This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.

The following article is Open access

The Importance of the Upper Atmosphere to CO/O2 Runaway on Habitable Planets Orbiting Low-mass Stars

, , , , and

Published 2023 November 21 © 2023. The Author(s). Published by the American Astronomical Society.
, , Citation Sukrit Ranjan et al 2023 ApJL 958 L15 DOI 10.3847/2041-8213/ad037c

Download Article PDF
DownloadArticle ePub

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

2041-8205/958/1/L15

Abstract

Efforts to spectrally characterize the atmospheric compositions of temperate terrestrial exoplanets orbiting M dwarf stars with JWST are now underway. Key molecular targets of such searches include O2 and CO, which are potential indicators of life. Recently, it was proposed that CO2 photolysis generates abundant (≳0.1 bar) abiotic O2 and CO in the atmospheres of habitable M dwarf planets with CO2-rich atmospheres, constituting a strong false positive for O2 as a biosignature and further complicating efforts to use CO as a diagnostic of surface biology. Importantly, this implied that TRAPPIST-1e and TRAPPIST-1f, now under observation with JWST, would abiotically accumulate abundant O2 and CO, if habitable. Here, we use a multi-model approach to reexamine photochemical O2 and CO accumulation on planets orbiting M dwarf stars. We show that photochemical O2 remains a trace gas on habitable CO2-rich M dwarf planets, with earlier predictions of abundant O2 and CO due to an atmospheric model top that was too low to accurately resolve the unusually high CO2 photolysis peak on such worlds. Our work strengthens the case for O2 as a biosignature gas, and affirms the importance of CO as a diagnostic of photochemical O2 production. However, observationally relevant false-positive potential remains, especially for O2's photochemical product O3, and further work is required to confidently understand O2 and O3 as biosignature gases on M dwarf planets.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

The launch of JWST has begun a new era in the characterization of exoplanet atmospheres. JWST has already revolutionized the study of gas giant exoplanet atmospheres, detecting novel chemical species and using them to infer operant chemical processes (Tsai et al. 2023). In the future, JWST will similarly seek to characterize the atmospheres of smaller exoplanets, including potentially habitable temperate terrestrial worlds, and such observations have already begun (Lafreniere 2017; Lewis et al. 2017; Lustig-Yaeger et al. 2019; Lim et al. 2021; Stevenson et al. 2021). Such observations may constrain surface processes such as volcanism, geochemical cycling, and the presence of life (Kaltenegger & Sasselov 2010; Kaltenegger et al. 2010; Misra et al. 2015; Krissansen-Totton et al. 2018; Fauchez et al. 2020; Lehmer et al. 2020; Zhan et al. 2021; Ranjan et al. 2022).

Particularly relevant in the quest to characterize temperate terrestrial exoplanets are planets with anoxic, CO2-rich atmospheres orbiting M dwarf stars. CO2-rich atmospheres are expected to be ubiquitous on terrestrial exoplanets, as they are in the solar system, thanks to robust outgassing of CO2 by basaltic magmatism (Gaillard & Scaillet 2014; Catling & Kasting 2017). Such high mean molecular mass secondary atmospheres are only accessible to atmospheric characterization with JWST for planets orbiting M dwarf stars. Planets with CO2-rich atmospheres orbiting M dwarf stars are therefore highly observationally relevant, and it is critical to understand their atmospheric photochemistry to accurately interpret anticipated observations of their atmospheres (Shields et al. 2016; Catling et al. 2018).

Atmospheric photochemical models of planets with CO2-rich atmospheres orbiting M dwarf stars are historically discrepant, with observationally relevant implications for the abundance of spectroscopically active trace gases proposed as probes of surface processes. Disagreement has been particularly intense about photochemical CO and O2 abundance in such atmospheres. While all models predicted the accumulation of photochemical CO and O2 on CO2-rich planets orbiting M dwarf stars, the specific predicted concentrations of CO and O2 varied by many orders of magnitudes between models run on identical planetary scenarios (Harman et al. 2015, 2018). This is problematic because atmospheric O2 and CO are proposed as remote probes of surface processes including the presence or absence of life (Meadows et al. 2018; Schwieterman et al. 2019; Wogan & Catling 2020). Uncertainty about the efficiency of photochemical generation of these gases neuters their value as probes of surface processes.

Recognizing the importance of this problem, extensive efforts have been invested to improve the understanding of photochemical CO and O2 on CO2-rich planets orbiting M dwarf stars. Initial efforts to simulate high-CO2 atmospheres on planets orbiting M dwarf stars resulted in a disagreement of up to 4 orders of magnitude in predicted pCO and pO2 (Domagal-Goldman et al. 2014; Tian et al. 2014; Harman et al. 2015). Harman et al. (2015) showed some of the inter-model disagreement was due to different lower boundary conditions. Harman et al. (2018) reconciled most of the remaining inter-model disagreement by showing that it was due to different assumptions regarding the photochemical effects of lightning, which generates NOX species in a CO2-N2 atmosphere. These NOX species can drive catalytic chemistry, which recombines CO and O2 to CO2. Upon including a lightning rate corresponding to modern Earth and its production of catalytic NOX species, Harman et al. (2018) found photochemical O2 to be low (pO2 < 2 × 10−4 bar, below the false-positive threshold) across all models and planetary scenarios they considered.

Most recently, the possibility of abundant photochemical CO and O2 was examined by Hu et al. (2020), who highlighted that previous work did not include the key NOX reservoir species HNO4 and N2O5. These species are relatively stable under M dwarf UV irradiation and serve to shuttle NOX into the ocean due to their high solubility, reducing the efficacy of the NOX-driven recombinative cycles. Hu et al. (2020) reproduced the results of Harman et al. (2018) for 0.05 bar CO2 when excluding HNO4 and N2O5, but found that the inclusion of HNO4 and N2O5 into the photochemical network led to the generation of abundant photochemical CO and O2, even assuming Earth-like lightning and its concomitant production of catalytic NOX species. This mechanism is minimally affected by updated CO2 and H2O near-UV (NUV) cross sections (Ranjan et al. 2020) because M dwarf NUV emission is very low, rendering the photochemical effect of the updated cross sections modest for planets orbiting low-mass M dwarf stars (W. Broussard et al. 2023, in preparation). Applying this finding to the TRAPPIST-1 system, Hu et al. (2020) found that if TRAPPIST-1e and TRAPPIST-1f had enough CO2 to be globally habitable, then they would necessarily accumulate abundant (≳0.1 bar, and possibly ≥1 bar) CO and O2, with high O3 as well. This atmospheric state (CO/O2 runaway) would constitute an observable false positive for O2 as a biosignature, would likely constitute a false positive for O3 as a biosignature, and would further complicate efforts to use CO as a diagnostic of surface biology. Importantly, this false positive would persist in comparative planetological approaches to O2 as a biosignature in the TRAPPIST-1 system, since abundant photochemical O2 would specifically co-occur with habitable conditions on the outer planets. With observations of the TRAPPIST-1 system already underway, it is critical to confirm the prospects for the photochemical accumulation of CO and O2 on the TRAPPIST-1 planets, to understand how to interpret potential detections of O2, O3, and CO.

Here, we explore further the possibility of substantial accumulation of CO and O2 on CO2-rich conventionally habitable planets orbiting M dwarf stars. By conventionally habitable, we refer to rocky planets with atmospheres and stable surface liquid water oceans, which are the prime targets for exoplanet biosignature search (Kopparapu et al. 2013; Kasting et al. 2014), as opposed to desiccated rocky planets, moist greenhouse rocky planets, Hycean worlds, or mini-Neptune aerial biospheres (Kasting 1988; Abe et al. 2011; Gao et al. 2015; Madhusudhan et al. 2021; Glidden et al. 2022). Our basic goal is to confirm whether it is possible for the photochemical decomposition of CO2 to CO and O2 to be favored in conventionally habitable planet atmospheres, even with a modern Earth-like lightning rate (Harman et al. 2018; Hu et al. 2020). We place specific emphasis on O2, which is strongly motivated by the solar system as a biosignature gas (Sagan et al. 1993), and on the TRAPPIST-1 system, now under observation by JWST. However, our findings are relevant to multiple gases and to planets orbiting M dwarf stars in general. We reproduce the CO/O2 runaway state in two independently developed photochemical models and determine its key controls. Our multi-model approach is critical because it enables us to be confident that our findings are not due to numerical or implementation errors of individual models. We describe our methods in Section 2, report and discuss our results in Sections 3 and 4, and summarize in Section 5. We provide additional background, supporting information, and methodological details in Appendices AE.

