Open Access
Issue
A&A
Volume 693, January 2025
Article Number A303
Number of page(s) 35
Section Planets, planetary systems, and small bodies
DOI https://doi.org/10.1051/0004-6361/202450307
Published online 28 January 2025

© The Authors 2025

Licence Creative CommonsOpen Access article, published by EDP Sciences, under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0), which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.

This article is published in open access under the Subscribe to Open model. Subscribe to A&A to support open access publication.

1 Introduction

After planet formation, rocky planets1 can start their evolution with a global magma ocean and outgassing from its interior that replaces the primary hydrogen-helium envelope inherited from the protoplanetary disk with a secondary atmosphere (e.g., Lammer et al. 2018; Stüeken et al. 2020). The composition of the secondary atmosphere depends on the redox state of the mantle (Deng et al. 2020; Ortenzi et al. 2020).

It is generally recognized that the magma ocean not only represents a direct link between atmospheric properties and the rocky planet’s mantle, but is further crucial for determining the abundances of important volatiles such as water and carbon dioxide in the mantle and the atmosphere (Chao et al. 2021; Barth et al. 2021; Moore et al. 2023; Krissansen-Totton & Fortney 2022). Therefore, despite its relatively short duration (1–100 Myrs), the magma ocean phase “sets the stage” for the long-term planetary evolution over billion of years (e.g Chao et al. 2021; Lammer et al. 2018; Stüeken et al. 2020).

For Earth-sized planets like Venus, Earth, and the potentially habitable exoplanets TRAPPIST-1 e, f, and g, predominantly H2O-CO2 outgassing is expected (Deng et al. 2020). However, a water-vapor-dominated atmosphere is subject to extreme ultra violet irradiation (XUV) and thus H2O photolysis with subsequent escape of H2 can occur. Therefore, water-dominated atmospheres are prone to significant mass loss, even in the Solar System with its relatively calm host star. Hamano et al. (2013) demonstrated that atmospheric erosion in the inner Solar System could deplete all of Earth’s oceans within 100 Myrs. Venus experienced a relatively long magma ocean stage of 100 Myrs, during which most of its water was lost, whereas the shorter magma ocean stage on Earth, lasting just a few million years, prevented significant water loss. For Earth-sized planets, the magma ocean stage thus plays a particularly critical role in determining their surface water content, and thus their habitability (e.g., Barth et al. 2021; Hamano et al. 2013; Tian & Ida 2015). The interaction between H2O outgassing and atmospheric escape during the magma ocean stage is even more crucial for rocky planets orbiting active M dwarf stars, which emit intense XUV flux during a prolonged pre-main-sequence stage (e.g., Johnstone et al. 2021; Tian & Ida 2015).

It was shown by Raymond et al. (2022) that the TRAPPIST- 1 planets – in contrast to Earth – were not strongly modified by large impactors during late accretion, which would alter the composition of the outgassed atmosphere after the magma ocean stage (e.g., Zahnle et al. 2020, 2015). Thus, any CO2 and H2O observed in the atmospheres today must be primordial (see e.g., Krissansen-Totton & Fortney 2022). Hence, atmospheric constraints on the rocky planets in the TRAPPIST-1 system may shed light on the volatile budget of the magma ocean, planet formation, and also the potential for habitability.

Initial observations with JWST suggest that TRAPPIST-1 b (Greene et al. 2023) and c (Zieba et al. 2023) do not possess dense CO2 atmospheres and even indicate the absence of a substantial atmosphere with psurf ≥ 0.1 bar. However, recent analyses suggest that the MIRI data for TRAPPIST-1 c are consistent with a variety of atmosphere compositions: water vapor (Acuña et al. 2023), O2 with CO2 or H2O (Lincowski et al. 2023) with surface pressures up to 40 bar. These results are, however, based on only five secondary eclipses for TRAPPIST-1 b and four for TRAPPIST-1 c, which may also be affected by stellar contamination (Lim et al. 2023).

The combination of ultraprecise density constraints (Agol et al. 2021) and interior models (Noack et al. 2016; Unterborn et al. 2018; Dorn et al. 2018) suggest that at least TRAPPIST- 1 g currently incorporates >10 wt% of water in its interior structure, which would correspond in mass to more than 100 terrestrial oceans (TOs) of water2 (Unterborn et al. 2018; Barth et al. 2021; Raymond et al. 2022). The high inferred volatile content of TRAPPIST-1 g is supported by planet formation models (Miguel et al. 2020; Unterborn et al. 2018; Schoonenberg et al. 2019) that suggest a volatile-rich formation scenario for the outer TRAPPIST-1 planets with initial water mass fractions of up to 50 wt%. Determining the magma ocean solidification time for rocky planets and the resulting distribution of volatiles are thus pressing science questions of current and future missions like PLATO (Turbet et al. 2019; Schlecker et al. 2024) and mission concepts like the Large Interferometer for Exoplanets (LIFE, Bonati et al. 2019).

There are, however, still several open questions in the modeling of magma oceans on diverse rocky planets. Many previous magma ocean models, including the open-source MagmOc1.0 (Barth et al. 2021) as part of the VPLANET modeling suite (Barnes et al. 2020), assumed a pure H2O steam atmosphere (Hamano et al. 2013; Goldblatt et al. 2013; Lichtenberg et al. 2021). Water is not the only critical molecule in terrestrial planet evolution Elkins-Tanton (2008); Nikolaou et al. (2019); Krissansen-Totton & Fortney (2022). Mixed H2O-CO2 outgassing can also modify the volatile distribution as the atmosphere composition changes from an initial CO2 -dominated atmosphere to a H2O-dominated atmosphere one (Bower et al. 2019). The impact of simultaneous H2O and CO2 outgassing during the magma ocean stage on the potentially habitable TRAPPIST-1 planets e, f, and g together with atmospheric escape has not yet been investigated.

We tackle for the first time the H2O-CO2 outgassing feedback between H2O and CO2 outgassing on TRAPPIST-1 e, f, and g with an upgraded version of the Barth et al. (2021) model, VPLanet/MagmOc2.0. In Sect. 2.2, we first introduce a new solution for coupled outgassing of two volatiles. We next present an update of the atmospheric evolution model, in which we take into account the vertical extent of a mixed CO2-H2O atmosphere (Sect. 2.4).

We employed full radiative transfer (RT) calculations with petitRADTRANS (Mollière et al. 2019) for two different atmosphere treatments: in the RT grid model, we compiled the outgoing longwave radiation (OLR) in a 3D emission grid for different surface pressures, surface temperatures, and water volume mixing ratios and interpolated during the simulation time (Sect. 2.5). In the corrected gray atmosphere model, we used the RT grid calculation to formulate an empirical approximation that reproduces the results of the RT grid atmosphere model generally within 10% accuracy (Appendix B).

We also investigated how the runaway greenhouse limit changes with different surface gravities (7.5–22.5 m/s2; Sect. B.4). After validation of various Earth scenarios and outgassing laws (Appendix C), we applied MagmOc2.0 to the potentially habitable TRAPPIST-1 planets e, f, and g, in which we investigated initial volatile contents between 1 and 100 TO H2O (Sect. 3). For each initial water scenario, we calculated evolutionary tracks with no CO2 and initial CO2 mass equal to 0.3× and 1× initial water mass, respectively, to illustrate the impact of additional CO2 on the thermal and volatile evolution for a relatively coarse H2O grid. For each evolution scenario, we further implemented two albedo (α) assumptions: α = 0.75 for a better comparison with Barth et al. (2021), and α = 0, the clear-sky assumption.

We first present for TRAPPIST-1 g and the clear-sky scenario three example evolutions with an initial water mass of 1, 5, and 100 TO (Sect. 3.1) as in Barth et al. (2021). Subsequently, a concise overview of solidification times and remaining water fraction in the solid mantle after the magma ocean evolution is given for TRAPPIST-1 e, f, and g, for an initial water mass between 1 and 100 TO and for the three CO2 ratios (Sect. 3.2). Finer evolution grids with respect to initial water and CO2 masses for TRAPPIST-1 e, f, and g are found in a detailed grid of end states of H2O, CO2, and O2 partial pressures as well as sequestered H2O and O2 in the mantle. We further investigate the strong impact of CO2 on the desiccation, and thus the magma ocean lifetime for water-poor composition (≤10 TO H2O).

We tie the explored magma ocean simulations to new constraints on the interiors of all TRAPPIST-1 planets. Here, the inner planets (b, c, and d) yield a constraint on the iron fraction of 27 wt-% that allows us to place TRAPPIST-1 e in the dry water regime (≤10 TO H2O), for which a partly desiccated CO2- atmosphere is expected. The amount of abiotically produced O2 in these scenarios depends strongly on the magma ocean lifetime. TRAPPIST-1 f and g are expected to be in the water-rich regime (>>10 TO H2O), based on the interior models, for which the magma ocean ends in a “wet” CO2 atmosphere.

Next, we explore the impact of an extended H2O and CO2 atmosphere on the planetary radii and measured bulk density. We also present a compositional model of the refractory elements and volatiles present during planet formation adjusted for the metallicity of TRAPPIST-1, and compare it to interior structures with respect to the iron fractions and volatile ratios of TRAPPIST-1 e, f, g that are compatible with the measured masses and radii of Agol et al. (2021) (Sect. 3.3). The results of the magma ocean, composition, and interior models are presented in Sect. 4 and discussed in Sect. 5. We conclude in Sect. 6 that comparative studies between TRAPPIST-1 e and g and the inner planets are warranted to identify end states of volatile-poor formation (inner planets), desiccated evolution (TRAPPIST-1 e), and volatile-rich formation with little modification by desiccation (TRAPPIST-1 g). We outline future avenues to improve MagmOc2.0 for a better link to planet formation and also to tackle the full diversity of outgassed atmospheres during the crucial magma ocean stage (Sect. 7).

2 Methods

We utilized and expanded the open source VPLANET framework (Barnes et al. 2020) that connects stellar and planetary processes in order to simulate the evolution of stars and planets over time spans of gigayears. For the present paper, we used the stellar module, which tracks the bolometric luminosity of the star according to the Baraffe et al. (2015) stellar evolution model grid, and the XUV evolution according to the Ribas et al. (2005) model originally developed for solar twins, but which appears compatible with lower-mass stars (see e.g., Richey-Yowell et al. 2022). The VPLANET AtmEsc module tracks water photolysis by the stellar radiation, hydrogen escape via energy- and diffusion-limited escape, and oxygen escape via hydrodynamic drag (Watson et al. 1981; Hunten et al. 1987; Luger & Barnes 2015). The VPLANET EqTide module simulates tidal effects, including frictional heating for the both the constant-phase-lag and constant-time-lag models (Ferraz-Mello et al. 2008; Leconte et al. 2010; Barnes et al. 2013). The RadHeat module tracks the radiogenic heating of the unstable isotopes 40K, 232Th, 235U, and 238U.

2.1 MagmOc approach

MagmOc is a geophysical and geochemical model for the coupled mantle and atmosphere when the mantle is at least partially molten (Barth et al. 2021). We assume a bulk silicate Earth composition from O’Neill & Palme (1998), following Schaefer et al. (2016). Table 1 provides the most relevant geophysical parameters for this work.

Our magma ocean model assumes as outlined in Barth et al. (2021) efficient cooling via convection of the magma ocean, which is only true until the melt fraction at the surface drops below 0.4. This condition is reached in our model typically with surface temperatures of 1650 K and a solidification radius rs that already comprises about 99% of the planetary radius. Below this temperature, we follow the argumentation of Debaille et al. (2009) that a thick thermal boundary layer may be neglected toward the end of the magma ocean because crystallization of iron-rich minerals lead to overturning near the surface and a resetting of the thermal boundary layer. The module thus switches to a solid-like viscosity (Barth et al. 2021, Eq. (3)) and advances solidification and thermal evolution further until rs is equal the planetary radius, which is typically for surface temperatures of 1400 K. At this point, our model does not further advance mantle and surface temperature evolution in contrast to other models (Lebrun et al. 2013; Bower et al. 2019, 2022; Krissansen-Totton & Fortney 2022; Lichtenberg et al. 2021).

