Photochemical Oxygen in Non-1-bar CO2 Atmospheres of Terrestrial Exoplanets

and

Published 2018 October 25 © 2018. The American Astronomical Society. All rights reserved.
, , Citation Tre'Shunda James and Renyu Hu 2018 ApJ 867 17 DOI 10.3847/1538-4357/aae2bb

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/867/1/17

Abstract

Atmospheric chemistry models have shown that molecular oxygen can build up in CO2-dominated atmospheres on potentially habitable exoplanets without input of life. Existing models typically assume a surface pressure of 1 bar. Here we present model scenarios of CO2-dominated atmospheres with the surface pressure ranging from 0.1 to 10 bars, while keeping the surface temperature at 288 K. We use a one-dimensional photochemistry model to calculate the abundance of O2 and other key species, for outgassing rates ranging from a Venus-like volcanic activity up to 20 times Earth-like activity. The model maintains the redox balance of the atmosphere and the ocean, and includes the pressure dependency of outgassing on the surface pressure. Our calculations show that the surface pressure is a controlling parameter in the photochemical stability and oxygen buildup of CO2-dominated atmospheres. The mixing ratio of O2 monotonically decreases as the surface pressure increases at very high outgassing rates, whereas it increases as the surface pressure increases at lower-than-Earth outgassing rates. Abiotic O2 can only build up to the detectable level, defined as 10−3 in volume mixing ratio, in 10-bar atmospheres with the Venus-like volcanic activity rate and the reduced outgassing rate of H2 due to the high surface pressure. Our results support the search for biological activities and habitability via atmospheric O2 on terrestrial planets in the habitable zone of Sun-like stars.

Export citation and abstract BibTeX RIS

1. Introduction

Recent discoveries of exoplanets (i.e., planets orbiting stars other than the Sun) have consistently sparked new research interest. Whether or not these planets are habitable and under what conditions allow them to be habitable has been driving exoplanet research. Today, scientists rely on remote sensing to detect and study exoplanets. The Kepler mission has confirmed over 2000 exoplanets; more than 20 of them are less than twice Earth-size and receive the right amount of stellar irradiation that would supports a liquid water ocean on their surface (based on the NASA Exoplanet Archive). If rocky, they are candidates for habitable exoplanets. The Transiting Exoplanet Survey Satellite (TESS) has been launched in 2018, and will cover an area of sky 400 time larger than Kepler and find many more potentially habitable worlds (Sullivan et al. 2015). The James Webb Space Telescope will have the capability to perform infrared spectroscopy of exoplanet atmospheres, including some of the potentially habitable worlds found by ground-based transit surveys (Dittmann et al. 2017; Gillon et al. 2017) and TESS (Schwieterman et al. 2016). Therefore, the search for signatures of habitability and bio-activities from exoplanet spectra is eminent.

Oxygen is considered a primary biosignature gas, thanks to its predominantly biogenic origins on Earth and its spectral features in the near-infrared (e.g., Lovelock 1965; Hitchcock & Lovelock 1967; Sagan et al. 1993). Abiotic processes such as photodissociation of H2O and CO2 can also produce oxygen, and if accumulated in the atmosphere, the abiotic oxygen would impede the use of oxygen as a biosignature gas (see Meadows et al. 2018, and references therein). In the case of an N2-dominated atmosphere, it has been shown that if the partial pressure of N2 is too small, water vapor cannot be trapped in the troposphere and the upper part of the atmosphere would be wet (Wordsworth & Pierrehumbert 2014). This leads to enhanced photodissociation of water and buildup of abiotic oxygen. In the case of a CO2-dominated atmosphere, early investigations suggest a massive buildup of oxygen (Selsis et al. 2002); the result was later called into question based on the redox balance in the atmosphere (Segura et al. 2007). Recent models that explicitly balance the redox budget of the atmosphere suggest that abiotic oxygen can sometimes accumulate to a detectable level, if the volcanic emission rate is very low (Hu et al. 2012; Domagal-Goldman et al. 2014) or the hydrogen content of the atmosphere is low (Gao et al. 2015). If the planet is around an M dwarf star, the accumulation of O2 will be much larger. Even in the case of volcanic emission rates as high as Earth, abiotic O2 can still accumulate to the detectable level (Tian et al. 2014; Harman et al. 2015). In this work, we define the detectable level as ∼10−3 in the volume mixing ratio, as suggested by evaluating the direct detection of the O2 A band in space (Des Marais et al. 2002; Segura et al. 2003). One may also consider a potential false-positive level to be 10−3 the present atmospheric level or 2 × 10−4, the oxygen level of Earth during most of the Proterozoic Eon with oxygenic photosynthesis (Planavsky et al. 2014). This level of oxygen is not detectable via direct imaging in the near-infrared, but may be detectable in the far future via O3 features in the thermal infrared (Léger et al. 1993; Segura et al. 2003) and ultraviolet wavelengths (Domagal-Goldman et al. 2014).