2. Methods

We employ two independent 1D photochemical models in our exploration of CO/O2 runaway: the MIT Exoplanet Atmospheric Chemistry Model (MEAC; Hu et al. 2012) and Atmos (Arney et al. 2016; Lincowski et al. 2018). These models implement conceptually similar chemistry, physics, and numerical schemes, but were developed independently and share no codebase. This means that results confirmed using both models are likely to be robust to implementation errors, though they may still suffer from common-mode errors in basic physicochemical understanding (Wen et al. 1989). This intercomparative approach has proved useful in past studies employing exoplanet photochemical models, which are extremely complex, often nonlinear, and feature numerous free parameters (Harman et al. 2015, 2018; Ranjan et al. 2020). In this work, we employ MEAC and Atmos in a highly focused investigation of CO/O2 runaway. A more exhaustive intercomparison engaging a much broader diversity of models in a much broader range of planetary scenarios is underway in the form of the Photochemical model Intercomparison for Exoplanets (PIE) project (PI: C. E. Harman) as part of the CUISINES initiative. 10

2.1. MEAC Model and Configuration.

We employ MEAC as updated by Ranjan et al. (2022), which is modified from the original Hu et al. (2012) version by elimination of errors and use of updated H2O cross sections (Ranjan et al. 2020; Hu 2021). Importantly, this model is similar to the version of MEAC employed by Hu et al. (2020), which motivated the present work. MEAC is a 1D photochemical model that calculates the steady-state vertical trace gas composition of a planetary atmosphere given temperature–pressure and eddy diffusion (vertical transport) profiles, stellar irradiation, chemical network, and chemical boundary conditions. MEAC encodes processes including photolysis (computed via delta two-stream approximation), eddy diffusion and molecular diffusion of H and H2, surface emission and wet and dry deposition of chemical species, diffusion-limited escape of H and H2, and formation and deposition of S8 and H2SO4 aerosols. MEAC does not account for the formation of organic haze, expected at elevated CH4/CO2 ratios (DeWitt et al. 2009; Arney et al. 2016); we do not explore this regime in our work. We assign a high deposition velocity of 1 × 10−5 cm s−1 for C2H6 to account for this omission in an ad hoc fashion following Hu et al. (2012). For our full CHOSN chemical network, we consider 86 species linked by 734 reactions, corresponding to the full chemistry of Hu et al. (2012) excluding the higher hydrocarbons. Appendix E further summarizes the MEAC model setup.

2.2. Atmos Model and Configuration

Atmos is a 1D photochemical-climate model first developed by Kasting et al. (1979) and recently modified by numerous works including Arney et al. (2016), Lincowski et al. (2018), Leung et al. (2022), Teal et al. (2022). For this work, we use the publicly available version of the code 11 with modifications as suggested by Ranjan et al. 2020. This code has 79 species with 397 reactions including 65 photolysis reactions. Atmos also incorporates eddy diffusion, wet and dry deposition, and sulfur aerosols to simulate terrestrial environments. While not explored in this work, the model has previously been used to simulate hazy Archean-like atmospheres and has an extensive history of use for other terrestrial exoplanet applications (e.g., Domagal-Goldman et al. 2011; Arney et al. 2016, 2018; Schwieterman et al. 2019; Leung et al. 2022; Teal et al. 2022).

2.3. Planetary Scenario

We apply our photochemical models to a planetary scenario closely corresponding to the CO2-dominated abiotic Earth-like planet benchmark scenario detailed in Hu et al. (2012). We assume the atmospheric structure from Hu et al. (2012) (T(z), P(z), Kz (z); Figure E2). To emphasize the enforcement of mass balance, we avoid fixed mixing ratio boundary conditions in favor of fixed surface flux and dry deposition velocity boundary conditions. We adopt surface emission of SO2, H2S, CH4, and H2 broadly consistent with Earth-like volcanism, surface production of CO and NO broadly consistent with Earth-like lightning, and generally low dry deposition velocities to simulate an abiotic planet (Hu et al. 2012; Harman et al. 2018; Hu & Diaz 2019) (Table 6). We set wet deposition of H2, O2, CO, CH4, C2H2, C2H4, C2H6, and NH3 to zero, to simulate saturation on a planet with abiotic oceans (Hu et al. 2012). For stellar irradiation, we adopt the TRAPPIST-1 model 1A spectrum of Peacock et al. (2019). Further details on the planet scenario are presented in Appendix E.

The high pCO2 = 0.9 bar in this scenario puts it well into the CO/O2 runaway regime identified by Hu et al. (2020). We generally avoid intermediate pCO2 corresponding to the runaway transition itself, where sensitivity to photochemical assumptions is enhanced (Ranjan et al. 2022). This choice ensures that our modeling reflects the fundamental cause of the CO/O2 runaway, as opposed to "red herrings", which have enhanced importance solely in the runaway transition regime, but do not control the overall runaway phenomenon.

2.4. Photochemical Model Deployment

We perform 17 targeted simulations with a combination of MEAC and Atmos to test model and parameter sensitivities and explore CO/O2 runaway (Table 1). This includes the stellar spectrum (including the Sun, TRAPPIST-1, and GJ 876), the partial pressure of CO2 (pCO2), the chemical network: the full CHOSN network as described for both models, then chemical networks without N2O5 and HNO4 (CHOSN - (N2O5+HNO4)), without nitrogen chemistry (CHOS), and without nitrogen and sulfur chemistry (CHO). We additionally test the impact of the shortwave UV cutoff in the stellar spectrum (λcrit), the maximum altitude of the model (zmax) and model grid resolution (Δz), and the magnitude of vertical transport (scaled by a factor Kscale). The results of these tests are summarized in the last two columns of Table 1 in which we show the surface-level CO and O2 partial pressures (pCO and pO2, respectively).

Table 1. Predicted pCO and pO2 for Our Baseline 0.9 Bar CO2, 0.1 Bar N2 Planetary Scenario for Different Model Assumptions

RunModelStarpCO2 Network λcrit zmax Δz Kscale pCOpO2
   (bar) (nm)(km)(km) (bar)(bar)
0MEACSun0.9CHOSN11054115 × 10−6 2 × 10−20
1MEACTRAPPIST-10.9CHOSN1105411 0.8 1
2MEACTRAPPIST-10.9CHOSN - (N2O5+HNO4)1105411 0.8 1
3MEACTRAPPIST-10.9CHOS1105411 0.8 1
4MEACTRAPPIST-10.9CHO1105411 0.8 1
5MEACTRAPPIST-10.9CHO110100110.071 × 10−5
6MEACTRAPPIST-10.9CHO12054110.073 × 10−4
7MEACTRAPPIST-10.9CHO120100110.071 × 10−5
8AtmosTRAPPIST-10.9CHOSN117.61000.510.0951.9 × 10−4
9AtmosTRAPPIST-10.9CHOSN117.6550.51 1.65 0.13
10MEACGJ 8760.05CHOSN11054110.062 × 10−3
11MEACGJ 8760.05CHOSN - (N2O5+HNO4)11054110.035 × 10−9
12MEACTRAPPIST-10.9CHO110540.110.087 × 10−4
13MEACTRAPPIST-10.9CHO110541103 0.262 × 10−4
14MEACTRAPPIST-10.9CHO110100110−3 5 × 10−8 2 × 10−19
15MEACTRAPPIST-10.9CHO110100510.017 × 10−11
16MEACTRAPPIST-10.9CHO1101000.210.042 × 10−5

Note. We have bolded instances of CO/O2 runaway, which we define on an ad hoc basis as pCO > 0.1 bar and pO2 > 0.1 bar. Assumptions regarding the details of the chemical scheme are not the main control on CO/O2 runaway. Rather, assumptions regarding the shortwave UV cutoff and especially regarding the model top are the main control on CO/O2 runaway, through their influence on the resolution of the CO2 photolysis peak. This finding is multi-model.

