7 shows the minimum H2O2 concentration necessary for maintaining a surface temperature of at least 273 K in a 2-bar atmosphere as a function of a condensation altitude. The condensation altitude stands for an altitude above which the H2O2 concentration is constant and below which all the H2O2 gas is virtually removed by rainout through condensation. To warm the surface environment, the required concentration of H2O2 needs to be about 2 ppm when the condensation altitude is no higher than about 20 km. The 2 ppm of H2O2 in the upper atmosphere is comparable to 1.51019 cm-2 which is also comparable to the H2O2 column density necessary to stabilize the CO2 atmosphere (Sec 4.1).
The condensation timescale will not be significantly changed if the dilution effect of H2O2 in an H2O solution is taken into account. When the temperature is above K, an H2O–H2O2 solution can exist, and then the saturation vapor pressure of H2O2 will be lowered relative to that of pure H2O2 (Foley & Giguére, 1951b; Manatt & Manatt, 2004). However, the temperatures in the photosphere for photons with wavenumbers in the range 500 cm-1 are lower than 220 K in thick and warm atmospheres (Fig. 5). So, it is likely that aqueous solutions would be frozen in the upper atmosphere where the concentration of H2O2 has the greatest influence on the surface temperature. Therefore, the dilution effect of H2O2 in an H2O solution would little affect the surface temperature.
4.3 Other processes possibly affecting H2O2 concentration
The atmospheric concentration of H2O2 can also be affected by several processes such as dissolution into water droplets, dry deposition and photochemical reactions with volcanic and reactive species (e.g., SO2 and NOx) (Vione et al., 2003).
Although H2O2 is a minor species with at most 3.5 ppb level in the Earth’s atmosphere, which is mainly due to the dissolution of gaseous H2O2 into water droplets, where SO2 enhances the dissolution rate (Vione et al., 2003), it might not be a minor species on early Mars during the era in which the observed Mn-oxide-rich rocks at Gale crater would have precipitated out. Since the temperatures at high altitudes in the early Martian atmosphere would be so low that H2O would freeze, its non-dissolution into water droplets would not deplete H2O2. Meanwhile, at lower altitude regions, H2O2 would dissolve into water droplets, and precipitation would remove it.
In Earth’s atmosphere, dry deposition is another removal process of atmospheric H2O2 at lower altitudes. Atmospheric H2O2 of early Mars would be vertically transported by eddy diffusion to the surface, whereby dry deposition and precipitation remove it. For the current Martian atmosphere at altitudes lower than 40 km, the scale height is km and the vertical eddy diffusion coefficient is cm2/s (Nair et al., 1994); hence the diffusion timescale is days. Since the timescale of H2O2 photochemical reactions is less than a day (Nair et al., 1994; Zahnle et al., 2008), the atmospheric concentration of H2O2 at high altitudes is likely controlled by photochemical reactions.
The actual eddy diffusion coefficient and dry deposition timescale on early Mars would depend on turbulence/large-scale-winds and the compositions/oxidations states of the surface rocks, respectively. As such, a more detailed examination requires that photochemical calculations be done alongside those of the atmospheric thermal structure, which will be the focus of a future study.
4.4 Oxidized early Martian environment
An early surface environment warmed by the greenhouse effect of H2O2 (Sec. 4.1) is consistent with the global, highly oxidized conditions implied by the high Mn materials found on the Martian surface by the Curiosity rover in Gale crater and by the Opportunity rover in Endeavour crater (Lanza et al., 2016; Arvidson et al., 2016).
The redox state of early Martian atmosphere is likely controlled by the escape of atmospheric components into space. In the early Martian atmosphere, UV radiation from the young Sun would have enhanced hydrogen escape and effectively oxidized the atmosphere and the surface environment. In addition to hydrogen escape, the escape of atomic carbon might also have contributed to the oxidation of the early Martian atmosphere because its escape flux would not be limited by diffusion in a CO2-rich atmosphere (N. Terada, private communications). Further studies are required to determine the redox state of the early Martian atmosphere, which could also be affected by the supply of reduced gases (e.g., CO and H2) through volcanic degassing, oxygen escape, and oxygen uptake through weathering of the planetary surface (Zahnle et al., 2008; Wetzel et al., 2013; Batalha et al., 2015).