Previous work in analyzing abiotic accumulation of oxygen in CO2-dominated atmospheres assume the surface pressure to be ∼1 bar. In reality, there is no reason to assume a priori the size of a habitable exoplanet's atmosphere. First, small exoplanets have diverse compositions. For example, the TRAPPIST-1 planets, with radii of approximately Earth's radius, have very different bulk densities (Grimm et al. 2018) and therefore, some of the planets may have massive atmospheres and some may not. Second, the very concept of the habitable zone implies that the mass of atmospheric CO2 increases as the stellar irradiation decreases to maintain the conditions for liquid water oceans at the surface (Kasting et al. 1993; Bean et al. 2017). Atmospheric circulation models have shown that TRAPPIST-1 e would be habitable if it has a 1-bar CO2-dominated atmosphere, and TRAPPIST-1 f would be habitable if it has a 2-bar CO2 atmosphere (Wolf 2017; Turbet et al. 2018). In all, the size of a habitable exoplanet's atmosphere is unknown and must not be simply assumed as 1 bar.

Here we study abiotic production and accumulation of oxygen in CO2-dominated atmospheres with varied surface pressures, using a photochemistry model. Our goal is to explore the effect that changing the surface pressure can have upon the abundance of O2 in the atmosphere. By investigating the abiotic production of oxygen in non-1 bar atmospheres, we further clarify the conditions that the false-positive scenarios would occur. This paper is organized as follows. Section 2 describes the photochemistry model and the specifics of the simulated atmospheres. In Section 3, we present model results with the surface pressure ranging from 0.1 to 10 bars. We discuss their implications in Section 4 and conclude in Section 5.

2. Methods

2.1. Photochemistry Model

The photochemistry model we utilize has previously been used to study abiotic oxygen in reducing, weakly oxidizing and highly oxidizing atmospheres (Hu et al. 2012, 2013), as well as biosignatures in super-Earths with H2-rich atmospheres (Seager et al. 2013). Using the known atmospheric pressure-temperature profile and background composition, the code is able to accurately predict the amount of trace gases in both Earth's and Mars' atmospheres. The code finds steady-state solutions of species at each altitude of the model atmosphere. Tracing the kinetics of hundreds of chemical reactions, the model can compute the concentrations of 111 molecules and aerosols. A complete list of molecules and reactions can be found in Hu et al. (2012). The model uses a self-adjusting timestep to track the progress of balancing products and loss from all chemical and photochemical reactions toward reaching a steady state. As the code gets closer to convergence, the timestep increases. The code is considered as converged once the timestep becomes very large, e.g., 1017 s.

2.2. CO2-dominated Atmospheres

Hu et al. (2012) present benchmark cases with 1-bar atmospheres on terrestrial exoplanets ranging from H2-dominated to CO2-dominated. Here we focus on the CO2-dominated atmosphere scenarios and expand to include surface pressures ranging from 0.1 to 10 bars. Following the benchmark case of Hu et al. (2012), we consider the background composition of 90% CO2 and 10% N2, and include 57 C-, H-, O-, N-, and S-bearing species in the model. We focus on the planets around Sun-like stars in this work; the stability of non-1-bar CO2 atmospheres around M stars will be addressed in a separate paper. We do not include the potential effects of lightning (Hodosán et al. 2016; Rimmer & Helling 2016; Wong et al. 2017) in this study. Specific parameters for our model atmospheres are tabulated in Table 1.