Download table as:  ASCIITypeset image

2.5. Synthetic Spectra

To illustrate the observational implications of our photochemical simulations, we simulate planetary transmission spectra. We simulate these synthetic spectra using the Planetary Spectrum Generator (PSG; Villanueva et al. 2018, 2022) by adapting the example presented at https://github.com/nasapsg/globes/blob/main/atmos.py. PSG is a widely used community tool for simulating exoplanet transmission spectra (e.g., Fauchez et al. 2020; Pidhorodetska et al. 2020; Suissa et al. 2020). We choose planetary and stellar parameters consistent with our photochemical simulations (i.e., 1 M, 1 R planet orbiting a TRAPPIST-1 like star). We include absorption due to H2O, CH4, C2H6, CO2, O2, O3, CO, H2CO, NO, NO2, SO2, N2O, and N2, as well as the effects of Rayleigh scattering, refraction, and all collision-induced absorption (CIA) included in PSG at a resolution of R = 500. PSG reports spectral fractional transit depths d(λ),

Equation (1)

where RP is the solid radius of the planet, R is the radius of the star, and zatm(λ) is the wavelength-dependent effective height of the atmosphere. d(λ) is specific to a given host star. To obtain a more general metric of the transmission spectrum that is not host star-specific and can be compared to other simulated transmission spectra, we also compute the effective height of the atmosphere zatm(λ) by solving Equation (1) for zatm(λ).

3. Results

3.1. Photochemistry

We find that the occurrence of CO/O2 runaway for conventionally habitable planets with high-CO2 atmospheres is primarily controlled by the resolution of the CO2 photolysis peak, via the choice of zmax and λcrit.

We find that the details of the photochemical network are not the primary control for CO/O2 runaway. Previously, it was reported that the inclusion of N2O5 and HNO4 into photochemical networks resulted in CO/O2 runaway for CO2-dominated atmospheres orbiting M dwarf stars because the reactive nitrogen, which could catalyze the recombination of CO and O2 back to CO2, was instead converted into HNO4 and N2O5 and sequestered into the ocean. Indeed, CO/O2 runaway is sensitive to the details of the chemical network at intermediate pCO2, corresponding to the onset of runaway (Table 1, Runs 8 and 9), as originally reported (Hu et al. 2020). In this intermediate atmospheric regime, sensitivity to photochemical details is enhanced because the dominant forcing on the atmosphere is changing (Ranjan et al. 2022). However, our simulations show that for CO2-rich atmospheres that are firmly in the runaway regime, the runaway persists when N2O5 and HNO4 are removed from the photochemical network (Table 1, Runs 1 and 2). Indeed, it persists even when nitrogen and sulfur chemistry are excised entirely from the network (Table 1, Runs 3 and 4). This demonstrates that the details of the photochemical network are not the fundamental driver of CO/O2 runaway.

Instead, CO/O2 runaway for high-CO2 atmospheres is driven by the choice of zmax and λcrit. Specifically, we found that either increasing zmax from 54 to 100 km or increasing λcrit from 110 to 120 nm was adequate to inhibit the atmosphere from becoming CO/O2 dominated, though concentrations of both gases remained high relative to the Sun-like star case (Table 1, Runs 0, 5, and 6). Both a low λcrit and a low zmax were required to induce a runaway. Critically, this finding is multi-model: by adjusting zmax from 100 to 55 km, we were able to induce CO/O2 runaway in the independently developed Atmos photochemical model 12 as well, despite its omission of HNO4 and N2O5 (Table 1, Runs 10 and 11). This confirms that λcrit and zmax, and not the details of the photochemical network, are the main drivers of model predictions of CO/O2 runaway.

To elucidate the mechanism by which λcrit and zmax drive CO/O2 runaway, we modeled atmospheres with pCO2 = 0.01–0.1 bar (pN2 = 0.99–0.9 bar). We explored both truncated atmospheres (model tops set at 0.34 μbar, as in the baseline 54 km case) and extended atmospheres (model tops set at 100 km; 4 nanobar for pCO2 = 0.1 bar). We chose this pCO2 range because it corresponded to the onset of runaway in the truncated atmospheres, enabling us to study the numerical drivers of this phenomenon, and because at higher pCO2, photochemical O2/O3 layers emerge in the truncated atmospheres, which strongly suppresses tropospheric NUV radiation and makes it hard to study the subtle changes in CO2 and H2O photolysis rates that ultimately drive the runaway phenomenon.

Based on our pCO2 = 0.01–0.1 bar modeling, we attribute the influence of λcrit and zmax on CO/O2 runaway to their control on the model resolution of the CO2 photolysis peak. 13 M dwarf stars emit proportionately more of their UV radiation at far-UV (FUV) wavelengths compared to the Sun. CO2 is a strong absorber of FUV light, and at high CO2 abundances, this FUV radiation is absorbed high in the atmosphere (Appendix C). Terminating a high-CO2 atmosphere at 54 km (0.34 μbar for our baseline 0.9 bar CO2, 0.1 bar N2 atmosphere) results in a failure to resolve the distribution of CO2 photolysis in the upper atmosphere with altitude (Figure 1). The failure to resolve the CO2 FUV photolysis peak means that CO2 photolysis, and therefore ultimately O production, is artificially confined to the topmost model grid layer. This confinement means that the O atoms that would normally be distributed across a wide altitude range are instead artificially sequestered into a single altitude bin, facilitating their mutual reaction via O+O+M → O2 + M (Figure 2). This is a chain termination reaction, which removes reactive power in the form of radicals from the atmosphere (Grenfell et al. 2018). In particular, the mixing ratio of OH (XOH) is strongly suppressed in truncated atmospheres, and since OH is the main catalyst driving CO2 recombination (Harman et al. 2018), it is unsurprising that as XOH falls, pCO and pO2 rise (Figure B1; Appendix B). That the CO/O2 runaway is fundamentally driven by OH is signaled by the fact that CO increases before O2 as pCO2 increases. This is because OH is the only effective photochemical sink on CO (Kasting 1990), meaning variations in XOH translate immediately into variations in pCO, while O2's sinks are more diverse and its response to XOH is more indirect. In this sense, CO is the canary in the coal mine, signaling variations in XOH in advance of O2 (Schwieterman et al. 2016).

Figure 1.

Figure 1. Non-resolution of the CO2 photolysis peak drives CO/O2 runaway. Concentrations of key atmospheric species (left column) and CO2 and H2O photolysis rates (right column) as a function of altitude, for pCO2 = 0.01, 0.03, and 0.1 bar, CHO chemistry only (S, N excluded). The key photochemical products CO and O2 are highlighted with thicker lines in the left column. "Dry pressure" refers to the atmospheric pressure due to CO2 and N2, excluding the contribution from water vapor. Solid lines correspond to calculations with the model top at 0.34 μbar, while dashed lines correspond to calculations with a higher 100 km model top. At 0.01 bar, the photolysis peak is resolved by both models, and they report identical results. As pCO2 increases, the CO2 photodissociation peak moves upward and is not resolved by the calculations with the 0.34 μbar model top, resulting in large increases in pCO and especially pO2.

Standard image High-resolution image
Figure 2.

Figure 2. Vertically resolved reaction rates of key OH-producing (left column) and radical-radical (right column) reactions, for both truncated (solid lines) and extended (dashed lines) model atmospheres. The top row shows pCO2 = 0.01 bar, for which both model tops return the same results, while the bottom shows pCO2 = 0.1 bar, for which CO/O2 runaway has begun in the truncated model top (but for which an NUV-attenuating ozone layer has not yet formed). CHO chemistry only (S, N excluded). The key reaction O + O + M→O2+M in the right column is highlighted with a thicker line. The concentration of O production in a single model bin in the truncated atmosphere leads to more intense O–O reactions, which are chain termination reactions that remove radicals from the atmosphere and impede the radical-driven recombination of CO and O2.

Standard image High-resolution image

We conducted additional numerical experiments to confirm our explanation that CO/O2 runaway is driven by the confinement of the O produced by CO2 photolysis to a single model layer. If our explanation is correct, then more efficient vertical transport (higher Kz (z)) should inhibit the photochemical runaway because more of the O produced in the topmost layer of the atmosphere will be transported to lower altitudes instead of reacting with other O radicals in chain termination events. Indeed, increasing Kz (z) to Kz (z) = 103 Kz,0(z), where Kz,0(z) is the eddy diffusion profile in our baseline scenario (Figure E2) suppresses pO2 and pCO in our baseline scenario (Table 1, Run 13). Similarly, if our explanation is correct, then increasing the number of layers in the model grid by decreasing the size of the altitude layers should inhibit runaway because the O produced by photolysis is distributed over multiple atmospheric layers instead of being limited to a single layer where it can more readily react with other O. Indeed, increasing the vertical altitude resolution from 1 to 0.1 km suppresses pO2 and pCO in our baseline scenario (Table 1, Run 12).