We note that it is worthwhile to extend simulations beyond the mantle solidification time with MagmOcV2.0 for cases with significant atmospheric erosion during the magma ocean stage to capture the final stages of complete desiccation via atmospheric escape that is an upper atmosphere process. The caveat is here that a) surface temperatures and heat fluxes are kept “artificially” high after solidification and b) no water condensation can occur, which would “save” the remaining water from being lost. Thus, water loss due to atmospheric erosion may be overestimated.

Table 1

Parameters of the geophysical model for MagmOcV2.0

Table 2

Parameters of the volatile model for MagmOcV2.0.

2.2 H2O-CO2 volatile budget

Here, we focus on advancing the volatile treatment of MagmOcV1.0, which is needed for coherent modeling of complete devolatilization on rocky exoplanets with a large variety of initial compositions and H2O-CO2 ratios during the magma ocean stage. Table 2 shows an overview of the relevant parameters.

The volatile budget was set up similarly to Barth et al. (2021), where we assume at all times that a volatile i (i=H2O,CO2) is distributed in the fully coupled magma ocean-atmosphere (MOA) system. This system consists of a liquid magma ocean, Mliq , the atmosphere, Matm, and crystals forming within the liquid magma ocean, Mcrystal ; the latter was set to zero initially.

We further assumed initially that the volatiles, i, in the liquid magma ocean and atmosphere are in balance with each other for a given initial volatile mass, Miinitial$M_i^{initial{\rm{}}}$, which results in an initial volatile mass fraction, Fiinitial$F_i^{initial{\rm{}}}$, governing both the amount of volatile outgassed to build an atmosphere and the amount of volatile dissolved in the melt. The mantle begins to solidify from bottom to top; that is, we assumed that the crystals fall to the bottom of the magma ocean, Mcrystal, incorporating a small amount of the available volatile mass from the melt (FiMliq). Sequestering of volatiles in the solidifying mantle, dMS ol, is regulated by the mantle-averaged constant partition coefficient, ki. The volatile mass in the solidified mantle (MiSol)$\left( {M_i^{Sol}} \right)$ is inaccessible for the MOA, and thus comprises a sink term. As the magma ocean depth, and thus the total mass of Mliq , decreases with increasing solidification radius, drs dt${{{\rm{d}}{r_s}} \over {{\rm{d}}t}}$ , the mass fraction, Fi , of the volatile in the melt typically increases, leading to more outgassing.

In summary, we assumed a mass balance in the MOA for a volatile, i, of the form (see also Barth et al. 2021; Bower et al. 2019): Mimoa=Micrystal +Miliq+Miatm            =kiFiMcrystal +FiMliq+4πrp2ɡpi, mass ,$\matrix{ {M_i^{{\rm{moa}}} = M_i^{{\rm{crystal}}} + M_i^{{\rm{liq}}} + M_i^{{\rm{atm}}}} \hfill \cr {\,\,\,\,\,\,\,\,\,\,\,\, = {k_i}{F_i}{M^{{\rm{crystal}}}} + {F_i}{M^{{\rm{liq}}}} + {{4\pi r_{\rm{p}}^2} \over }{p_{i,{\rm{mass}}}},} \hfill \cr } $(1)

where Fi is the volatile mass fraction of the liquid part of the magma ocean, Miliq$M_i^{liq}$. In addition, we assumed that the volatile, i, is partly sequestered in the crystallized magma ocean, Micry$M_i^{cry}$. The mass of the volatile in the atmosphere was also set by Fi, where the partial pressure of volatile, i, was determined by a Henrian fit to laboratory data (see Fig. A.1 and references in Appendix A): pi,part=(Ficiai)bi,${p_{i,{\rm{part}}}} = {\left( {{{{F_i} - {c_i}} \over {{a_i}}}} \right)^{{b_i}}},$(2)

where ai, bi, and ci are the fit coefficients. We note that ci is a term that suppresses outgassing and is only used in the solubility law of Elkins-Tanton (2008). For the TRAPPIST-1 planet simulations, we used the solubility law of Nikolaou et al. (2019), for which ci was set to zero. Since we implemented and tested both solubility laws (Appendix A), we show here the generalized form of the Henrian fit with the suppression term.

As is pointed out by Bower et al. (2019), the mass weighted pressure of volatile, i – that is, pi,mass – enters into the mass balance. The relation between the partial and mass weighted pressure is: pi,mass=μiμ¯atmpi,part,${p_{i,{\rm{mass}}}} = {{{\mu _i}} \over {{{\bar \mu }_{{\rm{atm}}}}}}{p_{i,{\rm{part}}}},$(3)

where µi is the molecular mass of volatile i and μ¯atm${\bar \mu _{{\rm{atm}}}}$ is the mean molecular mass of all atmosphere constituents.

In addition, atmospheric erosion can remove H2O from the system, and thus acts as another sink term to the volatile budget, Mimoa$M_i^{{\rm{moa}}}$ (Fig. 1; see Sect. 2.3 for a more detailed description of atmospheric escape). The two processes that remove volatiles from Mimoa$M_i^{moa}$ are described by the following differential equations: dMisoldt=kiFi4πρmrs2 drs dtdMcy stal /dt${{{\rm{d}}M_i^{sol{\rm{}}}} \over {{\rm{d}}t}} = {k_i}{F_i}\mathop {\mathop {4\pi {\rho _m}r_s^2{{{\rm{d}}{r_s}} \over {{\rm{d}}t}}}\limits_ }\limits_{{\rm{d}}{M^{{\rm{cystal}}/{\rm{d}}t}}} $(4) dMH2Oesc dt=4πrp2ϕ1μH2OμH${{{\rm{d}}M_{{{\rm{H}}_2}{\rm{O}}}^{esc}} \over {{\rm{d}}t}} = 4\pi r_p^2{\phi _1}{{{\mu _{{{\rm{H}}_2}{\rm{O}}}}} \over {{\mu _H}}}$(5) dMCO2esc dt=0.${{{\rm{d}}M_{{\rm{C}}{{\rm{O}}_2}}^{esc}} \over {{\rm{d}}t}} = 0.$(6)

We refer to Tables 1 and 2 for a concise description of all relevant parameters.

thumbnail Fig. 1

Schematic depicting the setup of the volatile exchange during initialization (t=0) and run time. For initialization, a surface temperature of 4000 K and a completely molten magma ocean was assumed, where the dissolved volatiles are in balance with the outgassed volatile content, set by the volatile melt fraction, Fi. As the magma ocean solidifies, part of the volatile budget is deposited in the solid mantle. Furthermore, atmospheric escape can remove H2O. These two sink terms thus reduce the amount of a volatile available in the fully coupled MOA system, Mimoa$M_i^{moa{\rm{}}}$. The full overview of all included processes, including radiogenic heating, is shown in Barth et al. (2021, Fig. 1).

2.2.1 Implementation of outgassing during run time

We used the starting conditions for the magma melt fraction of volatile i – that is, Fi(t = 0) (Fig. 1 left) - as input parameters for MagmOc2.0. We used a root-finding algorithm to solve the mass balance equations (Eq. (1)) with respect to Fi for (t = 0) for a given volatile initial mass, Miinitial$M_i^{initial{\rm{}}}$, to find the starting conditions for a given initial magma ocean depth.

In MagmOc1.0, root finding was performed at all time steps to solve for Fi (t) that satisfies the mass balance, which is still relatively efficient for a single outgassed volatile. In MagmOc2.0, we opted instead to operate on the time derivatives of the mass balance equation, dMimoadt${{{\rm{d}}M_i^{moa{\rm{}}}} \over {{\rm{d}}t}}$ , which is more numerically efficient for multiple volatiles. We also took into account outgassing feedback that occurs due to changes in the mean molecular mass of the atmosphere, dμ¯atm dt${{{\rm{d}}{{\bar \mu }_{{\rm{atm}}}}} \over {{\rm{d}}t}}$, when a magma ocean evolves from a CO2 toward an H2O-dominated atmosphere (Bower et al. 2019).

During run-time, the derivatives of the volatile mass fractions are advanced using a set of coupled ordinary differential equation: dFH2O dt=CH2OACO2CCO2AH2OAH2OBCO2CCO2BH2O,${{{\rm{d}}{F_{{{\rm{H}}_2}{\rm{O}}}}} \over {{\rm{d}}t}} = {{{C_{{{\rm{H}}_2}{\rm{O}}}}{A_{{\rm{C}}{{\rm{O}}_2}}} - {C_{{\rm{C}}{{\rm{O}}_2}}}{A_{{{\rm{H}}_2}{\rm{O}}}}} \over {{A_{{{\rm{H}}_2}{\rm{O}}}}{B_{{\rm{C}}{{\rm{O}}_2}}} - {C_{{\rm{C}}{{\rm{O}}_2}}}{B_{{{\rm{H}}_2}{\rm{O}}}}}},$(7)  dFCO2 dt=CCO2BH2OCH2OBCO2AH2OBCO2ACO2BH2O,${{{\rm{d}}{F_{{\rm{C}}{{\rm{O}}_2}}}} \over {{\rm{d}}t}} = {{{C_{{\rm{C}}{{\rm{O}}_2}}}{B_{{{\rm{H}}_2}{\rm{O}}}} - {C_{{{\rm{H}}_2}{\rm{O}}}}{B_{{\rm{C}}{{\rm{O}}_2}}}} \over {{A_{{{\rm{H}}_2}{\rm{O}}}}{B_{{\rm{C}}{{\rm{O}}_2}}} - {A_{{\rm{C}}{{\rm{O}}_2}}}{B_{{{\rm{H}}_2}{\rm{O}}}}}},$(8)

where Ai, Bi, and Ci comprise different components of the complete derivative of the (volatile) mass balance equations. The full derivation is outlined in Appendix A. The sink terms for the solidified mantle (Eq. (4)) and atmospheric erosion (Eq. (5)) are substracted from dMimoadt${{{\rm{d}}M_i^{moa{\rm{}}}} \over {{\rm{d}}t}}$.

We note that these coupled differential equations for outgassing of multiple volatiles differ from those of Bower et al. (2019), because we choose the volatile melt fractions Fi and not their partial pressures pi,part(Fi) as primary variables of integration. We have further verified that MagmOc2.0 with a pure H2O atmosphere yields the same results as MagmOc1.0 (not shown). Furthermore, the benchmarking that we performed for different Earth scenarios with and without mixed atmospheres further confirms the validation of the code (Appendix C).

2.3 Atmospheric escape

Atmospheric escape was calculated by the VPLanet module AtmEsc, which includes XUV photolysis of H2O into H and O. We assume that the hydrogen produced from photolysis escapes into space based on the hydrodynamic escape mechanism described in Barnes et al. (2020) (see also Watson et al. 1981, Zahnle & Kasting 1986, and Luger & Barnes 2015). Hydrogen may escape in one of two ways depending on the composition of the atmosphere. In water-dominated atmospheres, hydrogen atoms are liberated from water molecules where the optical depth is approximately unity, which is close to the exobase. Hence, these atoms can escape if the incident photons carry enough energy for the hydrogen atoms to achieve escape velocity, which is generally true for XUV photons. This escape mechanisms assumes as the limiting factor the amount of photon energy deposited in the atmosphere. In this energy-limited regime, hydrogen atoms can also carry away oxygen produced by photodissociation if the hydrogen escapes at sufficiently high velocity (Hunten et al. 1987). Our energy-limited escape model is identical to the one described in Barnes et al. (2020), Appendix A.