It is interesting to note that H2O2 might be able to warm a frozen planet and melt water ice. Liang et al. (2006) demonstrated that a weak hydrological cycle coupled with photochemical reactions could give rise to a sustained production of H2O2 during long and severe glacial intervals. Although an icy surface has a high albedo, the surface temperature can be warmed to temperatures above 273 K by a 4 and 15 ppm levels of H2O2 in a 2 bar atmosphere when the planetary albedo is assumed to be 0.45 and 0.5, respectively, as demonstrated by our model.
It has also been suggested that H2O2 deposited on the planetary surface could be stored in the ice during the time of a global snowball episode (Liang et al., 2006). If early Mars was once a snowball, and a large amount of H2O2 was stored in the ice, it would be released into the atmosphere upon melting caused by any mechanism, such as meteor impacts, volcanic emissions, or obliquity changes (e.g., Wordsworth, 2016, references therein). The release of abundant H2O2 would cause not only a global oxidation event but also enhance greenhouse warming. If so, there might be geological evidence that oxidation and warming occurred simultaneously in the aftermath of a snowball Mars.
5 Summary and Conclusion
We investigated the possible impact of H2O2 as an additional greenhouse gas in a CO2-dominant atmosphere using a one-dimensional atmospheric model. Because the timescale for condensation is longer at higher altitudes (subsection 4.2), photochemically produced H2O2 would likely be supersaturated in the upper atmosphere. We found that a reasonable amount of H2O2 in the upper atmosphere effectively cuts off the outgoing planetary radiation in the far-infrared and warms the planetary surface to a temperature hot enough to retain liquid water (Section 3).
Our results demonstrated that a warm and wet surface environment is compatible with an oxidized atmosphere on early Mars. The coexistence of liquid water and an oxidized atmosphere on early Mars has been suggested by the recent discovery of a high level of Mn in some Martian rocks (Lanza et al., 2016; Arvidson et al., 2016). Our results also indicated a key relationship between the redox state of the atmosphere and the surface temperature on early Mars, where the co-evolution of these factors may govern the surface environment over geological time scales. This important phenomenon will be the subject of future work, which will aim to understand the surface environment under an oxidized atmosphere on early Mars.
References
- Al-Refaie et al. (2016) Al-Refaie, A. F., Polyansky, O. L., Ovsyannikov, R. I., Tennyson, J., & Yurchenko, S. N. 2016, MNRAS, 461, 1012, doi: 10.1093/mnras/stw1295
- Andrews (2000) Andrews, D. G. 2000, An Introduction to Atmospheric Physics, 240
- Arvidson et al. (2016) Arvidson, R. E., Squyres, S. W., Morris, R. V., et al. 2016, American Mineralogist, 101, 1389, doi: 10.2138/am-2016-5599
- Bach et al. (1996) Bach, R. D., Ayala, P. Y., & Schlegel, H. B. 1996, Journal of the American Chemical Society, 118, 12758, doi: 10.1021/ja961838i
- Baranov et al. (2004) Baranov, Y. I., Lafferty, W. J., & Fraser, G. T. 2004, Journal of Molecular Spectroscopy, 228, 432, doi: 10.1016/j.jms.2004.04.010
- Batalha et al. (2015) Batalha, N., Domagal-Goldman, S. D., Ramirez, R., & Kasting, J. F. 2015, Icarus, 258, 337, doi: 10.1016/j.icarus.2015.06.016
- Bibring et al. (2005) Bibring, J.-P., Langevin, Y., Gendrin, A., et al. 2005, Science, 307, 1576, doi: 10.1126/science.1108806