We also investigated if we could induce a CO/O2 runaway in an extended model grid, which resolved the CO2 photolysis peak by varying Kz (z) or the altitude resolution, and found that we could not. This suggests that models that resolve the CO2 photodissociation peak are less sensitive to other details of the numerical scheme, as expected for physically accurate solutions. For ${z}_{\max }=100$ km, we were unable to induce a CO/O2 runaway by decreasing Kz (z) by setting Kz z = 10−3 Kz,0(z) (Table 1, Run 14). We were similarly unable to induce a runaway by decreasing the vertical resolution by a factor of 5 to 5 km, or increasing the vertical resolution by a factor of 5 to 0.2 km. However, we noticed a weak sensitivity of pO2 and pCO to vertical resolution, with pO2 increasing by 40% and pCO decreasing by 50% as the vertical resolution increased from 1 to 0.2 km.

3.2. Synthetic Spectra

To explore the observational relevance of CO/O2 runaway, we calculate synthetic transmission spectra of a planet in and out of CO/O2 runaway. We specifically simulate and compare spectra of the model atmospheres corresponding to Runs 4 and 5 of Table 1, which are in and out of CO/O2 runaway, respectively. These limited simulations do not constitute a rigorous spectral analysis of the potential observables from abiotic CO2-rich planets orbiting M dwarf stars. Rather, they are narrowly intended to demonstrate the large amplitude of the spectral features that CO/O2 runaway can generate, and therefore the observational relevance of better understanding this theoretical phenomenon and its triggers.

Our synthetic spectra show large variations between the runaway (Run 4) and non-runaway (Run 5) atmospheres (Figure 3). The scale of the spectral features from the runaway atmosphere is comparable to the scale of the spectral features expected from modern Earth-like planets orbiting late M dwarf stars (Fauchez et al. 2020). The overall amplitude of the transmission spectrum is higher for the runaway case due to the lower mean molecular mass of its atmosphere. The runaway atmosphere displays stronger CO spectral features compared to the non-runaway atmosphere and O2 and O3 spectral features, which are absent from the non-runaway atmosphere. The runaway atmosphere displays O3 features as strong as 50 ppm, which are absent from the non-runaway atmosphere. By comparison, JWST has already achieved a precision of <50 ppm in transmission spectroscopy, with even higher precision potentially achievable with additional observation time given the lack of evidence of a photometric noise floor down to a precision of 5 ppm (Lustig-Yaeger et al. 2023). This means that CO/O2 runaway is observationally relevant because it can lead to spectral features that are potentially accessible to JWST. The spectral features due to CO/O2 runaway are also relevant to other facilities such as the Large Interferometer for Exoplanets (LIFE) mission concept (Quanz et al. 2022), and potentially to ground-based facilities as well if observational challenges can be overcome (Currie et al. 2023; Hardegree-Ullman et al. 2023).

Figure 3.

Figure 3. Simulated transmission spectra of an Earth-sized planet orbiting a late M dwarf star with a CO2-dominated atmosphere in and out of CO/O2 runaway, corresponding to Runs 4 and 5 of Table 1. Key molecular absorption features that differ between the runaway/non-runaway cases are highlighted. Runaway spectral features are as strong as 50 ppm and the runaway/non-runaway spectra differ by 10 s of ppm, which is potentially detectable with JWST.

Standard image High-resolution image

4. Discussion

Our work strengthens the case for O2 as a biosignature gas on conventionally habitable planets orbiting M dwarf stars. Abundant atmospheric oxygen is the single strongest remotely detectable evidence of life on Earth (Sagan et al. 1993), and O2 detection often motivates design requirements for facilities and surveys to detect life on exoplanets (Brandt & Spiegel 2014; Rodler & López-Morales 2014; Bixel & Apai 2021; Checlair et al. 2021; Hardegree-Ullman et al. 2023). However, care is required when using O2 as a biosignature because of theoretically proposed mechanisms for the abiotic accumulation of O2 in the atmospheres of conventionally habitable worlds (Meadows et al. 2018). In particular, it was recently argued that conventionally habitable planets with CO2-rich atmospheres orbiting M dwarf stars would abiotically generate abundant (≥0.1 bar) atmospheric oxygen due to photochemical processes (Hu et al. 2020). This regime is highly observationally relevant because (1) CO2-rich atmospheres are expected in the absence of biology, especially for planets orbiting toward the outer edge of the habitable zone (Gaillard & Scaillet 2014; Lehmer et al. 2020), and (2) due to observational biases, only planets orbiting M dwarf stars are accessible to atmospheric characterization with near-term (∼10–20 yr) facilities (Shields et al. 2016). Notably, TRAPPIST-1e and f, the most observationally favorable potentially habitable exoplanets and which are already targeted by JWST observations, fall into this regime, if conventionally habitable (Wolf 2017). Our results demonstrate that the generation of abundant (≥0.1 bar) O2 on conventionally habitable worlds purely through CO2 photodissociation is unlikely, strengthening the case for O2 as a biosignature. However, photochemical false positives remain possible for cold, desiccated worlds that are not conventionally habitable (Gao et al. 2015), and escape-driven false-positive mechanisms remain possible for a broad range of worlds (Ramirez & Kaltenegger 2014; Wordsworth & Pierrehumbert 2014; Luger & Barnes 2015; Tian 2015; Wordsworth et al. 2018).

This paper is the latest in a long series to demonstrate the extreme sensitivity of predicted photochemical concentrations of O2 and CO in CO2-rich conventionally habitable planet atmospheres to diverse background model assumptions, including the lightning flash rate, the absorption cross-sections of CO2 and H2O, the rainout rate and atmosphere/ocean redox balance, and now the altitude of the model top (Wen et al. 1989; Selsis et al. 2002; Segura et al. 2007; Harman et al. 2015, 2018; Ranjan et al. 2020). Such sensitivity is distressing given the desire to use O2 and CO to constrain the presence or absence of biology on exoplanets (Meadows et al. 2018; Schwieterman et al. 2018, 2019; Wogan & Catling 2020), for which one desires a robust, assumption-insensitive prediction of photochemical CO or O2.

Unfortunately, the regime of conventionally habitable planets with thick CO2-rich atmospheres is highly sensitive to background assumptions because it corresponds to a regime with reduced atmospheric forcing from OH, permitting second-order processes to become important. The trace gas composition of modern Earth's atmosphere is to first order controlled by OH, the "detergent of the atmosphere" whose reactivity is the main photochemical control on the abundances of a wide range of trace gases on modern Earth and Mars, and is thought to have done the same for much of their histories (McElroy & Donahue 1972; Parkinson & Hunten 1972; Riedel & Lassey 2008; Catling & Kasting 2017). On planets with oxic atmospheres (e.g., modern Earth), OH is robustly produced by the reaction of H2O with O(1D), sourced primarily from O3 photolysis (Riedel & Lassey 2008). On anoxic planets with low CO2 abundances (e.g., Neoarchaean Earth) or CO2-dominated but thin atmospheres (modern Mars), OH is robustly produced by the direct photolysis of H2O. This robust production of OH efficiently destroys atmospheric trace gases and triggers catalytic cycles that drive the atmosphere back toward equilibrium (Harman et al. 2018). On anoxic planets with abundant CO2 (e.g., Eoarchaean Earth), this atmospheric forcing is severely inhibited because CO2 largely competes for the same photons as H2O, suppressing OH production (Ranjan et al. 2020). The suppression of the first-order atmospheric control from OH opens the door for second-order effects to become important, explaining the outsized relevance of details of the photochemical scheme in this regime. This effect is amplified on planets orbiting M dwarf stars because these stars emit proportionately less light at OH-producing NUV wavelengths (Segura et al. 2005). One might be forgiven for wanting to avoid this regime entirely. However, thick CO2-dominated atmospheres are predicted for Earth early in its history, and are predicted for conventionally habitable planets orbiting near the outer edges of their habitable zones, including the most near-term observationally accessible habitable zone planet, TRAPPIST-1e (Wolf 2017; Rugheimer & Kaltenegger 2018; Lehmer et al. 2020). Therefore, this is not a regime that can be neglected in the search for life on exoplanets.