If carbon dioxide and/or oxygen accumulates in the atmosphere, however, there may be fewer water molecules than photolyzing photons near the exobase. In this case, the limiting factor is the availability of water molecules, which must diffuse through the background gas(es) to reach the photolyzing layer of the atmosphere. In our model, we assume escape transitions from the energy-limited regime to this diffusion-limited regime when the atmosphere consists of more than 60% carbon dioxide or oxygen. We adopt the diffusion-limited escape model from Luger & Barnes (2015), which is a hybrid of the models originally presented in Zahnle & Kasting (1986) and Hunten et al. (1987). In our model, the flux is given by Fdiff=(mbgmH)(1Xbg)bbgɡmHkTflow ,${F_{{\rm{diff}}}} = {{\left( {{m_{{\rm{bg}}}} - {m_{\rm{H}}}} \right)\left( {1 - {X_{{\rm{bg}}}}} \right){b_{{\rm{bg}}}}{m_{\rm{H}}}} \over {k{T_{{\rm{flow}}}}}},$(9)

where mb𝑔 is the mass of the background gas (either oxygen or carbon dioxide), mH is the mass of a hydrogen atom, Xb𝑔 is the molar mixing ratio of the background gas a the base of the flow, bb𝑔 is the binary diffusion coefficient for the dominant background gas, k is the Boltzmann constant, and Tflow is the temperature at the base of the flow. Mason & Marrero (1970) provides the values for the diffusion constants as bO=4.8×1019Tflow 0.75,${b_{\rm{O}}} = 4.8 \times {10^{19}}T_{{\rm{flow}}}^{0.75},$(10)

and bCO2=8.4×1017Tflow 0.6.${b_{{\rm{C}}{{\rm{O}}_2}}} = 8.4 \times {10^{17}}T_{{\rm{flow}}}^{0.6}.$(11)

We assume that Tflow = 400 K for all cases (Luger & Barnes 2015). The only difference between this diffusion-limited escape model and the one described in Appendix A of Barnes et al. (2020) is that here we include CO2.

2.4 Vertically extended mixed atmospheres of magma oceans

Another improvement of MagmOc2.0 is its treatment of the atmosphere as vertically extended rather than as a single layer. Following the methodology established in previous studies (Goldblatt et al. 2013; Lichtenberg et al. 2021; Schaefer et al. 2016), we constructed pressure-temperature profiles for a given surface pressure, psur f , surface temperature, Tsur f , and atmosphere composition.

For atmospheres consisting of a single volatile, we assumed a dry adiabat for the lower atmospheric layers. For the upper atmospheric layers, we incorporated latent heat release via condensation, described by d lnT d lnp=R/cp,i(T)     dry${{{\rm{d}}\ln T} \over {{\rm{d}}\ln p}} = R/{c_{p,i}}(T)\,\,\,\,{\rm{dry}}$(12) =RT/Li     condensation, $ = RT/{L_i}\,\,\,{\rm{}}\,{\rm{condensation,}}$(13)

where p is the gas pressure, T is the gas temperature, R is the ideal gas constant, and cp,i(T) is the specific heat capacity of the volatile, i (either H2O or CO2), using the Shomate equation with values obtained from the NIST data base (Cox et al. 1984; Chase 1998). Values for latent heat, Li, were taken from Lichtenberg et al. (2021) and Pierrehumbert (2010).

A condensate becomes thermally stable when the p-T profile intersects the relevant condensation curves. For the H2O condensation curve, we adopted the August-Roche-Magnus formulation described in Alduchov & Eskridge (1996). For pure CO2, we used the Clausius-Clayperon relation. We validated this method by comparing pressure-temperature profiles for H2O and CO2 and Tsurf = 500-2000 K with previous work (Goldblatt et al. 2013; Lichtenberg et al. 2021).

For mixed H2O and CO2 atmospheres, we adopted the multispecies adiabat formulation with condensation of Graham et al. (2021), where we assumed for simplicity again that CO2 and H2O are not removed from an atmospheric layer when condensation occurs (that is, a supersaturation ratio of S = 1). Thus, we assume in a specific atmosphere layer equilibrium between condensation and instant re-vaporization of volatiles. This approach is valid for a hot steam atmospheres with vigorous mixing. We note that, for the calculation of the atmosphere’s pressure-temperature profile, we used, like Lichtenberg et al. (2021), the ideal gas law for the volatiles, which needs to be revisited for very volatile-rich compositions3. However, we first established here a basic framework to expand the boundaries of Solar-System magma ocean simulations by starting with simplified assumptions and offering it as an open-source project to the community. We further note that for the majority of the magma ocean simulations, the radiative properties are set by the water vapor opacities. Water, however, makes the atmosphere in the infrared – that is, for the outgoing long wave irradiation optically thick already for low atmospheric pressures – p > 0.1 bar. Furthermore, we find that the magma ocean solidification occurs, when the planet is in the runaway greenhouse limit. Thus, the cooling of the magma ocean is set by the radiative and thermodynamics properties of upper atmosphere layers, the temperature of which is mostly determined by the latent heat release of water vapor that is not impacted by supercritical conditions at the surface. Also Marcq et al. (2017) have verified that deviations from the ideal gas law are of minor importance for calculating the outgoing long wave irradiation of their magma ocean atmospheres. Supercritical surface layers may impact, however, the dry adiabat and thus the very hot initial conditions of the magma ocean stage. Differences in the initial conditions of the magma ocean, however, were found already for Earth to be of minor importance for the magma ocean lifetime and volatile evolution, as the magma ocean spends the majority of the evolution in the runaway greenhouse regime (Appendix C). In all cases, we utilized a fourth-order Runge-Kutta integrator to compute the pressure-temperature (pgas, Tgas) profile by integrating upward from surface temperature Tsurf = 500–4000 K and psurf = 0.26–26 000 bar4.

Figure 2 (left) shows example profiles for psurf =pH2O+pCO2=260${p_{{\rm{surf}}}} = {p_{{{\rm{H}}_2}{\rm{O}}}} + {p_{{\rm{C}}{{\rm{O}}_2}}} = 260$ bar for different water volume mixing ratios xH2O${x_{{{\rm{H}}_2}{\rm{O}}}}$. We defined the vertically uniform xH2O${x_{{{\rm{H}}_2}{\rm{O}}}}$ as the ratio of H2O partial pressure over the total surface pressure: xH2O=pH2O/psurf ,${x_{{{\rm{H}}_2}{\rm{O}}}} = {p_{{{\rm{H}}_2}{\rm{O}}}}/{p_{{\rm{surf}}}},$(14)

assuming a well-mixed two-component H2O-CO2 atmosphere: pH2O+pCO2=psurf .${p_{{{\rm{H}}_2}{\rm{O}}}} + {p_{{\rm{C}}{{\rm{O}}_2}}} = {p_{{\rm{surf}}}}.$(15)

We stress again that we operate under the assumption of a highly mixed, hot atmosphere, with instant reevaporation of condensates. Thus, we assumed to the first order that almost all water is in the gas phase.

The mixed H2O-CO2 pressure-temperature profiles indicate that water vapor releases significantly more latent heat per volume than CO2, causing the profiles to closely resemble the pure H2O profile even when 90% of the atmosphere is comprised of CO2 (Fig. 2 left). The resulting pressure-temperature profiles for a given H2 O-CO2 content are used as input for RT calculations with the open source, atmosphere model petitRADTRANS (Mollière et al. 2019) described in the next subsection.

thumbnail Fig. 2

Vertically extended pressure-temperature (pgas, Tgas) profiles for psur f = 260 bar and Tsur f = 1000 K and different water mixing ratios, xH2O${x_{{{\rm{H}}_2}{\rm{O}}}}$ (left). The solid black and solid red line denote 100% CO2 and 100% H2O atmosphere composition, respectively. The profiles converge in the upper atmosphere to the condensation curve, where we assume equilibrium between condensation and evaporation of H2O and CO2 (supersaturation ratio S = 1). Change in emission (right) from a pure H2O atmosphere (red line) to CO2-dominated (black line).

thumbnail Fig. 3

Integrated OLR for psurf = 260 bar and Tsurf = 1000 K and different H2O to CO2 content. Mixing CO2 into the steam atmosphere mainly acts to cool the upper atmosphere layers toward the CO2 condensation curve. Only for xH2O<103${x_{{{\rm{H}}_2}{\rm{O}}}}\, < \,{10^{ - 3}}$ (that is, xH2O>0.999${x_{{{\rm{H}}_2}{\rm{O}}}}\, > \,0.999$) will the overall thermal emission increase again as the steep adiabat of a CO2- dominated atmosphere extends into the upper thermally emitting part of the atmosphere.

Table 3

Opacity sources used in this work.

2.5 Radiative transfer in a vertically extended atmosphere

The thermal evolution of the magma ocean planet is determined by the net outgoing emission Fnet , which is defined as the difference between OLR and incoming absorbed stellar radiation (ASR) at the top of the atmosphere: Fnet =FOLRFASR.${F_{{\rm{net}}}} = {F_{{\rm{OLR}}}} - {F_{{\rm{ASR}}}}.$(16)

For the ASR flux, we used FASR=σTeff4,${F_{{\rm{ASR}}}} = \sigma T_{{\rm{eff}}}^4,$(17)

where the planet’s effective or black body temperature, Teff, was calculated as Teff=(1α4σL(t*)4πa2)14,${T_{{\rm{eff}}}} = {\left( {{{1 - \alpha } \over {4\sigma }}{{L\left( {{t_*}} \right)} \over {4\pi {a^2}}}} \right)^{{1 \over 4}}},$(18)

with the bolometric luminosity of the star, L(t*) at stellar age, t*, where t* = t + tini; that is, simulation t plus initial stellar age of 5 Myrs, the semi-major a and the planetary albedo, α, of the planet. In this work, we we assumed two albedos: 0.75 and 0. The albedo of 0.75 is mostly used to facilitate comparison between MagmOc1.0 and MagmOc2.0. However, our assumption of a well mixed atmosphere with immediate reevaporation of condensates is more consistent with a cloud-free atmosphere with low albedo. Moreover, for planets orbiting ultra cool M dwarfs like TRAPPIST-1, very low scattering on top of the atmosphere (albedo= 0-0.1) is predicted (Kopparapu et al. 2013). Thus, we adopted a clear sky albedo (α) of 0 as the new default.

We calculated the OLR with petitRADTRANS (Mollière et al. 2019) using opacities sampled with the correlated k- method for a wavelength resolution of R=1000. We further assume constant H2O and CO2 volume mixing ratios throughout the atmosphere together with the pressure-temperature profiles as outlined in the previous section. The OLR is equal to the integrated emission on top of the atmosphere between 0.2 and 35 microns, similar to the wavelength coverage in Goldblatt et al. (2013).

Similar to previous work (Goldblatt et al. 2013; Lichtenberg et al. 2021; Boukrouche et al. 2021), we emphasize the importance of using the continuum opacities for the dominant greenhouse gases, here H2O and CO2 are vital ingredients to correctly model the atmospheres of rocky (exo-)planets in the habitable zone. The opacity sources and their references are provided in Table 3. We compute the H2O and CO2 k-tables using the HITRAN2020 line list data and broadening coefficients (Gordon et al. 2022), with a constant line-wing cutoff of 25cm−1. These opacities were converted to a format for input into petitRADTRANS as outlined in Chubb et al. (2021). We have verified that for a pure H2O atmosphere, the OLR yields the canonical runaway greenhouse limit of 282 W/m2 for surface temperatures between 500 and 1800 K (Goldblatt et al. 2013). We further note that for CO2 we use an additional continuum opacity compared to Lichtenberg et al. (2021), which leads to an overall reduction in emission of 20% in a pure CO2 atmosphere with psurf = 260 bar compared to Lichtenberg et al. (2021). Already, Marcq et al. (2017) preformed non-gray RT calculations for mixed H2O-CO2 atmospheres on magma oceans. They also retrieve the blanketing effect of the runaway greenhouse limit of 280 W/m2 that is indeed the major factor shaping the magma ocean evolution also for mixed H2O-CO2 atmospheres in the simulations presented here. In fact, Marcq et al. (2017) also point out that the runaway greenhouse limit can be lowered by tens of W/m2 when CO2 starts to dominate over H2O (see also Appendices B and B.4).

The emission curves for psurf = 260 bar and Tsurf for different water volume mixing ratios xH2O${x_{{{\rm{H}}_2}{\rm{O}}}}$ (Fig. 2 right) are very similar to the emission curve of a pure H2O atmosphere for xH2O102${x_{{{\rm{H}}_2}{\rm{O}}}} \ge {10^{ - 2}}$. There are, however, two notable differences: a CO2 absorption band at 4.3 µm appears as soon as CO2 is added and its amplitude increases with higher CO2 abundances. In addition, the overall thermal emission decreases with lower xH2O${x_{{{\rm{H}}_2}{\rm{O}}}}$ because less water per volume is available that can condense out and thus heat the upper atmospheric layers. Consequently, the upper atmosphere layers cool down and emit less flux (see Fig. 2 left).