Table 1.  Parameters for Modeled Terrestrial Exoplanets

Parameters CO2-dominated Atmosphere
Main Component 90% CO2, 10% N2
Mean Molecular Mass 42.4
Surface Pressure 0.1 bar 0.3 bar 1 bar 3 bars 10 bars
Planetary Parameters
Stellar Type G2V
Mass     1 M    
Radius     1 R    
Semimajor Axis 1.1 au 1.2 au 1.3 au 1.5 au 2.1 au
Temperature-pressure Profile
Surface Temperature     288 K    
Tropopause Altitude     9.77 km    
Temperature Above Tropopause     175 K    
Top-of-atmosphere (0.1 Pa) Altitude 43 km 47 km 52 km 55 km 59 km
Water and Rainouta,b
Liquid Water Ocean     Yes    
Water Vapor Boundary Condition     f (H2O) = 0.01    
Rainout rate     Earth-like    
Gas Emissionc (molecule cm−2 s−1) Very High Emission High Emission Low Emission Very Low Emission
CO × 1010 1.5 × 109 2.0 × 108 2.5 × 107
H2 × 1010 d 1.5 × 109 2.0 × 108 2.5 × 107
SO2 × 1010 d 1.5 × 109 2.0 × 108 2.5 × 107
H2S × 109 1.5 × 108 2.0 × 107 2.5 × 106

Notes.

aThe rainout rate of the nonsoluble species CO, CH4, H2, O2, and C2H6 are generally zero. bThe deposition velocities of gases follow the ones used in Hu et al. (2012). cThe volcanic gas emission rates are assigned as lower boundary conditions in each scenario of emission. H2O and CO2 are considered to be abundant in the system. dAssumed to be 3 × 109 molecule cm−2 s−1 in the special case where we include the effect of surface pressure on the outgassing speciation.

Download table as:  ASCIITypeset image

Temperature-Pressure Profile. To model habitable exoplanets, we adopt a surface temperature of 288 K. The temperature profile is then set to follow an appropriate adiabatic lapse rate (i.e., the convective layer) until 175 K and is assumed to be constant above (i.e., the radiative layer). We simulate the atmospheres up to an altitude that corresponds to 0.1 Pa. The temperature profile is consistent with significant greenhouse effects in the convective layer and with no additional heating above the convective layer for habitable exoplanets. We caution that the latter point may not be entirely valid because appreciable amounts of ozone are produced in some of our scenarios, which can lead to heating above the convective layer.

Semimajor Axis. To preserve a surface temperature of 288 K as the surface pressure changes, we adjust the orbital distance. We assume a circular orbit and adjust the semimajor axis by balancing the incoming stellar irradiation and the outgoing thermal irradiation for full heat redistribution. The semimajor axis (a) is estimated by

Equation (1)

where L is the stellar luminosity, AB is the Bond albedo, and I is the thermal emission irradiation flux. We assume the terrestrial value for AB and use the precalculated formula for I as a function of the CO2 partial pressure and the surface temperature (Williams & Kasting 1997). The range of surface pressure we consider is within the range of validity of the formula.

2.3. Boundary Conditions

Outgassing. We simulate scenarios of various surface gas emission rate to explore its effect on the atmospheric abundance of abiotic oxygen. The gas emission rate (ϕ) is assumed to be dominated by volcanic outgassing, and is parameterized by

Equation (2)

where V is the magma production rate and φ is the volatile content released by unit volume of magma. For the magma production rate, we explore a wide range. We consider the volcanic production rates of the present-day Earth (30 km3 yr−1; e.g., Gaillard & Scaillet 2014) and the present-day Venus (0.5–3 km3 yr−1; Gillmann & Tackley 2014), and they are defined to be the high emission and the very low emission scenarios, respectively. We use the lower end of estimate of Gillmann & Tackley (2014) to represent an endmember scenario of very low emission. For completeness, we also consider a low emission scenario where the volcanic production rate is the geometric mean of the high emission and very low emission scenarios (3.9 km3 yr−1), as well as a very high emission scenario to have a volcanic production rate ∼20 times that of the present-day Earth (600 km3 yr−1). Such a high rate may be found in younger-than-Earth planets or planets that receive large tidal dissipation.