To address this regime properly, we will need to treat the details, which are not yet fully understood in most cases. For example, the NUV cross sections of CO2 and H2O are still uncertain (Wen et al. 1989; Ranjan et al. 2020; W. Broussard et al. 2023, in preparation). Importantly, this means that while our results contravene the previously proposed inevitability of CO/O2 runaway on CO2-rich M dwarf exoplanets, they do not completely eliminate the possibility of CO/O2 runaway because it is possible that we have omitted or incorrectly implemented one of these second-order processes. These processes are much less important (and therefore poorly constrained) in a solar system context, but may be highly relevant for M dwarf planets. In particular, our modeling neglects ion chemistry in the upper atmosphere, which may be an important source of CO and O2 due to the thermospheric dissociation of CO2 driven by intense M dwarf stellar energetic particle fluxes (J. Kasting, private communication, 2023 May 1). It also neglects novel photoreaction channels for H2O, CO2, and CO, which may be relevant in the uppermost atmosphere (An et al. 2021; Lo et al. 2021; Yang et al. 2023). We conclude that the upper atmosphere is an unexplored potential source of CO/O2 runaway.

While our results show abundant (≥0.1 bar) photochemical O2 is unlikely, they nonetheless admit the possibility of trace photochemical O2. Our models predict abiotic pO2 = 2 × 10−5–4 × 10−3 bar depending on the model branch, assumed stellar spectral energy distribution (SED), and assumed lightning flash rate. Such O2 abundances, while comparable to biotic O2 concentrations on Proterozoic Earth (Harman et al. 2015), do not constitute an observationally relevant biosignature false positive in the near-to-medium term because O2 itself is undetectable at such low concentrations—even much higher, modern Earth-like O2 concentrations (0.2 bar) will be extremely challenging to characterize with either JWST or ELTs (Fauchez et al. 2020; Currie et al. 2023; Hardegree-Ullman et al. 2023). Detection of trace O2 will likely require the advent of purpose-built instruments such as the Habitable Worlds Observatory (Checlair et al. 2021; Clery 2023), but HWO will primarily target Sun-like (FGK) stars and will therefore be relatively unaffected by M dwarf O2 false positives (our work may be relevant to the handful of M dwarf stars on the HWO target list). More observationally relevant is the false-positive potential for O3, which begins to produce IR spectral features in emission at 2 × 10−3 bar for M dwarf planets (Kozakis et al. 2022). This falls within the range of photochemical pO2 produced in our modeling, meaning that photochemistry may produce observationally relevant O3 false positives on M dwarf planets. Additionally, there remain parameters we have not explored which may expand pO2 beyond the pO2 = 2 × 10−5–4 × 10−3 range spanned by our sensitivity tests to date to model choice, lightning flash rate, and assumed stellar SED, such as volcanic outgassing level, combustion chemistry, and temperature–pressure profile (Selsis et al. 2002; Segura et al. 2007; Hu et al. 2012; Grenfell et al. 2018; Harman et al. 2022). Lastly, oceanic chemistry may also impact pO2. For example, the rate of direct recombination of CO and O2 in marine waters is unconstrained, and if efficient may also suppress pO2 and pCO (Harman et al. 2015). We highlight aqueous reactions like O2–CO recombination as priority targets for experimental characterization relevant to understanding exoplanet atmospheres, complementing the acknowledged community priority to characterize gas-phase kinetics (Fortney et al. 2016, 2019). In summary, further theoretical and experimental work is required to establish rigorous limits on predicted trace photochemical O2 and O3 for CO2-rich planets orbiting M dwarf stars, and understand their context-dependent false-positive potential.

Our work illustrates the challenges of projecting atmospheric photochemical models calibrated on solar system planets to the novel photochemical regimes accessible on exoplanets. A model top of 0.34 μbar is capable of reproducing the trace gas composition of modern Earth (Hu et al. 2012). A model top of 0.34 μbar is marginally adequate when modeling CO2-rich planets orbiting Sun-like stars, for which the pressure CO2 photodissociation peak pressure is at ≥1.9 μbar (Appendix C). However, a model top of 0.34 μbar is not adequate when modeling CO2-rich planets orbiting M dwarf stars, for which the photodissociation peak may be as high as 0.04 μbar (Appendix C). Care must be taken when projecting to the novel photochemical regimes accessible on exoplanets to ensure that the large diversity of background parameters in photochemical models are correctly calibrated.

Our work demonstrates the need to consider the upper atmosphere when modeling planets with CO2-rich atmospheres orbiting M dwarf stars. Most exoplanet photochemical models, including all models currently focused on understanding O2 false positives on temperate terrestrial worlds, focus on the lower atmosphere of the planet (troposphere, stratosphere, mesosphere) because this is the part of the atmosphere most relevant to trace gas detectability and planetary habitability. However, for CO2-rich planets orbiting M dwarf stars, the CO2 photodissociation peak can be well above the homopause, which corresponds to the thermosphere in the terrestrial atmosphere (Appendix D; Rumble 2017). This part of the atmosphere is partially ionized, and ion chemistry is important (Tian et al. 2008; Johnstone et al. 2018), which is rarely implemented in the exoplanet photochemistry models being used to study O2 false positives. Similarly, above the homopause, transport is dominated by molecular diffusion, not eddy diffusion, which is not well captured for all molecules in all models (e.g., MEAC implements molecular diffusion of H and H2 only). Accurately modeling CO2-rich atmospheres orbiting M dwarf stars will require resolving the thermosphere, and ideally the entire atmosphere from surface to exobase. Efforts toward resolving exoplanet photochemistry in the upper atmosphere alone have been implemented for oxic, steam, and CO2-dominated exoplanet atmospheres (Tian 2009; Garcia-Sage et al. 2017; Johnstone et al. 2019; Johnstone 2020; Nakayama et al. 2022), and from the surface to the upper atmosphere for oxic exoplanet atmospheres (Chen et al. 2019; Herbst et al. 2019; Cooke et al. 2023). We advocate for the extension of such work to abiotic CO2-dominated atmospheres, which will strongly enhance our understanding of abiotic false-positive scenarios for CO and O2.

Our findings affirm the importance of CO as a discriminant of photochemical O2 production (Schwieterman et al. 2016; Wang et al. 2016). CO/O2 runaway is ultimately driven by the suppression of OH, but OH is effectively the only photochemical sink on CO in conventionally habitable planet photochemical networks, while numerous pathways can suppress O2. Consequently, when CO/O2 runaways do occur in our model, CO rises before O2. We can concoct scenarios where CO is scrubbed from the atmosphere while O2 rises (e.g., through efficient surface deposition of CO but not O2 into the ocean), but this scenario is unlikely on an abiotic planet based on our current understanding of aqueous chemistry (Kharecha et al. 2005; Harman et al. 2015; Hu et al. 2020). We therefore continue to advocate for constraining CO abundances in tandem with O2/O3 abundances when seeking to employ O2 as a biosignature, as has also been argued when seeking to employ CH4 as a biosignature (Thompson et al. 2022). Similarly, our work affirms the value of capstone biosignatures to corroborate potentially ambiguous primary biosignatures like O2 (Leung et al. 2022).

5. Conclusions

We have used a multi-model approach to show that accumulation of abundant (>0.1 bar) photochemical CO and O2 on conventionally habitable rocky planets with CO2-rich atmospheres orbiting M dwarf stars (e.g., TRAPPIST-1e, -1f, if habitable) is unlikely, within the limits of current knowledge. An earlier prediction of abundant photochemical CO/O2 on M dwarf planets was a model artifact, due to a model top that was too low to resolve the CO2 photolysis peak. This model artifact occurred because on M dwarf planets with CO2-rich atmospheres, the CO2 photolysis peak is shifted to very high altitudes, meaning that model tops that are appropriate for solar system planets are too low for the M dwarf regime. Resolving the CO2 photolysis peak robustly eliminates predictions of abundant (>0.1 bar) CO and O2. However, accumulation of trace photochemical O2 (2 × 10−4 bar < pO2 <0.1 bar) together with >1% CO remains a possibility. Such O2 concentrations do not constitute an observationally relevant biosignature false positive in the near-to-medium term because they are likely too low to be detected directly with JWST, but are high enough that they may drive an observationally relevant false positive for O3 as a biosignature, and further work is required to address this possibility. Our work demonstrates the need to accurately resolve the upper atmosphere in order to model O2 false positives on M dwarf planets because on such worlds the CO2 photolysis peak can extend into the heterosphere. Overall, our work strengthens the case for O2 as a biosignature gas by reducing its false-positive potential, but further work is required to fully understand the context-dependent use of O2 and O3 as biosignature gases.