For xH2O103${x_{{{\rm{H}}_2}{\rm{O}}}} \le {10^{ - 3}}$, CO2 emission begins to dominate, causing the overall emission to increase as xH2O${x_{{{\rm{H}}_2}{\rm{O}}}}$ decreases, eventually exceeding the runaway greenhouse limit. This effect occurs because the hot dry adiabat of a CO2-dominated atmosphere extends up to p ≤ 50 mbar into the emitting atmosphere layers.

A closer inspection reveals that the runaway greenhouse limit for a mixed H2O-CO2 atmosphere with xH2O<1${x_{{{\rm{H}}_2}{\rm{O}}}} < 1$ is restricted to a lower surface temperature regime compared to an atmosphere composed of 100% H2O (see e.g., Goldblatt et al. 2013; Lichtenberg et al. 2021) and further becomes smaller than the canonical value of 282 W/m2. Further details on this phenomenon are provided in Appendices B and B.4.

During the runtime of MagmOc2.0, we do not compute full RT; instead, we construct for a given planet a thermal emission grid (see Table 4 for the setup). As the magma ocean evolves, the model interpolates surface temperature linearly, while surface pressure and water volume mixing ratio are interpolated logarithmically within the range 106xH2O1${10^{ - 6}} \ge {x_{{{\rm{H}}_2}{\rm{O}}}} \le 1$. For xH2O<106${x_{{{\rm{H}}_2}{\rm{O}}}} < {10^{ - 6}}$, thermal emission for xH2O=0${x_{{{\rm{H}}_2}{\rm{O}}}} = 0$ is used, representing a pure CO2 atmosphere. This is done because we interpolate in log-space and thus cannot interpolate to log(0). We also found that for xH2O=106${x_{{{\rm{H}}_2}{\rm{O}}}} = {10^{ - 6}}$, the integrated thermal emission deviates by less than 10% from the that of a pure CO2 atmosphere.

To speed-up computation time and as a “sanity check” for the thermal evolution based on full RT calculations, we have also developed an analytic approximation of the thermal emission of a mixed H2 O-CO2 steam atmosphere, including modifications to the runaway greenhouse atmosphere limit with increasing CO2 content. The equations describing the approximation are outlined in the Appendix B.

Table 4

Definition of the thermal emission grid.

Table 5

Physical and run parameters for TRAPPIST-1 e, f, and g used by MagmOc2.0.

3 Magma ocean evolution of TRAPPIST-1 e, f, and g

We tested MagmOc2.0 for different Earth scenarios (Appendix C), confirming previous research findings on the substantial H2O- CO2 outgassing feedback (Bower et al. 2019). Furthermore, we found that the addition of CO2 has a minor impact on the solidification timescale compared to a pure H2O atmosphere when atmospheric escape is negligible. This similarity between mixed H2O-CO2 and pure H2O simulations is primarily due to the dominance of the runway greenhouse radiation limit for the majority of the magma ocean lifetime. In contrast to the Earth’s magma ocean stage (e.g., Hamano et al. 2013; Barth et al. 2021), atmospheric erosion is not negligible for the potentially habitable TRAPPIST-1 planets e, f, and g.

In this work, we systematically revisit evolution trajectories investigated by Barth et al. (2021) with pure H2O atmospheres for 1–100 TO initial water mass. We focus on simulations with no CO2, an initial CO2 mass equal to 0.3× the initial H2O mass, and an extreme scenario, where the initial CO2 mass is assumed to be equal to the initial H2O mass (see Table 5 for the relevant parameters). We further note that we neglect tidal interactions in accordance with the results by Barth et al. (2021).

The wide range in initial volatile composition encompasses the large uncertainties in volatile content and the H2O-CO2 ratio acquired during planet formation (e.g., Bitsch et al. 2019). These diverse scenarios allow us to assess to what extent CO2 impacts the magma ocean lifetime and volatile distribution on magma oceans with oxidized outgassing in close proximity to an M dwarf star.

We adopted the H2O and CO2 outgassing laws of Nikolaou et al. (2019), which better represent the current understanding that CO2 is already outgassed during the initial magma ocean stage (see Sect. 2.2), whereas H2O only builds up when the mantle begins to fully solidify. Unless otherwise specified, we adopted the RT atmosphere grid model.

thumbnail Fig. 4

Magma ocean evolution for initial H2O of 1 TO, 5 TO, and 100 TO , respectively. All scenarios are for albedo=0. Initial CO2 mass content is scaled relative to the H2O content by a factor of 0, 0.3, and 1, denoted by red, green, and purple lines, respectively. Surface temperature (left) and volatile content evolution (right) are shown.H2O, CO2, and O2 are denoted by solid, dashed, and dotted lines, respectively.

3.1 TRAPPIST-1 g evolution

Following Barth et al. (2021), we selected TRAPPIST-1 g simulations for initial water masses of 1, 5, and 100 TO, respectively, to discuss critical stages in the magma ocean evolution for planets in the habitable zone of TRAPPIST-1. Barth et al. (2021) identified roughly three scenarios for the magma ocean evolution of TRAPPIST-1 g that represent significant stages for understanding the impact of atmospheric erosion on an oxidized magma ocean. We thus explore here the same initial water scenarios for better comparison. To assess the impact of CO2 on the three outlined scenarios (scenario 1–3), we compare the surface temperature and evolution tracks without CO2 to those with CO2 (Fig. 4). We note, however, that Barth et al. (2021) assumed an albedo = 0.75. Here, we focus mostly on simulations with a clear-sky albedo of 0, which is more consistent with our clearsky RT as well as with the low scattering efficiency calculated for planets around M dwarfs (Kopparapu et al. 2013).

We first report the general trends for all scenarios: there is a significant impact on the general volatile distribution (Appendix C.7) in simulations with additional CO2; the magma ocean starts with a CO2-dominated atmosphere, then H2O tends to become the most dominant species as the mantle solidifies. The planet can end its magma ocean stage again with a CO2 dominated atmosphere, when the planet desiccates. The feedback on the volatile distribution between H2O and CO2 during outgassing is mostly evident in changes in CO2 partial pressures. When H2O starts to become dominant during late stage mantle outgassing, CO2 partial pressures drop. If the planet desiccates, CO2 partial pressures rise. These changes generally align with our results for the Earth magma ocean simulations (Appendix C.3), as well as with those derived by Bower et al. (2019).

Bower et al. (2019) also report that the presence of a thick CO2 atmosphere delays the outgassing of H2O during the magma ocean evolution for Earth. We cannot reproduce this claim in our simulations. Instead, we find delayed H2O outgassing only in the volatile-poor “scenario 1” with albedo 0.75, specifically in the extreme case of adding 1 TO CO2 (not shown). Closer inspection reveals that the delay in H2O outgassing arises solely from the slower cooling of the magma ocean and is not attributable to the feedback effect. While we do not find significant changes in time of H2O outgassing, the thermal evolution and volatile distribution on TRAPPIST-1 g can be impacted with CO2, depending on initial water content.

In the “dry” scenario with 1 TO H2O, the atmosphere becomes completely desiccated, which leads to rapid mantle solidification within 7 Myrs with no CO2 and moderate CO2 content. If an extreme amount of CO2 is added, then diffusion limited escape delays desiccation such that complete solidification occurs after 17 Myrs. Furthermore, we find that even in the case of moderate CO2 content, even a slight delay of solidification by 1 Myrs due to diffusion limited escape facilitates moderate abiotic O2 build-up of a few bar. Without any CO2, all abiotically created O2 is deposited in the mantle before complete solidification.

In the “intermediate” scenario with 5 TO, complete desiccation occurs at a much later date compared to the dry scenario; that is, after 70 Myrs simulation time. Moderate CO2 content leads again to a slight delay in desiccation such that 10 bar abiot- ically created O2 can accumulate in the atmosphere compared to the CO2-free simulation for which only 2 bar O2 remains. With extreme CO2 content, complete desiccation is delayed even further by CO2 diffusion limited escape to 120 Myrs. In this simulation, the planet enters the habitable zone of its host star at 76 Myrs with about 1 TO of H2O in the atmosphere (200 bar). Beyond 200 Myrs simulation time, the model is no longer suitable to follow the further evolution, as is evidenced by the plateauing of the surface temperature evolution (see Sect. 2).

For the “wet” scenario with 100 TO, all simulations enter the habitable zone at 76 Myrs with more than 2000 bars of H2O in the atmosphere and the mantle solidifies very late after 150 Myrs. Additional CO2 extends the solidification time by only few % compared to the H2O-free simulation time. Generally, additional CO2 has little effect on the final magma ocean state. The final state is characterized by an accumulation of more than 10000 bars of water in the atmosphere as the mantle solidifies. All abiotically created O2 enters the mantle already during the magma ocean evolution and thus no O2 build-up occurs even after several 100 Myrs. The lack of final O2 build-up beyond solidification may be, however, due to the limitation of the model that is strictly valid “only” until the magma ocean solidification state at 150 Myrs is reached.

A closer look on the evolution of the net fluxes for the three scenarios (Fig. 5) reveals the key processes determining the magma ocean lifetime. For the first, dry scenario, the final thermal evolution is primarily determined by the fate of the water, even in the presence of 200 bar CO2. The mixed H2O- CO2 atmosphere enters the runaway greenhouse radiation limit already after 100 000 years, which reduces the outgoing thermal radiation to less than 300 W/m2 as long as water is present. Once all water has been eroded, the remaining 60–200 bar thick CO2 atmosphere with a 1400 K surface temperature yields net outgoing flux of more than 1000 W/m2 that quickly leads to mantle solidification (Sect. 2.5, Lichtenberg et al. 2021). All initial differences with thermal evolution due to differences in the CO2 content are removed, once the simulations enter the runaway greenhouse limit.

For the second, intermediate scenario, both the evolution of the runaway greenhouse radiation limit and atmospheric erosion determine the mantle solidification time. Here, again the runaway greenhouse radiation limit with less than 300 W/m2 is reached after 100 000 years. Because there is more water in the system compared to the dry scenario, the radiation limit is maintained for at least 70 Myrs. The net flux continuously drops during that time mainly because the stellar luminosity continuously decreases as the host star evolves to the main sequence. We further note that simulations with additional CO2 consistently show a lower runaway greenhouse limit (in particular around 1 Myrs evolution time) by 10s of W/m2 as long as the atmosphere remains CO2 dominated (see Appendix B.4). When the majority of H2O is outgassed during the final stages of mantle solidification (at 70 Myrs), the thermal evolution curves converge for the moderate CO2 and CO2-free simulation until complete desiccation occurs that triggers quasi-instant solidification. For extreme CO2 content, however, diffusion limited escape can prevent desiccation and thus delays solidification to 120 Myrs.

The third, wet scenario with 100 TO H2O is entirely determined by the runaway greenhouse limit. Additional CO2 leads only initially to differences in evolution. Once, the simulations enter the runaway greenhouse limit (after a few 100 000 Myrs), these differences cease to matter. We note here again that the simulations with additional CO2 have a lower runaway greenhouse radiation limit, as long as CO2 remains the dominant atmosphere constituent. All simulations converge, however, once the majority of the water is outgassed after 100 Myrs. Even with extreme CO2 content, solidification occurs only slightly later compared to the CO2-free simulation.

In Barth et al. (2021), the radiation limit was assumed to apply only for surface temperatures cooler than 1800 K, regardless of surface pressure. Within that framework, thick H2O-dominated atmospheres entered the runaway greenhouse radiation limit very late leading to an extension of the magma ocean lifetime. In this work, however, we find that magma oceans with thick H2O dominated atmospheres enter the greenhouse limit for hotter surface temperatures (Appendix B). Thus, we have not reproduced the magma ocean lifetime extension proposed in Barth et al. (2021) for very high volatile content.

In summary, the fate of water effectively dominates magma ocean solidification lifetimes for all cases. CO2 generally only influences the timescale of water erosion. Because the atmosphere keeps more water vapor, in particular for extreme CO2 content, the magma ocean lifetime of TRAPPIST-1 g can be significantly delayed for the dry and intermediate scenarios with 1 TO and 5 TO H2O, respectively.