The volcanic production rate of 0.5 km3 yr−1 can be converted to 4.3 × 104 kg s−1 for a density of 2700 kg m−3 (typical for mid-ocean ridge basalts). For a volcanic gas content of approximately 1000 ppm in mass and a mean molecular weight of 20 (e.g., Gaillard & Scaillet 2014), this corresponds to 2.1 × 103 mol s−1. Averaging over an Earth-radius planet, this is ∼2.5 × 108 molecules cm−2 s−1.

For speciation, we follow the planetary magmatic degassing calculations of Gaillard & Scaillet (2014). Consider typical mid-ocean ridge basalts degassing at a surface pressure ranging from 0.1 to 3 bars: H2, CO, and SO2 have mole fractions of ∼0.1 in the volcanic gas (Figure 4 of Gaillard & Scaillet 2014), and at 10 bars, the mole fractions of H2 and SO2 reduce to ∼0.01. The mole fractions of the elemental sulfur and H2S are ∼0.01 for all surface pressures. In this study, to isolate the effects of changing surface pressure, we adopt in the standard model a mole fraction of 0.1 for H2, CO, and SO2 and 0.01 for H2S. This is applied to all surface pressures. We additionally study a special case for the 10 bar surface pressure atmosphere, where we take into account less degassing of H2 and SO2 under high surface pressures by assuming a mole fraction of 0.1 for CO, and 0.01 for H2, SO2, and H2S. According to our estimate, the emission rate of H2 of an abiotic Earth is 1.5 × 109 molecules cm−2 s−1, consistent with Sleep & Bird (2007). The gas emission rates we model are tabulated in Table 1.

In addition to the gas emission, we include diffusion-limited escape for H and H2 at the top of the atmosphere in the same way as Hu et al. (2012).

Deposition and Redox Balance. We include dry deposition and rainout of species from the atmosphere in the same way as Hu et al. (2012). In particular, the deposition velocity of CO to the ocean is assumed to be 10−8 cm s−1, and that of O2 is assumed to be zero. This assumption considers that CO is slowly converted to acetate and precipitate in the ocean (Kharecha et al. 2005), and ignores various potential sinks for O2. Harman et al. (2015) considered recombination of CO and O2 in hydrothermal systems, oxidation of formate in solution, and oxidation of ferrous iron as the potential sinks of O2 in the ocean, and suggested that these sinks should be minor if using the Earth's process rates as a guide. We have tested the model by changing the deposition velocity of O2 to 10−8 cm s−1, the same as that of CO. The resulting mixing ratio profiles are not distinguishable with the standard assumption. If rapid processes exist in the ocean to directly recombine CO and O2, their deposition velocities would be on the order of 10−4 cm s−1 (Domagal-Goldman et al. 2014; Harman et al. 2015), and this would greatly reduce the amount of CO and O2 in the atmosphere in the steady state. Since there is no such processes known on Earth, we do not include them in the model.

We enforce redox balance in the atmosphere and in the ocean for all simulated scenarios. The redox balance says that the total redox influx to system (i.e., surface emission) should be balanced by the total redox outflux from the system, otherwise the system is being oxidized or reduced. The redox balance is equivalent to the conservation of the total number of electrons in the system. When applied to the atmosphere, the balance is strictly enforced by the mass balance and the convergence of the photochemical model itself. Specifically, we solve all molecules in the mass conservation equation, without assuming any fast species. It follows that the imbalance between the surface emission and the escape of hydrogen, if any, does not cause the redox of the atmosphere to change over time (it must be balanced by dry deposition and rainout), or,

Equation (3)

where Φ is the flux of net reducing species into the atmosphere, following the definition in Hu et al. (2012).

Recently, it has been realized that Φ(Deposition) may not be arbitrary in realistic planetary scenarios, as it represents a net transfer of reducing or oxidizing species from the atmosphere to the ocean (Domagal-Goldman et al. 2014; Harman et al. 2015). The key idea is that there shall be no net reducing species into the ocean on a planet without life. This is based on an analog to Earth as a terrestrial planet, where the only ways that the ocean can remove reducing species are burial of organic matter and burial of sulfide, both involving life (e.g., Harman et al. 2015). If this is universally applicable to exoplanets, we must impose

Equation (4)

Moreover, if Φ(Deposition) > 0, net oxidizing species are deposited into the ocean. While this is possible geologically (e.g., via burial of magnetite), this would reduce the abundance of O2 in the atmosphere. To simulate the limiting case for oxygen buildup in the atmosphere, we impose