Acknowledgments

We thank an anonymous referee for constructive criticism that substantially improved this paper. We thank Roger Yelle and James Kasting for their helpful discussions. We thank Thomas Fauchez and Ana Glidden for their answers to questions. S.R., E.W.S., and M.L. gratefully acknowledge support by NASA Exoplanets Research Program grant No. 80NSSC22K0235. R.H. was supported by NASA Exoplanets Research Program grant No. 80NM0018F0612. This research has made use of NASA's Astrophysics Data System. The MEAC input and output files underlying Figures 1, B1, 2, and Table 1, along with the scripts used to analyze them and to conduct the calculations presented in Appendices C and D, are publicly available at https://github.com/sukritranjan/co-o2-runaway-revisited-toshare and via Zenodo: doi:10.5281/zenodo.8148741 (Ranjan et al. 2023).

Software: MEAC (Hu et al. 2012), Atmos (Arney et al. 2016), PSG (Villanueva et al. 2018, 2022).

Appendix A: Additional Background: CO2 Photochemical Instability on M dwarf Planets

In this appendix, we briefly review the problem of CO2 photochemical instability, with application to M dwarf planets.

CO2 alone is unstable to photolytic conversion to CO and O2 in conventionally habitable planet atmospheres. CO2 photolyzes efficiently under UV irradiation (Ityaksov et al. 2008):

Equation (A1)

The recombination reaction CO + O + M → CO2 + M is spin forbidden and slow, leaving O2 to accumulate via (Krasnopolsky 2019)

Equation (A2)

However, photodissociation of non-CO2 trace gases can trigger catalytic cycles that recombine CO and O2 back to CO2. For example, on Mars, OH ultimately sourced from H2O photolysis 12 (Kasting & Walker 1981; Nair et al. 1994; Harman et al. 2015; Ranjan et al. 2020):

Equation (A3)

stabilizes 13 CO2, via cycles like (Harman et al. 2018):

Equation (A4)

Equation (A5)

Equation (A6)

The challenge for M dwarf exoplanets is that M dwarfs emit proportionately less NUV (>200 nm) radiation due to their cooler photospheres, while their emission of FUV (<200 nm) remains robust due to their enhanced magnetic activity (Loyd et al. 2016). At habitable temperatures, CO2 known to photoabsorb significantly only at FUV wavelengths (Ityaksov et al. 2008), while H2O is known to photoabsorb at both FUV and NUV wavelengths (Ranjan et al. 2020). The high FUV-to-NUV emission ratio of M dwarf stars therefore results in reduced efficacy of the H2O photolysis-driven chemical cycles, which recombine CO2 relative to CO2 photolysis, resulting in predictions of enhanced CO and O2 relative to planets orbiting M dwarf stars. The problem is exacerbated for high-CO2 planets because abundant CO2 shields H2O from photodissociation, as H2O is restricted to the lower atmosphere by condensation (Harman et al. 2015). If the relative efficacy of H2O photolysis-driven CO2 recombination cycles decreases below a critical point, the atmosphere can enter a state of photochemical runaway, where CO and O2 accumulate to appreciable atmospheric concentrations; this state is termed CO/O2 runaway (Zahnle et al. 2008; Harman et al. 2015; Hu et al. 2020). The debate in the field has been about which conventionally habitable M dwarf planets, if any, access this CO/O2 runaway atmospheric state (Domagal-Goldman et al. 2014; Tian et al. 2014; Gao et al. 2015; Harman et al. 2015, 2018; Hu et al. 2020). In the most recent iteration of this debate, Hu et al. (2020) find the CO/O2 runaway regime is robustly accessed for high-CO2 terrestrial planets orbiting M dwarf stars; in this paper, we test this finding.

Appendix B: Catalytic Cycles in Truncated versus Non-truncated Atmospheres

In this appendix, we more closely examine the mechanism by which suppressed O abundances in truncated atmospheres lead to decreased production of OH and therefore reduced stability of CO2-rich atmospheres to photolysis. Specifically, we follow Harman et al. (2018) in constructing tables of the main catalytic cycles stabilizing CO2 and their rates in the truncated and non-truncated atmospheres with pCO2 = 0.1 bar (Table 2). The rate of the net catalytic cycle is set by the slowest reaction within the cycle.

Table 2. Column-Integrated Rates of CO2 Photolysis and Recombination, Following Harman et al. (2018)

ReactionColumn-integrated Rate (${z}_{\max }{| }_{P=0.34\,\mu bar}$)Column-integrated Rate (${z}_{\max }=100$ km)
 (cm−2 s−1)(cm−2 s−1)
CO2 Photolysis
CO2 + hν → CO+O2.1E123.8E12
CO2 + hν → CO + O(1D)2.6E111.2E11
Total CO2 Photolysis2.4E123.9E12
CO2 Recombination Cycle 1
H + O2 + M → HO2 + M3.0E112.3E12
O + HO2 → OH + O2 1.0E122.6E12
CO+OH → CO2 + H1.4E123.5E12
Net (CO+O → CO2)3.0E112.3E12
CO2 Recombination Cycle 2
H+CO+M → CHO+M9.7E117.4E11
CHO+O2 → CO+HO2 9.6E117.1E11
O+HO2 → OH+O2 1.0E122.6E12
CO+OH → CO2 + H1.4E123.5E12
Net (CO+O → CO2)9.6E117.1E11
CO2 Recombination Cycle 3
H2O2 + hν → 2OH9.5E101.3E11
2CO + 2OH → 2CO2 + 2H7.1E111.8E12
2H + 2O2 + 2M → 2HO2 +2M1.5E111.1E12
2HO2 → H2O2 + O2 8.6E101.0E11
Net (2CO+O2 → 2CO2)8.6E101.0E11
Direct CO2 Recombination
CO+O+M → CO2 + M9.5E113.7E11
Total CO2 Recombination2.4E123.6E12
(Cycle 1 + Cycle 2 + 2×Cycle 3 + Direct Recomb.)

Note. We compare truncated and non-truncated atmospheres for pCO2 = 0.1 bar, CHO chemistry only. Cycles 1 and 3 are ultimately drawn from the work of Stock et al. (2012) on the Martian atmosphere. Cycle 2 is identified in this work, where it is relevant to the high-CO2, high-CO atmospheres we consider. The CO2 recombination mechanisms we identify here balance >90% of CO2 photolysis, meaning we have identified the main CO2 recombination mechanisms.

Download table as:  ASCIITypeset image

The chemical cycles in the C-H-O system are complex and nonlinear, but the catalysis tables lend some insight (Harman et al. 2018). H plays a key role in recombining CO2 because it facilitates the generation of intermediate species en route to OH:

Equation (B1)

Equation (B2)

Equation (B3)

and

Equation (B4)

Equation (B5)

Equation (B6)

Equation (B7)

However, OH and O are also main sources of H, via

Equation (B8)

Equation (B9)

A decrease in O therefore decreases H production, which decreases OH production, which further decreases H production, which relies on OH. This is a positive feedback loop, whereby small changes in [O] can be amplified into larger changes in [OH]. We suggest this positive feedback loop as the mechanism whereby the relatively modest changes in [O] driven by the truncated model top (Figure B1) can be amplified into relatively large changes in [OH], and therefore pCO and pO2.

Figure B1.

Figure B1. A closer look at the mechanics of CO/O2 runaway. Change in column-integrated mixing ratio of key species (top) and column-integrated reaction rates for main radical species (middle) and main OH-producing reactions (bottom) as a function of pCO2, relative to pCO2 = 0.01 for which both truncated (solid lines) and extended (dashed lines) model calculations produce the same result. Suppression in OH concentrations, likely due to suppression of OH production pathways, underlies CO/O2 runaway.