We caution, however, that we cannot predict how long the 1400 K surface temperatures can be maintained, because the geophysical model terminates the evolution of the mantle at this point. It thus remains to be confirmed if the extreme jump in outgoing flux as the planet desiccates that triggers immediate solidification can be re-captured with other models that follow the mantle evolution beyond the magma ocean solidification state.

thumbnail Fig. 5

TRAPPIST-1 g: Net flux (OLR-ASR) evolution for the magma ocean scenarios shown in Fig. 4; that is, for 1 TO, 5 TO, and 100 TO initial H2O and various initial CO2 mass fractions for an albedo of 0. We note simulations with an assumed albedo of 0.75 (not shown) are qualitatively similar but have a shorter regime with net flux limited by the runaway greenhouse radiation limit.

3.2 Overview of magma ocean evolution in TRAPPIST-1 e, f, and g

We find that additional CO2 has a particular strong impact on the timescale of atmospheric water loss for simulations with dry to intermediate initial water content (≤10 TO H2O). Because desiccation is in our model followed by complete mantle solidification, the impact of CO2 on atmospheric erosion also impacts the magma ocean solidification timescale. In addition, the magma ocean duration changes with different scattering (albedo) assumptions, which may also impact atmospheric water loss. Thus, we first inspect the change in atmospheric desiccation time for TRAPPIST-1 e, f, and g and for our nominal clearsky (albedo=0) and the high albedo simulations (albedo =0.75) (Fig. 6).

In our simulations that assume clear-sky albedo, Fnet is low and thus the magma ocean cools and evolves over longer timescales compared to the high albedo simulations. This difference in cooling timescale by itself does not appear, however, to strongly impact atmospheric mass loss rate as is evident by comparing pure H2O simulations for albedo 0, and 0.75, respectively (Fig. 6, solid lines). The addition of CO2, however, leads to significant changes in atmospheric desiccation.

If a clear sky is assumed, scenarios with extreme CO2 content only lead to complete atmospheric desiccation with very dry initial water content (≤2 TO H2O). For the high albedo simulations, the shorter magma ocean duration compared to the clear-sky simulations appears to limit the impact of CO2. Desiccation is only significantly delayed for TRAPPIST-1 e, even with extreme CO2 content.

The combination of delayed atmospheric water mass loss and low magma ocean cooling rate in the clear-sky case also leads to a significant extension of the magma ocean solidification timescale with intermediate initial water content (5-10 TO H2O)5 (Fig. 7). This is true for TRAPPIST-1 e, f, and g with extreme CO2 content (Fig. 7 dotted lines, bottom left panel) compared to the pure H2O and high albedo simulations (Fig. 7 solid lines, upper left panel). The magma ocean solidification times are so long that the planets can enter their habitable zone before atmospheric desiccation is complete.

For the volatile-rich scenarios (≥ 10 TO H2O) and with clear skies (albedo = 0), TRAPPIST-1 e, f, and g enter the habitable zone with partly molten surfaces and a mixed H2O-CO2 dominated atmosphere, because atmospheric escape is of lesser importance. But even there, more CO2 leads to an extension of mantle solidification time in particular for TRAPPIST-1 e. In these scenarios, the addition of CO2 reduces the runaway greenhouse radiation limit by a few tens of W/m2 (see also Marcq et al. 2017). For simulations with albedo=0.75, the extension of the magma ocean lifetime for the volatile-rich scenarios is less evident and is at most 10% even for TRAPPIST-1 e.

The duration of the magma ocean also shapes the amount of remaining water in the solidified mantle due to atmospheric erosion (Fig. 7, right panels). The impact of atmospheric erosion is already evident by comparing the remaining water fraction for the individual planets. TRAPPIST-1 g can retain the most water because it is farther away from its host star, and thus experiences less erosion compared to TRAPPIST-1 f. TRAPPIST-1 f experiences less atmospheric erosion than TRAPPIST-1 e, and thus retains more water than the latter.

The impact of water mass loss over different timescales explains the differences for remaining water in the solidified mantle when comparing high albedo (short magma ocean phase <100 Myrs) with clear sky simulations (long magma ocean phase, up to 350 Myrs). As has already been pointed out by Barth et al. (2021), the volatiles in a magma ocean planet are embedded in a strongly coupled system. If this system is subject to continuous water erosion for a longer time, then also more water can be removed from the planet, because more water is outgassed from the melt to compensate for the mass loss from the atmosphere. Consequently, for the same planet and the same initial water input, there is more remaining water in the melt in the high albedo simulations compared to clear sky simulations with longer magma ocean lifetimes.

However, only 2–6%6 of the initial water can be retained in the melt in the pure H2O outgassing scenario for albedo 0.75 and 1–5% for albedo 0. Apparently, the impact of extended water erosion over timescales of 100 Myrs and longer does not lead to a drastic reduction in remaining water. This relatively minor impact on the water mantle content with extended magma ocean lifetime indicates that most water is sequestered in the solidified part of the mantle during the very first million years of simulation time, before atmospheric erosion has had time to remove significant amounts of water from the atmosphere and thus from the remaining molten part of the mantle. It takes typically more than a few million years to significantly erode the atmosphere, except for the very dry scenarios with 1 TO H2O.

Adding CO2 tends to increase the percentage of remaining water fraction by up to 2%. For TRAPPIST-1 e, the total amount of water stored in the solid mantle can thus be doubled with extreme initial CO2 content, assuming a clear-sky albedo. The impact of CO2 on the remaining water is so strong, because more CO2 in the system leads to higher water melt fractions for the whole duration of the magma ocean evolution. Consequently, more H2O can be sequestered in the solid part of the mantle throughout the evolution.

In any case, our work shows that it is important to capture the feedback of CO2 on the atmospheric water loss and the magma solidification timescale. Large amounts of CO2 (>1000 bar) can prevent or at least delay complete desiccation for the TRAPPIST- 1 e, f and g planets. We further find that among all investigated planets, TRAPPIST-1 e, orbiting closest to its host star, is more strongly impacted by various processes that shape the magma ocean stage: These are the assumed albedo, which represents here cases with low scattering (albedo=0) versus high scattering of incoming stellar light (albedo =0.75), the reduction in the runaway greenhouse radiation limit with high CO2 abundances, as well as the reduction of atmospheric water loss because atmospheric escape of H2O is diffusion limited with high amounts of CO2 in the atmosphere.

thumbnail Fig. 6

Overview of total atmospheric desiccation times of TRAPPIST-1 e (red), f (green) and g (purple) for a dry composition with 1–10 TO of initial H2O, assuming an albedo of 0.75 (left) and 0 (right). Colored shaded regions denote the time, when the respective planet enters the habitable zone. Solid lines represent scenario with pure H2O atmospheres, dash-dotted lines denote scenarios with added CO2 scaled by 0.3x compared to the initial H2O mass, and dotted lines show scenarios with added CO2 equal to the initial water mass. The solid circles denote for a given H2O-CO2 mass fraction and planet the maximum initial water mass, for which total atmosphere desiccation occurs before the respective planet enters the habitable zone. Please note the change in y-axis scale between the plots.

thumbnail Fig. 7

Overview of magma ocean solidification time (left) and the remaining water in the solid mantle compared to initial water mass in percent (right) for TRAPPIST-1 e (red), f (green), and g (purple), considering various CO2 contents for albedo 0.75 (top panels) and albedo 0 (bottom panels). Solid lines represent scenarios with pure H2O atmospheres, dash-dotted lines denote scenarios with added CO2 scaled by 0.3x compared to the initial H2O mass, and dotted lines show scenarios with added CO2 equal to the initial water mass. Colored shaded regions denote the time, when the respective planet enters the habitable zone.

3.3 Atmosphere extent and interior modeling constraints

To be able to understand the impact of an extended atmosphere on the observed radii of the TRAPPIST-1 planets, we first made use of an interior structure model (Noack et al. 2017) employing look-up tables created with Perple_x (Connolly 2009) for thermodynamic properties of the silicate mantles. For a first- order estimate on the potential chemical composition of the TRAPPIST-1 planets, we use an adapted version of the condensation model presented in Bitsch & Battistini (2020), and employed the measured metallicity of TRAPPIST-1 to derive the likely chemical composition of the star and planets (depending on their distance of the star; see Appendix E for more information on the numerical models). For the silicate mantle of the TRAPPIST-1 planets, and under the assumption of an Earthlike mantle iron number of 0.1 (i.e., a magnesium number of 0.9), our model predicts the following molar composition: 5.35% FeO, 48.15% MgO, 39.09% SiO2, 2.85% CaO, 1.83% Al2O3, and 1.0% Na2O. Residual iron as well as any condensed FeS contributes to the metal core, with a predicted mass fraction of 27%. However, in our interior structure model, we vary the coremass fraction further to obtain realistic interior structure profiles for the TRAPPIST-1 planets that can match their observed radii. For simplicity, in our model we consider only pure iron (see Appendix E). For the innermost planets, b, c and d, a coremass fraction of 27% leads to a model radius of 1.114 Earth radii (compared to an observed radius of 1.1160.014+0.012$1.116_{ - 0.014}^{ + 0.012}$, Ago et al. 2021), 1.098 Earth radii (compared to 1.0970.014+0.012$1.097_{ - 0.014}^{ + 0.012}$) and 0.774 Earth radii (compared to 0.7880.011+0.01$0.788_{ - 0.011}^{ + 0.01}$), and therefore our model radii match the measured radii within the observational error. This suggests that our compositional model is able to correctly predict the planetary composition of the TRAPPIST-1 planets. For the outer planets of the system, the appearance of volatiles adds a degeneracy to our interior structure.

For the building blocks of the outer planets TRAPPIST-1 e, f, and g (and h with almost exactly the same composition as g), our model predicts that the fraction of silicates and metals in the planetary building blocks would decrease to 64 wt-% (e, f) and 57 wt-% (g, h) (see Fig. 8). For e and f, the remainder of the mass is water in the form of ice layers (including high-pressure ice) as well as liquid water. For g and h, volatiles are separated into 32 wt-% H2O as well as 11 wt-% NH4. However, during planet accretion, due to high-energy impacts and melting processes in the interior, we expect the final planetary compositions to be drier. In addition, planetary orbits may change during the planet formation process, leading to an additional uncertainty in accreted planetary building material, and hence the fraction of volatile materials.

Table 6 lists the range of iron fractions and water/ice fractions that would match the observed masses and radii of TRAPPIST-1 e, f, and g while taking into account the star- derived mantle chemical setup. Figure F.1 shows the possible H2O ranges for each planet within error bars. While for all three planets, the measured radii could be explained by a dry composition (with 25.2, 20.4 and 15.9 wt-%, respectively), increasing iron fractions are possible when adding a water/ice layer of increasing extend with increasing distance to the star. It should be noted that here we only considered water ice and did not take into account any contribution of NH4 , which may be abundant in addition to H2O in TRAPPIST-1 g following our compositional model. Given the apparent lower iron fraction in the system compared to the solar system (as suggested also in Unterborn et al. 2018), we only investigate core-mass fractions of up to 50 wt-%. Within this range, the maximum water fraction is 17.1 wt-% and therefore below 20 wt-%, in accordance with Dorn et al. (2018), Unterborn et al. (2018), Agol et al. (2021), Barth et al. (2021) and Acuña et al. (2021), even though here we apply our TRAPPIST-1 adapted compositional model instead of an Earth-like mineralogy.

In Fig. 9, we display how the planetary radius and density of the planet underneath the atmosphere would change when considering extended atmospheres of different compositions and pressures, calculated as in Ortenzi et al. (2020). We first calculate the predicted atmospheric thickness for the measured mass and radii of each planet and calculate the related radius change. To give examples on how the density of the planets would be affected, we subtract the predicted radius change from the planet radius and calculated the increased density of the sub- atmospheric planet. We note that for today’s measured planet masses and radii, the extended steam atmospheres depicted here are not realistic and the density variations should therefore only be seen as exemplary values.