Equation (5)

Equation (5) is essentially equivalent to Φ(Outgassing) + Φ(Escape) = 0, and is not automatically enforced by the photochemical model. We enforce this condition the same way as Harman et al. (2015), in that we include a return flux of hydrogen from the ocean to the atmosphere, if finding ${\rm{\Phi }}(\mathrm{Deposition})\ne 0$ for a converged solution. We then relaunch the simulation and repeat the process until the condition in Equation (5) is sufficiently satisfied. In practice we require an imbalance of the global redox budget to be no larger than 1% of the outgassing redox flux (Table 2). This way, our converged models satisfy redox balance for both the atmosphere and the ocean.

Table 2.  Redox Fluxes (Equivalent H cm−2 s−1) and Column Depths (cm−2) of Modeled Atmospheres

Scenarios Very High Emission Very Low Emission
  0.1 bar 10 bar 0.1 bar 10 bar
Escape
H −7.38E+8 −2.95E+8 −7.92E+6 −7.24E+5
H2 −1.37E+11 −1.38E+11 −1.08E+8 −1.14E+8
Total −1.38E+11 −1.38E+11 −1.16E+8 −1.15E+8
Outgassing
H2 6.0E+10 6.0E+10 5.0E+7 5.0E+7
CO 6.0E+10 6.0E+10 5.0E+7 5.0E+7
H2S 1.8E+10 1.80E+10 1.5E+7 1.5E+7
Total 1.38E+11 1.38E+11 1.15E+8 1.15E+8
Dry and Wet Deposition
O3 3.25E+10
HO2 4.96E+8 1.28E+7
H2O2 4.27E+9 7.04E+8
CO −4.17E+10 −6.12E+10 −2.40E+9 −3.52E+10
CH2O −9.25E+7 −2.63E+8
H2S −7.39E+9 −1.73E+10 −5.98E+6 −1.44E+7
H2SO4 6.84E+7 9.33E+7 1.34E+6
S8 −1.92E+10 2.17E+6
Total −6.83E+10 −7.84E+10 2.10E+9 −2.03E+9
Return H2 Flux
H2 6.83E+10 7.84E+10 −2.09E+9 2.03E+9
Column Depth
O2 6.4 × 1018 3.0 × 1018 2.5 × 1019 4.9 × 1023
O3 2.0 × 1014 8.0 × 1014 3.1 × 1015 9.4 × 1018

Download table as:  ASCIITypeset image

3. Results

We find that the trend of photochemically produced O2 in the CO2-dominated atmosphere is different between the high and the low emission scenarios: when the emission rate is high (e.g., 20× Earth's volcanic activity level, a.k.a. the very high emission scenario in this study), the mixing ratio of O2 decreases as a function of the surface pressure, and when the emission rate is low (e.g., Venus's volcanic activity level, a.k.a. the very low emission scenarios in this study), the mixing ratio of O2 increases as a function of the surface pressure (Figure 1). In between the two endmembers, the Earth-like emission rates (a.k.a. the low and high emission scenarios in this study) have the O2 mixing ratio peaks at ∼0.3 bar to 10−4 (Figure 1). In addition, the column depth of O3 photochemically produced in the atmosphere increases with the surface pressure, and also is higher for lower surface emission rates (Figure 2).

Figure 1.

Figure 1. Mixing ratios of the key species in the CO2-dominated atmosphere as a function of the surface pressure. Four surface gas emission scenarios defined in Table 1 are shown in four different colors. The black dashed line in the O2 panel shows the level above which the abiotic O2 would be detectable (10−3), and the black solid line shows the oxygen level of Proterozoic Earth (2 × 10−4). Each point is summarized from a converged simulation of the full photochemical model. The abundance of O2 increases with the surface pressure at very low emission rates, whereas it decreases with the surface pressure at very high emission rates.

Standard image High-resolution image
Figure 2.

Figure 2. Column depth of O3 as a function of the surface pressure and the surface emission scenarios. The black dashed line shows the column depth of O3 of present-day Earth for comparison.