Standard image High-resolution image

Appendix C: SED-driven Changes in Photodissociation Peak Pressure

The higher FUV/NUV ratio associated with M dwarf irradiation leads to lower CO2 photolysis peak pressures (higher CO2 photolysis peak altitudes) relative to Sun-like stars. To illustrate this phenomenon, we derive an estimate of the pressure of peak dissociation for a well-mixed gas g in hydrostatic equilibrium, Ppeak,g, and apply it to CO2 in the atmospheres of CO2-dominated planets orbiting different stars. This calculation is crude; for example, CO2 mixing ratios may decrease with altitude in the uppermost planetary atmosphere due to robust photolysis and diffusive separation, therefore necessitating going deeper into the atmosphere to get the same absorption compared to the well-mixed case. However, the general trends of this calculation give some intuition for why M dwarf SEDs lead to higher ${P}_{\mathrm{peak},{\mathrm{CO}}_{2}}$.

The pressure of peak photodissociation of a gas g occurs when the UV optical depth of that gas τg,UV (measured traveling from space downward normal to the planet surface) satisfies (Catling & Kasting 2017)

Equation (C1)

where θ0 is the stellar angle of incidence. For an atmosphere in hydrostatic equilibrium (Catling & Kasting 2017),

Equation (C2)

where P is the pressure, μ is the mean molecular weight of the atmosphere, g is the acceleration due to gravity, and N is the number column density of the atmosphere above pressure P. Since ${\tau }_{\mathrm{UV},{\rm{g}}}={N}_{g}\overline{{\sigma }_{\mathrm{UV},{\rm{g}}}}$, where $\overline{{\sigma }_{\mathrm{UV},{\rm{g}}}}$ is the mean UV photolysis cross section of g, for a well-mixed gas with molar concentration rg (i.e., Ng = rg N), we can substitute Equation (C2) into Equation (C1) to write

Equation (C3)

We apply Equation (C3) to estimate the peak CO2 photolysis pressure for planets orbiting different kinds of stars for our CO2-dominated baseline scenario. For our CO2-dominated baseline scenario, ${r}_{{\mathrm{CO}}_{2}}=0.9$, g = g, μ = 42.4 amu, we have chosen θ0 = 57fdg3° (Zahnle et al. 2008). To calculate $\overline{{\sigma }_{\mathrm{UV},{\mathrm{CO}}_{2}}}$, we calculate the mean CO2 cross-section weighted by the stellar number flux F*:

Equation (C4)

We calculate $\overline{{\sigma }_{\mathrm{UV},{\mathrm{CO}}_{2}}}$ and ${P}_{\mathrm{peak},{\mathrm{CO}}_{2}}$ for solar irradiation, TRAPPIST-1 irradiation (Peacock et al. 2019 model 1a) with λ0 = 110 nm, and TRAPPIST-1 irradiation with λ0 = 120 nm (Table 3). σCO2(λ) and F*(λ) for both scenarios are illustrated in Figure C1. In all cases, λ1 = 202 nm, corresponding to the end of detected CO2 absorption at room temperature (Ityaksov et al. 2008). The higher FUV/NUV ratio of TRAPPIST-1 irradiation increases $\overline{{\sigma }_{\mathrm{UV},{\rm{g}}}}$ by 50× relative to Solar irradiation, decreasing Ppeak,g proportionately. This means that to resolve the CO2 photolysis peak correctly for planets orbiting M dwarf stars, it is necessary to extend the model grid to much lower pressures compared to planets orbiting Sun-like stars.

Figure C1.

Figure C1. Top: top-of-atmosphere (TOA) number flux from the Sun and TRAPPIST-1 (left y-axis). Also co-plotted are the photolysis cross sections of CO2 (right abscissa). Bottom: CO2 cross sections weighted by the stellar emission of TRAPPIST-1 and the Sun. The bluer spectrum of TRAPPIST-1 increases the weight of shortwave ${\sigma }_{{{\rm{CO}}}_{2}}$, which is larger, driving an increase in $\overline{{\sigma }_{\mathrm{UV},{\rm{g}}}}$.

Standard image High-resolution image

Table 3. Estimates of $\overline{{\sigma }_{\mathrm{UV},{\rm{g}}}}$ and ${P}_{\mathrm{peak},{\mathrm{CO}}_{2}}$ for Different Stellar Irradiation

Star λ0 $\overline{{\sigma }_{\mathrm{UV},{\rm{g}}}}$ ${P}_{\mathrm{peak},{\mathrm{CO}}_{2}}$
 (nm)cm2 (μbar)
Sun1102 × 10−20 2
TRAPPIST-11207 × 10−20 0.6
TRAPPIST-11109 × 10−19 0.04

Note. The higher FUV/NUV ratio of M dwarf stars (here represented by the late M endmember TRAPPIST-1) leads to higher $\overline{{\sigma }_{\mathrm{UV},{\rm{g}}}}$ and lower ${P}_{\mathrm{peak},{\mathrm{CO}}_{2}}$ relative to Sun-like stars.

Download table as:  ASCIITypeset image

In this calculation, we have approximated CO2 as well mixed. This is a good assumption for the lower atmosphere, but our extended atmospheric simulations extend into the heterosphere where CO2 is not well mixed, and its abundance decreases with altitude. Therefore, our calculation, which assumes constant CO2 abundance with altitude, formally constitutes a lower bound on ${P}_{\mathrm{peak},{\mathrm{CO}}_{2}}$. Nevertheless, it illustrates the general mechanism by which ${P}_{\mathrm{peak},{\mathrm{CO}}_{2}}$ is decreased on planets orbiting M dwarf stars, i.e., the higher FUV/NUV ratio increasing $\overline{{\sigma }_{\mathrm{UV},{\rm{g}}}}$.

Appendix D: Homopause Estimates

In this appendix, we estimate the homopause pressure for CO2-dominated temperate terrestrial planets to demonstrate that the CO2 photodissociation peak can extend into the heterosphere for M dwarf planets, demonstrating the need to resolve the upper atmosphere when modeling such worlds.

The homopause pressure is the pressure level in the atmosphere when

Equation (D1)

where D12(T, n) is the diffusion coefficient of gas g1 diffusing through a background gas g2 at temperature T and atmospheric number density n. D12(T, n) takes the general form (Banks & Kockart 1973)

Equation (D2)

where A12 and s12 are specific to each (g1, g2) pair. Combining Equations (D1) and (D2) with the ideal gas law (P = nkT), we can write

Equation (D3)

In Earth's upper atmosphere near the homopause, Kz (z) = 4 × 105 − 2 × 106 cm2 s−1, depending on which tracer is used (Hunten 1975; Swenson et al. 2019). We adopt an Earth upper atmosphere Kz,homo,⊕ = 1 × 106 cm2 s−1, which may be scaled to ${K}_{{\rm{z}},\mathrm{homo},{\mathrm{CO}}_{2}-\mathrm{dominated}}=\tfrac{29.0}{42.4}(1\times {10}^{6}\,{\mathrm{cm}}^{-2}\,{{\rm{s}}}^{-1})\,=7\,\times \,{10}^{5}\,{\mathrm{cm}}^{-2}\,{{\rm{s}}}^{-1}$ for our CO2-dominated benchmark scenario (Hu et al. 2012). Further, in the upper atmosphere, T = 175 K in our baseline scenario, meaning we can set Thomo = 175 K. With these values, we evaluate Equation (D3) for Ar, H2, and H diffusing through CO2 and find homopause pressures of 0.07, 0.4, and 0.6 μbar, respectively (Table 4). For comparison, the photolysis peak pressure is 0.1 μbar for pCO2 = 0.1 bar (Figure 1), and 0.01 μbar for pCO2 = 0.9 bar as in our baseline scenario. This means that to accurately model the photochemistry of CO2-rich planets orbiting M dwarf stars, it is necessary to extend model grids well into the heterosphere. We have done so here, but our calculation is not accurate because it does not accurately represent vertical transport in this part of the atmosphere, especially molecular diffusion, and because it does not include the ion chemistry important in the thermosphere and above. We advocate for the construction of exoplanet photochemistry models that resolve the atmosphere from surface to exobase to accurately treat the important, observationally relevant problem of CO2-rich atmospheres orbiting M dwarf stars, and especially their propensity to accumulate photochemical CO and O2.