The atmospheric pressure has a logarithmic effect on the thickness of the atmosphere, leading to increasing atmospheric pressures showing only a weak effect for pressures above a few tens of bars. Water steam atmospheres have approximately twice as strong an influence on the atmosphere thickness as the heavier CO2 atmospheres. In addition, the atmospheric temperature strongly impacts the atmospheric thickness, with an almost linear effect (a doubled absolute temperature leads to an approximately doubled atmosphere thickness).

thumbnail Fig. 8

Predicted composition of the building blocks of the different TRAPPIST-1 planets based on the stellar metallicity and the assumption that the planets did not strongly migrate during accretion. Due to secondary processes including collisions and stripping of material, as well as melting and evaporation processes, especially for the outer planets, the final planetary composition is expected to be considerably less volatile-rich than predicted here for the planetary building blocks.

Table 6

Combinations of core mass fractions and water mass fractions for TRAPPIST-1 e, f, and g that match the observed radii from Agol et al. (2021).

4 Results summary

In this study, we introduce MagmOc2.0, a magma ocean model with versatile multi-volatile outgassing and thermal cooling informed by RT calculations in a vertically extended atmosphere. Comparison of different atmosphere models for the Earth with mixed H2O-CO2 content shows that using RT calculation in a vertically extended atmosphere generally yields longer solidification times compared to a gray atmosphere model. The solidification times still remain within one order of magnitude when atmospheric escape is neglected (Appendices C.1 and C.3).

For a pure CO2 atmosphere, however, the magma ocean reaches the solidification surface temperature of 1400 K significantly faster (by more than one order of magnitude) compared to a gray atmosphere model (Appendix C.2). We find that a vertically extended CO2 dominated atmosphere with surface temperatures larger than 2000 K emits more thermal flux on top of the atmosphere than is captured with the original gray model.

Consequently, adding CO2 to a water steam atmosphere leads to faster cooling of the magma ocean during the initial magma ocean stage.

As soon as the runaway greenhouse limit is reached (Tsurf ≤ 1800 K), however, a second thermal evolution stage begins, lasting much longer than the initial stage. Here, the addition of CO2 leads to slower cooling because its presence cools the upper atmosphere and this reduces overall thermal emission compared to a pure H2O atmosphere.

Application to the potentially habitable TRAPPIST-1 planets e, f, and g shows that the albedo is of major importance in determining magma ocean lifetimes. If an albedo of 0.75 is assumed, as for the Earth-like simulations (Appendix C), than the magma ocean lifetime is always shorter than 100 Myrs for TRAPPIST-1 e, f, and g. An albedo of zero, which is probably more appropriate for hot H2O-CO2 dominated atmospheres around M dwarfs (Kopparapu et al. 2013) extends the magma ocean lifetime for all investigated TRAPPIST-1 planets such that they can even enter their respective habitable zone with intermediate to water-rich compositions (T > 10 TO initial water mass). The presence of CO2, on the other hand, results for water-rich scenarios overall only in a small (1–10%) net extension of the magma ocean lifetime. For dry compositions (≤10 TO initial H2O) and large amounts of CO2, CO2 diffusion limited escape can also significantly extend the magma ocean lifetime for TRAPPIST-1 e (albedo 0.75) and TRAPPIST-1 e, f, and g (albedo 0).

The presence of CO2, however, modifies the volatile distribution in all Earth and TRAPPIST-1 e, f, and g simulations. The magma ocean evolves from an initially CO2-dominated atmosphere into a H2O-dominated atmosphere and for water-poor scenarios (≤10 TO H2O) the end state is again a CO2-dominated atmosphere. This shift in atmosphere composition is accompanied by a feedback in outgassing as outlined by Bower et al. (2019). Generally, the addition of CO2 increases the water melt fraction. Larger melt fractions lead further to larger partial pressures of H2O and may also shift the on-set of water outgassing. The impact of O2 is less clear, because in our simulations O2 only accumulates as the magma ocean solidifies. Here, the magma ocean lifetime that can be extended with low albedos and/or diffusion limited escape is important. We thus find that an extension of the magma ocean lifetime with a clear-sky albedo reduces the amount of abiotically created O2 strongly, as already outlined by Barth et al. (2021).

The increase in the water melt fraction due to the H2O-CO2 feedback is particularly evident for TRAPPIST-1 e (Fig. 7, right panels). Larger water melt fractions also increase the amount of water sequestered in the solid mantle. Here, the amount of remaining water can be increased by 100% and more for the clear-sky albedo with equal amounts of H2O and CO2 compared to a pure H2O scenario. However, the fraction of initial water retained in the solid mantle still remains low: only between 1% and 6% of initial water mass can be retained in the mantle.

New interior composition constraints for all TRAPPIST-1 planets reveals that at least the inner planets (b,c, d) are apparently dry and exhibit an iron fraction of 25 wt-%. If the sane iron fraction is assumed, than TRAPPIST-1 e has currently a low water content, consistent with a dry evolution scenario (≤10 initial H2O). We find that for such a dry composition, the TRAPPIST-1 e magma ocean lifetime can be significantly extended due to CO2 diffusion limited escape. The planetary atmosphere may not completely desiccate even with initial water masses as low as 5 TO H2O within the first 280 Myrs of its evolution; that is, until it enters its habitable zone. TRAPPIST- 1 e is thus the ideal planet to test formation theories and the impact of atmospheric erosion on the water budget during the early evolution of rocky planets orbiting M dwarfs.

In this work, we further developed an analytical corrected gray atmosphere model. This model incorporates information from a thermal emission grid obtained via full RT calculations. Compared to the traditional gray atmosphere model, this corrected gray atmosphere represents a significant improvement. It yields sufficiently similar results for all investigated scenarios in this work compared to the RT atmosphere model, including the special scenario for TRAPPIST-1 e. Both methods are computationally efficient, however, the corrected gray model is 10–26 times faster compared to the RT atmosphere model (Appendix D).

thumbnail Fig. 9

Predicted increase in planet radius (in km and %) as well as resulting increase in density in % calculated for the sub-atmospheric planet layers (i.e., core, mantle, and water/ice layers) when considering CO2 or H2O atmospheres of variable average atmospheric temperature and atmospheric pressures up to 1000 bar.

5 Discussion

The magma ocean model MagmOc2.0 that is introduced in this work is designed to facilitate efficient testing of various scenarios and assumptions. This capability is particularly valuable in the era of detailed characterization of rocky planets beyond our Solar System. MagmOc2.0 provides a fast approach to study the evolution of rocky planets in a large parameter space. On the other hand, MagmOc2.0 can benefit from more complex magma ocean and atmosphere models, leading to continuous improvement of this model for better overall understanding of planetary processes.

In this work, specifically, MagmOc2.0 contributes to a better understanding of the potential final atmospheric state and its impact on inferred interior structure for the potentially habitable TRAPPIST-1 e, f, and g planets. This research is crucial for unraveling the complex interplay between planet formation, planetary structure, atmospheric composition and ultimately their potential habitability. TRAPPIST-1 e may be very volatilepoor compared to TRAPPIST-1 f and g if we assume an iron mass fraction of 27 wt% as for the inner planets. TRAPPIST- 1 f and g, on the other hand, must have formed volatile-rich to explain large water fractions that we estimate for the same iron mass fraction, resulting potentially in thick CO2 atmospheres that may prevent the existence of surface liquid water.

If TRAPPIST-1 e is indeed (partly) desiccated, then it may have formed close to the water-ice line, during which the planet migrated rapidly inward, while f and g formed exterior to the ice line and migrated later. Alternatively, the planets did not undergo significant migration, instead the water ice line shifted during the disk evolution outward starting from a location near TRAPPIST- 1 e, leading to the same outcome (Bitsch et al. 2019).

5.1 Mutual benefit between versatile MagmOc2.0 and other magma ocean models

The investigation of mixed outgassing on magma oceans in the habitable zone of TRAPPIST-1 underscores the potential synergies between magma oceans models with a more comprehensive treatment of geophsical processes, more complex models that capture the geochemistry more completely and simpler frameworks such as VPLanet/MagmOc2.0 that can explore a broad range of initial volatile content and the impact of various treatments of atmospheric and outgassing processes. For instance, the incorporation of feedback between H2O and CO2 outgassing was motivated by the magma ocean model results of Bower et al. (2019). Despite employing different differential equation and a simpler geophysical model, we generally reproduce the findings of Bower et al. (2019) for Earth (Appendix C.7).

Comparison between results from the Earth magma ocean model of Elkins-Tanton (2008) and MagmOc2.0 further yields agreement in volatile distribution when the same outgassing laws are used that suppress CO2 outgassing (Table C.1). However, when modern outgassing laws are applied, there is less agreement in volatile distribution, in particular for CO2. Suppressed CO2 outgassing leads to 10–100 times higher CO2 melt fractions at the end of the magma ocean stage compared to a magma ocean evolution, where the majority of CO2 is outgassed already at the beginning (Table C.3). Thus, MagmOc2.0 can shed light on how initial outgassing conditions can alter the volatile distribution in diverse magma ocean models.

Application of MagmOc2.0 to TRAPPIST-1 e, f, and g reveals that additional CO2 also significantly influences these planets’ volatile distributions, resulting in increased abiotic O2 build-up and a higher remaining water fraction in the solid mantle at the end of the magma ocean stage. Furthermore, we find that the inclusion of CO2 does not consistently lead to delayed H2O outgassing, as inferred by Bower et al. (2019) for an Earth-like scenario. Instead, the onset of H2O outgassing may be delayed, start even earlier, or remain unaffected compared to simulations without CO2. In our model, it is the radiative properties of the mixed H2O-CO2 atmosphere at the time of mantle solidification that primarily determine when H2O outgassing occurs. This insight is facilitated by exploring more volatile scenarios with different atmosphere models (Appendices C.4 and C.5).

We also demonstrate that MagmOc2.0 can identify unexpected feedbacks between atmospheric erosion and mantle solidification due to diffusion limited escape in mixed H2O-CO2 outgassing. For TRAPPIST-1 e with 10 TO H2O initially, the reduction of the runaway greenhouse radiation limit by 10s of W/m2 as water erosion from the atmosphere substantially prolongs the magma ocean stage and also diminishes water sequestration in the mantle.

We note that this special case was not identified in a similar magma ocean study with mixed H2O-CO2 outgassing for the TRAPPIST-1 planets conducted by Krissansen-Totton & Fortney (2022). Their model did not consider the reduction of the runaway greenhouse limit with CO2, which is a prerequisite for this scenario to occur. Additionally, it does not account for changes in the volatile distribution due to the evolution in atmospheric composition. Apart from these differences, the model of Krissansen-Totton & Fortney (2022) yields similar magma ocean solidification lifetimes for TRAPPIST-1 e, f, and g compared to our study.

5.2 The concept of solidification time

When comparing different magma oceans simulations, it is important to discuss the concept of ‘solidification time’ with care. In MagmOc, the magma ocean effectively ends with the convectively dominated viscosity regime. This point is reached when the critical the surface melt fraction ψ = 0.4, following Lebrun et al. (2013) with surface temperatures of 1650 K. At this point 99% of the planetary radius is solidified. In principle, the remaining melt is no longer well mixed and in full equilibrium with the atmosphere. For the last % of radius solidification, thus, MagmOc switches to a high viscosity treatment of the mantle with no further outgassing but with further surface cooling and ongoing atmosphere erosion. Complete solidification then occurs when the solidification radius is equal to the planetary radius, which is in our model typically for surface temperatures of 1400 K (see Sect. 2). Thus, solidification temperatures in our model effectively comprises between 1650 and 1400 K.

Since our magma ocean effectively ends at ψ = 0.4 and complete solidification occurs shortly thereafter, MagmOc2.0 stops earlier and at higher surface temperatures of 1400 K compared to the model of Lichtenberg et al. (e.g., 2021). In this model solidification is defined when the rheological front reaches the surface with ψ = 0.22. Since these authors present the full thermal evolution for pure H2O and CO2 atmospheres, respectively, we can still compare MagmOc2.0 to their model state at ψ = 0.4 which is reached in their models with surface temperatures between 1700 K and 1500 K, depending on atmosphere composition, with a similar strong temperature decrease in this regime as also observed in our models (Lichtenberg et al. 2021, Fig. 5). Thus, we identify general agreement for the thermal evolution of our Earth magma ocean model and that of Lichtenberg et al. (2021) (Appendix C.2).