Standard image High-resolution image

The redox balance of the atmosphere and the ocean is maintained for all scenarios. Table 2 shows the redox balance for the four endmember scenarios of the very high and very low emission rate and the 0.1 bar and 10 bar surface pressure. We see that for all models the escape flux balances the outgassing flux. Figure 1 shows that the mixing ratio of H2 depends on the outgassing flux but not on the surface pressure, which is a direct consequence of maintaining the balance between hydrogen escape and outgassing of reducing species. Also, the net deposition flux is balanced by the return H2 flux. In most (but not all) cases there is a net deposition flux of reducing species, which is balanced by a positive H2 flux from the ocean to the atmosphere.

Figure 3 provides a detail look into the mixing ratio profiles of the key species and shows the difference between the endmember scenarios of the very high emission rate and the very low emission rate. With the very high emission rate, the mixing ratio of O2 drops to extremely small values near the surface for surface pressures ranging from 0.1 to 10 bars. The mixing ratio of CO is >1% and is larger for lower surface pressures. With the very low emission rate, however, O2 can have substantial mixing ratios at the surface when the atmosphere is larger than ∼0.3 bars; the mixing ratio of this O2 generally increases for larger atmospheres, to about 6 × 10−4 (Figure 1). Due to this O2 accumulation, an ozone layer is formed, with the peak ozone mixing ratio between 10−6 and 10−5, on the same orders of magnitude as the present-day Earth's ozone layer.

Figure 3.

Figure 3. Comparison between the very high emission scenario and the very low emission scenario in terms of the mixing ratio profiles of the key species. Solid lines are used for both 0.1 bar and 10 bar cases, and dashed, dotted, and dashed–dotted lines correspond to 0.3, 1, and 3 bar cases, respectively. At the very high emission rate, no O2 is accumulated at the surface regardless of the surface pressure, whereas at the very low emission rate, O2 can accumulate when the surface pressure is >0.1 bar.

Standard image High-resolution image

Pressure-dependent degassing of magma can impact the oxygen buildup when the overall emission rates are small. Figure 4 compares the standard model, where the same volcanic gas composition is assumed for all surface pressures, and the special case, where 10-fold less H2 and SO2 in volcanic gas is assumed for the 10 bar surface pressure atmosphere, according to the calculations of Gaillard & Scaillet (2014) for degassing under the surface pressure. There is no significant difference in the abundance of atmospheric O2 for the very high emission scenario, but the O2 abundance increases by 80% (to 1.1 × 10−3) when the pressure dependency of degassing is taken into account. The main cause is a further reduction of the H2 emission rate when the surface pressure is high. The atmospheres with less H2 emission have less H2 and more CO and O2.

Figure 4.

Figure 4. Comparison of two speciation models for the 10 bar surface pressure atmosphere. The solid lines correspond to the outgassing rates tabulated in Table 1, and the dashed lines include 10-fold less H2 and SO2 due to the effect of surface pressure to the outgassing speciation (Gaillard & Scaillet 2014). No significant difference is found between the two speciation models at very high emission, but the abundance of O2 increases by 80% for the 10-fold less H2 and SO2 at very low emission.

Standard image High-resolution image

Why are the behaviors of the high emission scenarios and the low emission scenarios so different when changing the surface pressure? Figure 5 shows the reaction rate profiles of key chemical reactions that lead to the combination of CO and O. It is well known that the direct combination

Equation (R1)

requires a third molecule and is only efficient when the ambient number density is large. The combination can alternatively proceed with

Equation (R2)

For all modeled scenarios, we find that R1 dominates the lower atmosphere and R2 dominates the middle and the upper parts of the atmosphere. These two combined can account for the photodissociation of CO2 (Figure 5). For R2 to work as a catalytic cycle, H has to be converted to OH. The following cycle is known to operate in the atmosphere of Mars (Nair et al. 1994):

Equation (R3)

Equation (R4)

The net result of R3 and R4 is H + O → OH, and this completes the catalytic cycle of R2 that combines CO and O. Figure 5 shows that in the very low emission scenario, this cycle indeed operates and dominates the combination in the middle atmosphere. However, in the very high emission scenario, we find that the rate of R3 is substantially smaller than the rate of R4 in the middle atmosphere. Instead, the following cycle produces the HO2 needed in R4:

Equation (R5)

Equation (R6)

The net result of R5 and R6 is exactly the same as R3. This cycle replaces R3 and dominates the middle atmosphere in the very high emission scenario.