Table 4. Homopause Estimates for Different Gases in Our CO2-dominated Baseline Planetary Scenario

GasBackground Gas A12 s12 Phomo (μbar)References
ArCO2 7.16E160.6460.07Marrero & Mason (1972)
H2CO2 2.15E170.750.4Marrero & Mason (1972), Hu et al. (2012)
HCO2 3.87E170.750.6Marrero & Mason (1972), Zahnle et al. (2008), Hu et al. (2012)

Note. The CO2 photodissociation peak for CO2-dominated atmospheres lies in the heterosphere.

Download table as:  ASCIITypeset image

Appendix E: Detailed Planet Scenario

In this appendix, we give more details of our planetary scenario to facilitate the reproduction of our work. Table 5 summarizes further details on our general planetary scenario and model setups. Table 6 presents the detailed species-by-species boundary conditions employed in our simulations. Figure E1 shows the stellar SEDs assumed in this work. Figure E2 shows the baseline atmospheric structure profiles (T(z), P(z), and Kz (z)) assumed in this work. In constructing the atmospheric structure profiles for the variable pCO2 runs shown in Figures 1, B1, and 2, we followed Hu et al. (2012), i.e., assuming dry adiabatic evolution to a 175 K stratosphere (with thermodynamic parameters drawn from Pierrehumbert 2010) and scaling modern Earth's eddy diffusion profile by bulk atmospheric mean molecular mass.

Figure E1.

Figure E1. TRAPPIST-1 spectra employed by MEAC and Atmos in this study.

Standard image High-resolution image
Figure E2.

Figure E2.  TP and Eddy diffusion profiles employed by MEAC and Atmos in this study.

Standard image High-resolution image

Table 5. Simulation Parameters For Planetary Scenario

Scenario ParameterMEACATMOS
Model referenceHu et al. (2012)Arney et al. (2016)
Reaction networkAs in Ranjan et al. (2022) a Archean Scenario
 (excludes, C>2-chem)(includes Ranjan et al. 2020 updates)
Stellar typeTRAPPIST-1 (Peacock et al. 2019)TRAPPIST-1 (Peacock et al. 2019)
Stellar constant relative to modern Earth0.5920.592
Planet size1 M, 1 R 1 M, 1 R
Surface albedo0.0.25
Major atmospheric components0.9 bar CO2, 0.1 bar N2 b 0.9 bar CO2, 0.1 bar N2 b
Surface temperature (z = 0 km)288 K288 K
Surface ${r}_{{{\rm{H}}}_{2}{\rm{O}}}$ (lowest atmospheric bin)0.010.01
Eddy diffusion profileSee Figure E2 See Figure E2
Temperature–pressure profileSee Figure E2 See Figure E2
Vertical resolution1 km0.5 km
RainoutEarth-like; rainout turned off forEarth-like (all species)
 H2, CO, CH4, NH3, N2, C2H2, 
 C2H4, C2H6, and O2 to simulate 
 saturated ocean on an abiotic planet 
LightningSimulated via ϕCO (z = 0) = ϕNO(z = 0) = 1 × 108 On
Global redox conservationNoNo

Notes.

a In MEAC, pN2 is fixed. In Atmos, pN2 is adjusted to maintain dry P = 1 bar. b Default full chemistry CHOSN model runs contain 86 species linked by 734 reactions. For model runs where we exclude HNO4 and N2O5, our network contains 84 species linked by 723 reactions. For model runs where we exclude S and N chemistry entirely, we set the initial concentrations and emission fluxes of S- and N-containing species to 0, but otherwise retain the full 734-reaction photochemical network.

Download table as:  ASCIITypeset image

Table 6. Species Choice and Treatment for the Photochemical Models Used in This Study

SpeciesType Surface FluxSurface Mixing RatioDry Deposition VelocityTOA Flux
 MEACATMOS(cm−2 s−1)(Relative to CO2+N2)(cm s−1)(cm−2 s−1)
CO2 XX0.900
N2 CC0.100
H2OXX0.0100
H2 XX3 × 1010 0Diffusion limited
COXX1 × 108 1 × 10−8 0
CH4 XX3 × 108 00
SO2 XX3 × 109 10
H2SXX3 × 108 0.0150
HXX01Diffusion limited
OXX010
O(1D)XF000
O2 XX000
O3 XX00.40
OHXXX10
HO2 XX010
H2O2 XX00.50
CH2OXX00.10
CHOXX00.10
CXF000
CHXX000
CH2 XX000
1CH2 XF000
3CH2 XX000
CH3 XX000
CH3OXX00.10
CH4OX00.10
CHO2 X00.10
CH2O2 X00.10
CH3O2 XX000
CH4O2 X00.10
C2 XX000
C2HXX000
C2H2 ?X000
C2H3 ?X000
C2H4 ?X000
C2H5 XX000
C2H6 XX01 × 10−5 0
C2HOX000
C2H2OX00.10
C2H3OX00.10
C2H4OX00.10
C2H5OX00.10
SXX000
S2 XX000
S3 XF000
S4 XF000
SOXX000
1SO2 XF000
3SO2 XF000
SO3 XX010
HSXX000
HSOXX000
HSO2 X000
HSO3 XF00.10
H2SO4 XX010
H2SO4(A)AA00.20
S8 X000
S8(A)AA00.20
OCSXX00.010
CSXX00.010
CH3SX00.010
CH4SX00.010
NXX000
NH3 X010
NH2 X000
NHX000
N2OX000
NOXX1.0 × 108 0.020
NO2 XX00.020
NO3 X010
HNOXX000
HNO2 XF00.50
HNO3 XX040
HCNX00.010
CNX00.010
CNOX000
HCNOX000
CH3NO2 X00.010
CH3NO3 X00.010
CH5NX000
C2H2NX000
C2H5NX000
N2H2 X000
N2H3 X000
N2H4 X000
C3H2 X000
C3H3 X000
CH3C2HX000
CH2CCH2 X000
C2H5CHOX000
C2H5CHOX000
C3H6 X000
C3H7 X000
C3H8 X000
C2H4OHX000
C2H2OHX000
C2H2COX000
C2H3COX000
C2H3CHOX000
HCSX000
CS2 X000
HCAER1A000
HCAER2A000

Note. (1) For species type, for each model, "X" means the full continuity-diffusion equation is solved for the species; "F" means it is treated as being in photochemical equilibrium; "A" means it is an aerosol and falls out of the atmosphere; "C" means it is treated as chemically inert; and "⋯" means that it is not included in that model. Note that boundary conditions like dry deposition velocity are not relevant for Type "F" species, since transport is not included for such species. The exclusion of a species from a model does not necessarily mean that the model is incapable of simulating the species, but just that it was not included in the atmospheric scenario selected here. (2) For the bottom boundary condition, either a surface flux is specified, or a surface mixing ratio. N2 is a special case in the Kasting model and in ATMOS; in these models, [N2] is adjusted to set the total dry pressure of the atmosphere to be 1 bar (to account for outgassed species and photochemical intermediates). Consequently, pN2 ≲ 0.1 bar in these models. (3) TOA flux refers to the magnitude of outflow at the TOA; hence, a negative number would correspond to an inflow. (4) For model runs where we exclude S and N chemistry entirely, the emission fluxes of S- and N-containing species to 0, but otherwise retain the full 734-reaction photochemical network.

Download table as:  ASCIITypeset images: 1 2

We draw our TRAPPIST-1 spectrum from Peacock et al. (2019), model number 1A. This spectrum is binned to 1 nm resolution for use in MEAC, and to the standard Atmos model 750 point grid for use in Atmos (Lincowski et al. 2018).

Footnotes

  • 10  
  • 11  
  • 12  

    Arney et al. (2016); main branch, Archean template, commit b98a47527d14e6cf82e0edf4640fc7fba6d09cc8.

  • 13  

    Specifically, the channel CO2 + hν → CO + O(1D)

  • 12  

    While H2O can undergo three-body photolysis as well, this only occurs at EUV wavelengths confined to the upper atmosphere, and initial sensitivity tests indicate no substantial impact on the lower atmosphere (An et al. 2021).

  • 13  

    On Venus, CO2 recombination is also facilitated by HCl photodissociation; this recombination mechanism is less applicable to conventionally habitable worlds with hydrological cycles, where highly soluble HCl is efficiently scrubbed from the atmosphere by rain (Demore & Yung 1982; Yung & Demore 1982; Lightowlers & Cape 1988).

Please wait… references are loading.
10.3847/2041-8213/ad037c