Our effective solidification temperature of 1400 K is similar to that of the model of Krissansen-Totton & Fortney (2022) and only 100–150 K cooler compared to the solidification temperatures in the models of Nikolaou et al. (2019) and Hamano et al. (2013), respectively. In the magma ocean model of Elkins-Tanton (2008) solidification is assumed when 98% of the planet’s radius is solid, which aligns well with our choice of ψ = 0.4 with 99% mantle solidification, allowing comparisons between MagmOc2.0 and their Earth simulations (Appendix C).

5.3 The impact of clouds

Efficient cooling of a magma ocean requires a large net flux, FNet, leaving the top of the atmosphere, which is the difference between OLR and absorbed stellar atmosphere (ASR). Our simulation results show that mainly two effects determine the magma ocean lifetime: The presence of water vapor and scattering on top of the atmosphere via Rayleigh scattering and clouds, encapsulated in our work via the bond albedo α. The presence of water “sets” the OLR, the albedo α sets the ASR.

When comparing simulations with α = 0.75 (Venus-like scattering) to α = 0 (cloud-free scattering), we find that the cloud-free assumption yields a much larger magma ocean lifetime for all planets even without adding CO2. The addition of CO2 is in this respect of minor importance, it may prolong, however, the timescale of water loss as outlined in the next section. We further note that Marcq et al. (2017) propose in their model that the inclusion of clouds lowers the runaway greenhouse radiation limit to 197 W/m2, whereas for the majority of our simulations the limit of 280 W/m holds unless the atmosphere is highly water-depleted. Kopparapu et al. (2013), on the other hand, argue that water clouds can be neglected for the calculation of the OLR.

In the clear-sky simulations, the magma ocean lifetime is so prolonged that all planets may enter the habitable zone7 before complete mantle solidification for water-rich composition (>10 TO H2O). This result is of outermost importance because many geophysical models assume by default that water can condense out if the incoming stellar flux is low enough. While this is a safe assumption for Earth, this is apparently no longer true for planets that orbit an M dwarf star.

Instead, volatile-rich planets can be ‘nominally’ in the habitable zone, but condensation on the surface is not yet possible. In addition, atmospheric water loss continues for several hundred years after the habitable zone entrance for TRAPPIST-1 e. In principle, this result is similar to that of Turbet et al. (2021) who showed that even in the Solar system, Venus probably never had a chance to form a surface ocean because thick water clouds prevented efficient cooling of the surface. On the other hand, Yang et al. (2014) showed that for tidally locked planets around M dwarf stars, the high albedo from water clouds forming predominantly over the dayside of an Earth-like planet may allow the inner edge of the habitable zone to shift closer to its host star.

In this work, we simplified the problem by adopting two extreme values to outline the scope of the impact. In our model-setup, however, the cloud-free assumption is the most physically consistent, because we assume efficient reevapora-tion after condensation (that is, with a supersaturation ratio of S=1). Furthermore, Kopparapu et al. (2013) point out that water clouds in a water dominated atmosphere have only a minor impact because of the high intrinsic optical thickness of water vapor in the infrared. Conversely, Marcq et al. (2017) find that the runaway greenhouse radiation limit is lowered to 200 W/m2, when water clouds are included. CO2 clouds, on the other hand, may allow for strong scattering. This effect may be relevant in our framework for desiccated CO2 atmospheres. However, even for relatively low water abundances of xH2O = 10−3, the upper atmosphere is warmed by the latent heat release of water condensation such that CO2 clouds are not expected to form (see Fig. B.2). In any case, clearly more work is warranted to assess the impact of clouds in particular to investigate if and when the TRAPPIST-1 e, f, and g planets can form surface oceans after the magma ocean stage. This will require further work on the details of precipitation and evaporation of water on rocky planets around M dwarfs. Already for ‘Earth-like’ conditions with temperate temperatures, different cloud models can yield very different results (Sergeev et al. 2022). Furthermore, Sergeev et al. (2024) point out that resolving moist convection on the dayside of potentially habitable planets requires a higher resolution than is adopted so far. In any case, clearly more work is needed to explore the role of clouds on the outgoing long wave irradiation and albedo for the magma ocean stage on rocky planets around red dwarfs.

5.4 Volatile evolution

A magma ocean evolution with a mixed H2O-CO2 atmosphere can exhibit feedback between H2O and CO2 outgassing that shapes the overall volatile distribution; that is, the distribution of H2O, CO2, and O2. In addition, in a CO2-dominated atmosphere that is prone to emerge at the beginning and the end of the magma ocean stage, CO2-diffusion limited escape can limit water loss, and thus prevent complete desiccation on the investigated planets for a longer time. In the following subsections, we discuss the feedback impact for each volatile species in more detail.

5.4.1 H2O

The fate of water during the magma ocean evolution on rocky planets around M dwarf stars is of utmost importance for constraining their habitability. If there is too little water (<10 TO H2O), the planets completely desiccate during the magma ocean stage with only a few percent initial water mass remaining in the mantle.

We find that the mantle is driest for our TRAPPIST-1 e simulations compared to simulations of TRAPPIST-1 f and g because planet e is more strongly affected by UV-photolysis of H2O. Additionally, the mantle tends to become even drier when the magma ocean stage is prolonged, which is the case for the clear-sky (albedo=0) simulations. The desiccation of the mantle of TRAPPIST-1 e can be, however, compensated by large amounts of CO2 as is evident by comparing the remaining mantle water inventory for the albedo 0 and 0.75 evolution scenario with assumed initial CO2 mass being equal to the initial H2O content (Fig. 7). Despite much longer desiccation times, the remaining water content remains of similar order.

A desiccated evolution scenario is further of particular interest for TRAPPIST-1 e because its composition is apparently consistent with a dry Earth-like H2O content (<<1 wt-%) (see Sect. 3.3). Thus, here our evolution scenarios with an initial water mass of 1–10 TO, estimated for Earth (see Appendix C), are particularly relevant. They show that the atmosphere of TRAPPIST-1 e may retain 5 TO water until the star’s luminosity dropped such that it enters the habitable zone (Fig. 6). Similarly, even if the atmosphere is desiccated, up to 4% of the initial water mass could be stored in the mantle, in principle enabling a secondary surface water reservoir to build up (Godolt et al. 2019). Since such a relatively wet mantle would require large amounts of CO2, the surface may be Venus-like and thus too warm for surface liquid water. On the other hand, there may be several other ways to retain more water in the melt like delayed H2O outgassing (Ikoma et al. 2018; Bottinga & Javoy 1990; Lensky et al. 2006) and a basal magma ocean that doesn’t require huge amounts of CO2 (see also Sect. 5.4). Conversely, the outer planets TRAPPIST-1 f and g may have the opposite problem of having too much water for a habitable surface ocean. A very thick water ocean not only dilutes nutrients, the further enrichment from the crust may also be suppressed due to the high pressures at the bottom of the ocean (Noack et al. 2016).

The result that an extension of the magma ocean stage generally leads to a drier planet is in marked contrasts with the findings from the 0D magma ocean model by Moore et al. (2023). Figure 1 illustrates the reason for this discrepancy: the amount of dissolved water in the magma ocean, our primary variable of integration, FH2O${F_{{{\rm{H}}_2}{\rm{O}}}}$, is in balance with the outgassed atmosphere. As has previously been pointed out by Barth et al. (2021), as long as such a balance exists, atmospheric escape will drive additional outgassing from the magma ocean to establish a new equilibrium. It is thus mainly the reduction of water loss in the presence of a thick CO2 atmosphere that can prevent atmosphere desiccation, in particular on TRAPPIST-1 e. It is then the retention of significant amount of water vapor in the atmosphere (>100 bar) that extends the magma ocean lifetime, and not the other way around.

We note that there are alternative mechanisms to ‘save’ water from atmospheric escape during the magma ocean stage apart from CO2 diffusion limited escape. In MagmOc2.0, water is sequestered by partitioning 1% of available water mass in the solid mantle during solidification (kH2O)$\left( {{k_{{{\rm{H}}_2}{\rm{O}}}}} \right)$, consistent with values for relevant minerals such as lherzolite and peridotite (Elkins-Tanton 2008; Johnson & Dick 1992). This water is saved from atmospheric escape because it is no longer part of the fully coupled volatile system. Nevertheless, the assumption of a constant kH2O${k_{{{\rm{H}}_2}{\rm{O}}}}$ at the bottom of a highly convective magma ocean may be overly simplistic (Ikoma et al. 2018). Additionally, whether crystals settle at the bottom of the magma ocean (fractional solidification), as is assumed in this model, or remain suspended in the melt (batch solidification) may further impact how much water is incorporated into the solid mantle. Moreover, the role of minerals capable of incorporating higher water mass fractions like phyllosilicates has not been accounted for yet (Herbort et al. 2020). Lastly, the efficiency of outgassing in a highly convective magma ocean may be reduced compared to ‘static’ laboratory experiments (Salvador & Samuel 2023).

In this study, we identify an H2O-CO2 feedback as one mechanism that increases the amount of water in the mantle. Adding CO2-outgassing leads to changes in the mean molecular weight of the atmosphere during the magma evolution (Nikolaou et al. 2019; Bower et al. 2019; Lichtenberg et al. 2021): Initially, the magma ocean begins its evolution with a CO2-dominated rather than a H2O-dominated one because CO2 is less soluble in the melt compared to H2O. One major consequence is that the water melt fraction FH2O${F_{{{\rm{H}}_2}{\rm{O}}}}$ increases in a magma ocean with a CO2-dominated atmosphere, leading to increased H2O out-gassing with higher H2O partial pressures of H2O at later times. Simultaneously, because there is more water in the melt, a greater amount of water can be partitioned and sequestered into the solidifying mantle. Generally it holds that the more CO2 is in the MOA system initially, the more H2O remains in the mantle. Consequently, the addition of CO2 results in 15–50% more remaining water in the mantle at the end of the magma ocean stage for all investigated cases.

Other possibilities to retain more water in the melt include the delay or suppression of H2O outgassing (Ikoma et al. 2018) through mechanisms such a bubble nucleation (Bottinga & Javoy 1990; Lensky et al. 2006). Furthermore, Bower et al. (2022); Maurice et al. (2017) that use more complex models pointed out that the mode of crystallization can significantly increase water retention. In addition, Maurice et al. (2017) suggest that Rayleigh-Taylor instabilities may replenish the mantle with volatiles from the deeper interior. Furthermore, Hier-Majumder & Hirschmann (2017) suggest that volatile-rich may be trapped at the solidification front. Another mechanism to retain water in the melt may also be the early stratification of the magma ocean (O’Rourke 2020; Samuel et al. 2023) such that part of the magma ocean solidifies downward, resulting in a basal magma ocean. If stratification occurs before the majority of H2O is out-gassed, then a basal magma ocean could, in principle, retain large amounts of water over the course of billion of years (Moore et al. 2023). Therefore, it is very likely that we underestimate the amount of water that can be retained in the planet. Still, we emphasize that increasing the amount of water with the addition of CO2 will in all these cases also increase the amount of volatiles that can be stored in the planet. It is thus highly important to diagnose in the near-future the CO2 atmosphere budget on the potentially habitable TRAPPIST-1 planets in the future.

Here, our magma ocean simulations can help as they also incorporate atmospheric escape. Most importantly we find that the inclusion of CO2 can significantly reduce atmospheric escape compared to a pure H2O atmosphere, as already pointed out by Moore et al. (2023). On the other hand, Joule heating, which is not considered in this work, may instead lead to increased atmospheric erosion compared to our model, at least for TRAPPIST-1 e (Cohen et al. 2024).

5.4.2 CO2

The addition of CO2 has a profound impact on the overall volatile budget, as it is much less soluble in magma and less prone to be partitioned into the solid mantle than H2O (Table 1). Furthermore, CO2 can create a ‘diffusion barrier’ for atmospheric water loss as outlined above for water-poor compositions (<10 TO H2O) and in particular for TRAPPIST-1 e at the inner edge of the habitable zone of its host star. Still, even in this case, generally large amounts of CO2 are required for a significant impact of the magma ocean; that is, CO2 mass ≥0.5× initial water mass (see detailed grid). Such extreme scenarios can lead to very thick (≥103 bar) CO2-dominated atmospheres. In very volatile-rich cases (>>10 TO H2O and CO2), the magma ocean may not even completely solidify, retaining surface temperatures above 1400 K (Fig. 3).