- Claire et al. (2012) Claire, M. W., Sheets, J., Cohen, M., et al. 2012, ApJ, 757, 95, doi: 10.1088/0004-637X/757/1/95
- Encrenaz et al. (2004) Encrenaz, T., Bézard, B., Greathouse, T. K., et al. 2004, Icarus, 170, 424, doi: 10.1016/j.icarus.2004.05.008
- Fedorova et al. (2020) Fedorova, A. A., Montmessin, F., Korablev, O., et al. 2020, Science, 367, 297, doi: 10.1126/science.aay9522
- Foley & Giguére (1951a) Foley, W. T., & Giguére, P. A. 1951a, Canadian Journal of Chemistry, 29, 895, doi: 10.1139/v51-104
- Foley & Giguére (1951b) —. 1951b, Canadian Journal of Chemistry, 29, 123, doi: 10.1139/v51-016
- Forget & Pierrehumbert (1997) Forget, F., & Pierrehumbert, R. T. 1997, Science, 278, 1273, doi: 10.1126/science.278.5341.1273
- Galeazzo et al. (2018) Galeazzo, T., Bekki, S., Martin, E., Savarino, J., & Arnold, S. R. 2018, Atmospheric Chemistry and Physics, 18, 17909, doi: 10.5194/acp-18-17909-2018
- Gao et al. (2015) Gao, P., Hu, R., Robinson, T. D., Li, C., & Yung, Y. L. 2015, ApJ, 806, 249, doi: 10.1088/0004-637X/806/2/249
- Gendrin et al. (2005) Gendrin, A., Mangold, N., Bibring, J.-P., et al. 2005, Science, 307, 1587, doi: 10.1126/science.1109087
- Gruszka & Borysow (1997) Gruszka, M., & Borysow, A. 1997, Icarus, 129, 172, doi: 10.1006/icar.1997.5773
- Johnson et al. (2008) Johnson, S. S., Mischna, M. A., Grove, T. L., & Zuber, M. T. 2008, Journal of Geophysical Research (Planets), 113, E08005, doi: 10.1029/2007JE002962
- Kahan et al. (2012) Kahan, T. F., Washenfelder, R. A., Vaida, V., & Brown, S. S. 2012, Journal of Physical Chemistry A, 116, 5941, doi: 10.1021/jp2104616
- Kasting (1988) Kasting, J. F. 1988, Icarus, 74, 472, doi: 10.1016/0019-1035(88)90116-9
- Kasting (1991) —. 1991, Icarus, 94, 1, doi: 10.1016/0019-1035(91)90137-I
- Kasting et al. (1992) Kasting, J. F., Brown, L. L., Acord, J. M., & Pollack, J. B. 1992, in Martian Surface and Atmosphere Through Time, ed. R. M. Haberle & B. M. Jakosky, 84
- Kasting et al. (1984) Kasting, J. F., Pollack, J. B., & Ackerman, T. P. 1984, Icarus, 57, 335, doi: 10.1016/0019-1035(84)90122-2
- Keller-Rudek et al. (2013) Keller-Rudek, H., Moortgat, G. K., Sander, R., & Sörensen, R. 2013, Earth System Science Data, 5, 365, doi: 10.5194/essd-5-365-2013
- Kopparapu et al. (2013) Kopparapu, R. K., Ramirez, R., Kasting, J. F., et al. 2013, ApJ, 765, 131, doi: 10.1088/0004-637X/765/2/131
- Lanza et al. (2016) Lanza, N. L., Wiens, R. C., Arvidson, R. E., et al. 2016, Geophys. Res. Lett., 43, 7398, doi: 10.1002/2016GL069109
- Liang et al. (2006) Liang, M.-C., Hartman, H., Kopp, R. E., Kirschvink, J. L., & Yung, Y. L. 2006, Proceedings of the National Academy of Science, 103, 18896, doi: 10.1073/pnas.0608839103
- Lin et al. (1978) Lin, C. L., Rohatgi, N. K., & Demore, W. B. 1978, Geophys. Res. Lett., 5, 113, doi: 10.1029/GL005i002p00113
- Lohmann et al. (2016) Lohmann, U., Lüönd, F., & Mahrt, F. 2016, An Introduction to Clouds: From the Microscale to Climate (Cambridge University Press), doi: 10.1017/CBO9781139087513
- Maltagliati et al. (2011) Maltagliati, L., Montmessin, F., Fedorova, A., et al. 2011, Science, 333, 1868, doi: 10.1126/science.1207957
- Manatt & Manatt (2004) Manatt, S. L., & Manatt, M. R. R. 2004, Chemistry – A European Journal, 10, 6540, doi: 10.1002/chem.200400104
- Maroulis (1992) Maroulis, G. 1992, The Journal of Chemical Physics, 96, 6048, doi: 10.1063/1.462646