Figure 5.

Figure 5. Key chemical reactions that contribute to the combination of CO and O. For all cases the direct combination of CO + O + M dominates the lower atmosphere and the reaction OH + CO → CO2 + H dominates the middle atmosphere. At the very high emission rate, the cycle of CHO (Reactions R5 and R6) dominates the production of HO2, whereas at the very low emission rate, direct combination of H and O2 (Reaction R3) dominates the production of HO2.

Standard image High-resolution image

Therefore, the starting reaction for the catalytic combination of CO and O is different between the very high emission scenario and the low and very low emission scenarios. For the former it is H + O2 (R3) and the latter it is H + CO (R5). Moreover, the pressure dependency of the reaction rate constants of R3 and R5 is different. The reaction rate constant of R3 has a high-pressure limit of 7.5 × 10−11(T/300)0.21 cm3 s−1, where T is the temperature (Burkholder et al. 2015). The reaction rate constant of R5 does not have an established high-pressure limit and is 5.3 × 10−34exp(−370/T)N cm6 s−1, where N is the ambient number density (Baulch et al. 1994). The rate constant of R5 is 3.7 × 10−15 cm3 s−1 at 1 bar and 300 K, and is 10 times higher at 10 bar. Comparing the rate constants and the O2 and CO mixing ratios, we find that with the very high emission rate, the mixing ratio of CO is more than four orders of magnitude higher than that of O2 (Figure 1), and so R5 dominates over R3. As the cycle starting with R5 becomes more efficient at higher pressures, the mixing ratio of O2 at the steady state decreases. With the very low emission rates, R3 dominates over R5, and its rate constant reaches the high-pressure limit. The catalytic cycle cannot become more efficient as the pressure increases, leading to an increase of the steady-state O2.

4. Discussion

The simulations presented here qualitatively confirm the results of Hu et al. (2012) at the surface pressure of 1 bar and extend to the surface pressures ranging from 0.1 to 10 bars. Compared to Hu et al. (2012), we have in this work changed the main reducing species in outgassing from CH4 to CO according to the subaerial magma degassing model of Gaillard & Scaillet (2014), and enforced the redox balance of the ocean by including a H2 return flux controlled by the redox balance of the deposition (Harman et al. 2015). We have also worked with a more realistic range of volcanic activities, in that the zero emission case in Hu et al. (2012) is replaced by the very low emission scenario in this work corresponding to Venus's volcanic rate. With these updates, the results continue to indicate accumulation of O2 near the surface of the planet. The cases where the O2 mixing ratio is indefinitely small at the surface are the very high emission scenarios (20 times higher than present-day Earth's volcanic activity) and the 0.1 bar surface pressure scenarios. All other cases have O2 near the surface to various abundances, and also have O3 peaks in the middle atmosphere. This result is different from Harman et al. (2015), where the authors suggest the O2 accumulation for terrestrial planets of Sun-like stars is completely prevented by enforcing the redox balance of the ocean. The difference may be due to the reaction networks being used, and also the criteria of convergence in the model.

The steady-state mixing ratio of O2 is below 10−3, the presumed level for O2 to be detectable via direct imaging, for all model scenarios. The only exception is the 10-bar atmosphere under the very low volcanic activity level and with the further reduced emission rate of H2 due to the high surface pressure (i.e., the dashed lines in Figure 4). Therefore, our model results, while different from Harman et al. (2015) in terms of the mixing ratio profiles of O2, support their generic conclusion that O2 is unlikely to have a photochemical false positive if found in abundance (>10−3) on water-rich and terrestrial planets in the habitable zone of a Sun-like star. This finding supports the use of O2 as a key indicator for potentially habitable worlds by future direct imaging space missions.

Our models further indicate that the potential for photochemical oxygen to accumulate to nearly the detectable level is particularly large when the atmosphere has a large amount of CO2 and when the volcanic activity of the planet is low. The planets close to the outer edge of the habitable zone are expected to have large CO2 partial pressures in the atmosphere (Kasting et al. 1993). This risk can be mitigated by detecting signatures of volcanic activities. The signatures include sulfur-bearing species (Kaltenegger & Sasselov 2010; Hu et al. 2013) and methane (Domagal-Goldman et al. 2014). Finally, we note that while the steady-state mixing ratios of O2 in our fiducial cases are not particularly larger than 10−3, they are in some cases larger than the oxygen level of Earth in most of the Proterozoic Eon after the Great Oxidation Event (Planavsky et al. 2014). Detail photochemistry models are thus necessary to interpret future detections of O2.

In addition to O2, our simulations suggest substantial photochemical production of O3 over a wide range of surface pressures and emission rates. Domagal-Goldman et al. (2014) showed that ozone would be detectable via its Hartley band at 0.25 μm for a column depth of 1015 ∼ 1018 cm−2. The column depth of O3 we find from the very low to the high emission scenarios spreads in this range (Figure 2). Ozone alone cannot be considered as a biosignature gas due to this false positive, in agreement with (Domagal-Goldman et al. 2014). The ozone feature at 0.25 μm is particularly sensitive to the atmospheric scenarios determined by the emission rates and the surface pressure. A measurement in this band would thus greatly help to pinpoint the atmospheric scenario of a terrestrial exoplanet.

Finally, the trend we find for the low and very low emission scenarios is consistent with the work of Zahnle et al. (2008), in the context of the Martian atmosphere. Despite the difference in terms of the atmospheric water abundance, we essentially agree with Zahnle et al. (2008) that a larger CO2 atmosphere would be more photochemically unstable, if the surface emission rates are low. We discover the reverse trend under the high emission rates. The high atmospheric CO versus O2 ratio leads to the catalytic cycle initiated by H + CO (Section 3) and further stabilizes the CO2 atmosphere at high surface pressures. If this trend also applies to dryer conditions, an atmosphere of early Mars more massive than the present might have been stabilized by strong volcanic outgassing.

5. Conclusion

We have used a one-dimensional photochemical model to simulate the composition of CO2-dominated atmospheres on terrestrial exoplanets in the habitable zone of a Sun-like star, for the surface pressures ranging from 0.1 to 10 bars and the emission rates corresponding to volcanic activities from Venus-like to 20 times higher than Earth-like. Our models maintain the redox balance of both the atmosphere and the ocean. We find that the emission rates control how the mixing ratio of photochemically produced O2 changes with the surface pressure. The mixing ratio of O2 increases with the surface pressure when the emission rates are very low, consistent with previous studies. However, driven by a catalytic cycle initiated by the combination reaction between H and CO, the mixing ratio of O2 decreases with the surface pressure when the emission rates are high. We have also studied the effect of the surface pressure on the speciation of magma degassing and the composition of volcanic outgassing. For the very low volcanic activity and 10-bar atmosphere, this effect almost doubles the steady-state mixing ratio of O2.

To search for potentially habitable exoplanets, our models support the use of O2 detectable via its A band at 0.76 μm as a key indicator for oxygenic photosynthesis on terrestrial planets in the habitable zone of Sun-like stars. The maximum amount of photochemical O2 we find is 1.1 × 10−3 in terms of volumetric mixing ratio. We could define the amount of O2 above this level to be abundant, and such a definition is natural and commensurate with the strength of the A band and reasonable prospect for detection capabilities in the next decade or two. Abundant oxygen, with contextual information including the rocky nature of the planet and the existence of water vapor in the atmosphere, is then probably the signature we should aim for. We may miss some habitable planets in this way, since Earth managed to maintain its oxygen level below 10−3 for much of the time after the rise of oxygenic photosynthesis (Reinhard et al. 2017). We likely have to accept this false negative, as the models presented here show that oxygen lower than 10−3 has a direct photochemical false positive. This photochemical false positive may however be mitigated by also ruling out abundant CO in the atmosphere (e.g., Schwieterman et al. 2016), because all our models have large mixing ratios of CO. In all, with the volcanic activity and the pressure dependency of the reaction rate constants quite universal for planets, our numerical experiments provide a useful baseline to understand under what conditions oxygen on potentially habitable terrestrial exoplanets can be regarded as a biosignature.

The research was carried out at the Jet Propulsion Laboratory, California Institute of Technology, under a contract with the National Aeronautics and Space Administration. T.J. thanks the support of the US National Science Foundation's (NSF) Division Of Undergraduate Education (DUE), under grant No.1457943.

Please wait… references are loading.
10.3847/1538-4357/aae2bb