- Meadows & Crisp (1996) Meadows, V. S., & Crisp, D. 1996, J. Geophys. Res., 101, 4595, doi: 10.1029/95JE03567
- Nair et al. (1994) Nair, H., Allen, M., Anbar, A. D., Yung, Y. L., & Clancy, R. T. 1994, Icarus, 111, 124, doi: 10.1006/icar.1994.1137
- Noda et al. (2019) Noda, N., Imamura, S., Sekine, Y., et al. 2019, Journal of Geophysical Research (Planets), 124, 1282, doi: 10.1029/2018JE005892
- Parkinson & Hunten (1972) Parkinson, T. D., & Hunten, D. M. 1972, Journal of Atmospheric Sciences, 29, 1380, doi: 10.1175/1520-0469(1972)029<1380:SAAOOO>2.0.CO;2
- Perrin & Hartmann (1989) Perrin, M. Y., & Hartmann, J. M. 1989, J. Quant. Spec. Radiat. Transf., 42, 311, doi: 10.1016/0022-4073(89)90077-0
- Postawko & Kuhn (1986) Postawko, S. E., & Kuhn, W. R. 1986, J. Geophys. Res., 91, D431, doi: 10.1029/JB091iB04p0D431
- Press et al. (1996) Press, W. H., Teukolsky, S. a., Vetterling, W. T., & Flannery, B. P. 1996, Numerical Recipes in Fortran 77: the Art of Scientific Computing. Second Edition, Vol. 1
- Ramirez (2017) Ramirez, R. M. 2017, Icarus, 297, 71, doi: 10.1016/j.icarus.2017.06.025
- Ramirez & Craddock (2018) Ramirez, R. M., & Craddock, R. A. 2018, Nature Geoscience, 11, 230, doi: 10.1038/s41561-018-0093-9
- Ramirez & Kasting (2017) Ramirez, R. M., & Kasting, J. F. 2017, Icarus, 281, 248, doi: 10.1016/j.icarus.2016.08.016
- Ramirez et al. (2014) Ramirez, R. M., Kopparapu, R., Zugger, M. E., et al. 2014, Nature Geoscience, 7, 59, doi: 10.1038/ngeo2000
- Rothman et al. (2013) Rothman, L. S., Gordon, I. E., Babikov, Y., et al. 2013, J. Quant. Spec. Radiat. Transf., 130, 4, doi: 10.1016/j.jqsrt.2013.07.002
- Sagan & Mullen (1972) Sagan, C., & Mullen, G. 1972, Science, 177, 52, doi: 10.1126/science.177.4043.52
- Schaefer et al. (2016) Schaefer, L., Wordsworth, R. D., Berta-Thompson, Z., & Sasselov, D. 2016, ApJ, 829, 63, doi: 10.3847/0004-637X/829/2/63
- Spracklen et al. (2005) Spracklen, D. V., Pringle, K. J., Carslaw, K. S., Chipperfield, M. P., & Mann, G. W. 2005, Atmospheric Chemistry and Physics, 5, 2227, doi: 10.5194/acp-5-2227-2005
- Tennyson & Yurchenko (2018) Tennyson, J., & Yurchenko, S. N. 2018, Atoms, 6, doi: 10.3390/atoms6020026
- Tian et al. (2010) Tian, F., Claire, M. W., Haqq-Misra, J. D., et al. 2010, Earth and Planetary Science Letters, 295, 412, doi: 10.1016/j.epsl.2010.04.016
- Vione et al. (2003) Vione, D., Maurino, V., Minero, C., & Pelizzetti, E. 2003, Annali di chimica, 93, 477
- Wetzel et al. (2013) Wetzel, D. T., Rutherford, M. J., Jacobsen, S. D., Hauri, E. H., & Saal, A. E. 2013, Proceedings of the National Academy of Science, 110, 8010, doi: 10.1073/pnas.1219266110
- Wordsworth et al. (2013) Wordsworth, R., Forget, F., Millour, E., et al. 2013, Icarus, 222, 1, doi: 10.1016/j.icarus.2012.09.036
- Wordsworth et al. (2017) Wordsworth, R., Kalugina, Y., Lokshtanov, S., et al. 2017, Geophys. Res. Lett., 44, 665, doi: 10.1002/2016GL071766
- Wordsworth (2016) Wordsworth, R. D. 2016, Annual Review of Earth and Planetary Sciences, 44, 381, doi: 10.1146/annurev-earth-060115-012355
- Yung & Demore (1999) Yung, Y. L., & Demore, W. B., eds. 1999, Photochemistry of planetary atmospheres: Oxford University Press, QB603.A85 Y86 1999, doi: 10.1021/ja9957938
- Yurchenko et al. (2018) Yurchenko, S. N., Al-Refaie, A. F., & Tennyson, J. 2018, A&A, 614, A131, doi: 10.1051/0004-6361/201732531
- Zahnle et al. (2008) Zahnle, K., Haberle, R. M., Catling, D. C., & Kasting, J. F. 2008, Journal of Geophysical Research (Planets), 113, E11004, doi: 10.1029/2008JE003160