In contrast, Krissansen-Totton & Fortney (2022) assume less extreme values of additional CO2, with partial pressures of up to 103 bars, allowing at least the crust to fully solidify. Another difference is that the runaway greenhouse radiation limit depends in our model on the water volume mixing ratio and can be lowered by 10s of W/m2, when CO2 is added (Appendix B.4). Instead, Krissansen-Totton & Fortney (2022) assume that all water condenses out of the atmosphere onto the solidified crust of TRAPPIST-1 e, f, and g, once the ASR drops below the constant runaway greenhouse radiation limit of 282 W/m2. Due to the presence of liquid surface water in their model, surface weathering and the carbon cycle remove hundreds of bars of atmospheric CO2. In our model, the presence of liquid surface water is questionable for the thick remaining CO2 atmospheres, which appears to be a reasonable outcome for the outer planets TRAPPIST-1 f and g. Furthermore, in contrast to the model of Krissansen-Totton & Fortney (2022), the magma ocean stage in our cloud-free (albedo=0) model can be extended well into the habitable zone. Thus, we caution to automatically assume rainout of water from the atmosphere, only due to low stellar luminosity.

We further note again that the role of clouds (Turbet et al. 2021; Yang et al. 2014) is of outermost importance (see Sect. 5.3). If water cannot condense out on the surface after the magma ocean stage, then the remaining H2O-CO2-O2 atmospheres would only be modified by further atmospheric loss. Constraints of CO2-H2O dominated atmospheres on TRAPPIST-1 e, f, and g may thus provide insights into the abundances of volatiles that were present during the magma ocean stage and on conditions to form surface liquid water, as also pointed out by Krissansen-Totton (2023).

5.4.3 O2

In our work, atmospheric O2 begins to accumulate as soon as more than 5 TO H2O initial water are in the system for albedo=0.75, reaching several hundred bars of partial pressure for 100 TO initial H2O, as has already been outlined by Luger & Barnes (2015) and Barth et al. (2021). O2 can only be stored in the FeO-buffer in the mantle, which remains accessible only as long as the magma ocean is not solidified. For the high albedo simulations, the magma ocean solidifies within 100 Myrs and thus tends to yield large partial pressures of O2. Conversely, for magma ocean lifetimes larger than 100 Myrs, fewer than 50 bars of O2 build up abiotically. Instead, more O2 is stored in the mantle (see detailed grid). This grid also shows the tendency of more O2 being stored in the mantle for extended magma ocean lifetimes even for simulations with a high albedo of 0.75 and initial water masses between 5 and 10 TO.

Krissansen-Totton & Fortney (2022) find in their model that even if there is significant abiotically build-up of oxygen, up to 100 bar O2 can be removed due to reduced outgassing later in the evolution and crustal oxidation. However, the authors also find that in some scenarios 100 bar of atmospheric O2 may remain on TRAPPIST-1 e, f, and g.

Both models neglect, however, possible O2 sequestration in the iron-core (Wordsworth et al. 2018). Moreover, abiotic O2 build-up occurs when the majority of H2O is outgassed - that is, when the mantle is mostly solidified – already limiting the exchange of volatiles between the atmosphere and the mantle. Therefore, it is questionable if removal of atmospheric O2 via the even deeper iron-core can play a significant role this late in the magma ocean evolution.

5.5 The need for tighter link to planet formation models

The prospects of obtaining reliable atmospheric constraints for rocky exoplanets in the near future holds great promise. Establishing a coherent link between planet formation, rocky planet evolution (including the crucial magma ocean stage), interior models and atmospheric composition constraints has the potential to unveil the complete history of rocky exoplanets. Studies of the TRAPPIST-1 system are particularly promising because it is postulated to be only weakly modified by late accretion (Raymond et al. 2022).

The water content of the Trappist planets needs to have been delivered during the protoplanetary disk phase. Planet formation during this initial stage can be driven either by pebble accretion (e.g., Ormel et al. 2017; Schoonenberg et al. 2019; Bitsch et al. 2019) or planetesimal accretion (e.g., Coleman et al. 2019), where both scenarios are capable of roughly reproducing the observed structure of the system. Both scenarios require planet migration as a key ingredient due to the observed resonance configuration of the system (Pichierri et al. 2024). This begs the question of whether the planets formed interior or exterior to the water and CO2 ice lines in the disk, and how their water content is influenced by their formation location and disk cooling. Consequently, planet formation theories can give a range of initial water contents for the TRAPPIST-1 planets, depending on the exact stellar composition of Trappist-1 and on how efficient atmospheric recycling of incoming water rich pebbles is (Müller et al. 2024a).

In the Schoonenberg et al. (2019) scenario, the planets are formed originally at the water ice line and the migrate inward, which sets their water content. The migration direction and speed in comparison with the growth rate then sets the final water content of planets forming exterior to the water ice line (e.g., Bitsch et al. 2019). Planets that start their formation exterior to the water ice line, but then migrate inward very quickly will have a low water content despite their water-rich formation location. Planets that finish their accretion in the outer regions of the disk and that then only migrate across the water ice line once their accretion is complete will have the maximal water content that the system allows, which should be around 35–50% for solar composition, depending on the exact chemical model and stellar composition (Bond et al. 2010; Bitsch & Battistini 2020; Cabral et al. 2023). The exact volatile ratio could also be affected by atmospheric recycling of incoming water ice rich pebbles, where water fractions above 15% are hard to achieve. A high volatile mass fraction appears to be the currently favored scenario for TRAPPIST-1 g inferred from interior models (Unterborn et al. 2018; Barth et al. 2021; Raymond et al. 2022).

To assess the initial CO2 content of the TRAPPIST-1 planets, the CO2 evaporation line needs to be considered, which is located further away from the host star due to the lower evaporation temperature of CO2 compared to water. This difference inherently implies that a planet forming in the outer regions of the disk that accretes CO2 ice, will also accrete water ice. The ratio of the water to CO2 content is then set by the initial CO2 to water ratio of the system and also by the growth and migration behavior of the planet exterior to the CO2 line, similar to that described above for the water content.

In addition, the effect of pebble evaporation and condensation can dramatically change the water and CO2 content of growing planets. As pebbles drift inward and cross evaporation fronts, they release their volatile content into the gas phase (e.g., Piso et al. 2015; Aguichine et al. 2020; Schneider & Bitsch 2021), enriching the disk with vapor. This vapor can then diffuse outward and re-condense, resulting in very large fractions of the corresponding molecule in the solids around the evaporation front (Ros & Johansen 2013). This effect can potentially explain the low water to CO ratio of the Comet C/2016 R2 (Mousis et al. 2021). This effect occurs for all volatiles, indicating that planets that form close to the CO2 evaporation front could have very large CO2 to water ratios.

Here, we build upon previous work (Bitsch & Battistini 2020) to constrain the composition of the building blocks of the TRAPPIST-1 planets thus to derive constraints on the iron fractions 25 wt% and thus on the volatile composition of the inner planets. Based on our results, TRAPPIST-1 b, c, and d already formed volatile-poor and are thus not expected to currently exhibit any atmosphere.

In contrast to that TRAPPIST-1 e, f, and g started with volatile-rich building blocks, of which they may have lost a significant fraction already before the magma ocean phase. However, of these planets TRAPPIST-1 e is particularly prone to complete desiccation during the magma ocean evolution - in particular, when the already long magma ocean stage of 50 Myrs is extended even further. Interior models confirm that currently measured radii and masses are compatible with a dry interior and (moderately) thick CO2 atmosphere of 100 bar. TRAPPIST-1 g, on the other hand, may exhibit up to 17 wt% of water, which indicates an evolution that has allowed it to retain significant fractions of volatiles until today.

This work thus shows that a strong link between CO2 and H2O content during formation that is modified during the rocky planet evolution justifies the need for a coherent chain between long term planet evolution models, current atmosphere characterization efforts of rocky exoplanets, and planet formation theories.

The TRAPPIST-1 planetary system may be key here to compare the outcome of three different scenarios in volatile accretion during formation and their subsequent loss or retain-ment via JWST observations and future missions like LIFE, ARIEL. PLATO may be highly useful to constrain extended steam atmospheres (up to 10%) in young rocky exoplanets that are at the end of their magma ocean stage (Turbet et al. 2019; Schlecker et al. 2024).

6 Conclusion

MagmOcV2.0 can provide novel insights into feedback between H2O and CO2 for outgassing in magma oceans on Earth-mass rocky planets in the habitable zone around TRAPPIST-1. Together with modeling of the refractory and volatile abundances during formation and modeling of the current interior composition, we derived the following key insights for the TRAPPIST-1 system:

  • A compositional model adjusted by the measured metallicity of TRAPPIST-1 yields a dry composition with an iron fraction of 27 wt% for the inner TRAPPIST-1 planets (b, c, and d), which is compatible with the measured radii and masses from Agol et al. (2021). This result is also compatible with recent JWST measurements that may indicate the absence of a substantial atmosphere on TRAPPIST-1 b and c (Zieba et al. 2023; Greene et al. 2023).

  • Assuming an iron fraction of 27 wt%, interior structure modeling for TRAPPIST-1 f and g yields a substantial water fraction of 2.3 wt%, and 3.8 wt%, respectively. This result aligns well with a volatile-rich formation environment that is inferred for these planets in this work and also aligns with formation modeling (see also Barth et al. 2021; Raymond et al. 2022; Miguel et al. 2020; Schoonenberg et al. 2019).

  • Without additional sink terms, thick (>1000 bar), mixed CO2-H2O atmospheres remain at the end of the magma ocean stage for volatile-rich scenarios ( >10 TO initial water mass) that are favored for TRAPPIST-1 f and g.

  • More than 1000 bar of CO2 retained after the magma ocean stage may contribute up to 2% to the total planetary radius if water vapor is removed after the magma ocean stage. An H2O-dominated atmosphere would comprise even up to 10% of the planetary radius, as has already been pointed out by Turbet et al. (2019); Schlecker et al. (2024). While such extended steam atmospheres are incompatible with expectations for the evolved TRAPPIST-1 system, they may be identifiable for younger systems with future space missions like PLATO.

  • In all cases, we find that adding CO2 increases the H2O mass fraction in the melt. Thus, between 15 and 100% more water can be retained in the solidified crust, which results in the sequestration of between 2 and 6% of the initial water mass in the solidified mantle for TRAPPIST-1 e, f, and g.

  • TRAPPIST-1 e may have formed, like TRAPPIST-1 f and g, in a volatile-rich formation environment. If, however, an iron fraction of 27 wt% is assumed, as was derived for the inner planets, then its current radius and mass is consistent with a more desiccated evolution scenario (water content of 0.60.6+2.1$0.6_{ - 0.6}^{ + 2.1}$ wt%). The values also include Earth-like water content (≤ 10 TO H2O).

  • TRAPPIST-1 e magma ocean simulations with Earth-like water content are very sensitive to assumptions about scattering (albedo) as well as to changes in atmospheric water loss brought about via diffusion-limited escape in a CO2 background atmosphere. Here, CO2 can prevent, or at least delay, complete atmospheric water loss, even for a long magma ocean stage of 350 Myrs. The atmospheric end result for TRAPPIST-1 e after the magma ocean stage could thus be a either a dry CO2 or a mixed H2O-CO2 atmosphere with surface pressure of several 100 to 1000 bar. The mixed atmosphere could, in principle, further evolve into a more Earth-like state if water condenses out of the atmosphere and if up to 1000 bar of atmospheric CO2 can be sequestered in the mantle, as is proposed, for example, by Krissansen-Totton & Fortney (2022).

  • TRAPPIST-1 e thus emerges as a “Rosetta stone” for deciphering formation as well as evolution processes for rocky planets in the habitable zone of active M dwarfs.

7 Outlook

In this work, we have illustrated how combined H2O-CO2 outgassing can alter the magma ocean evolution and volatile distribution on rocky planets in the habitable zone of their Mdwarf host stars, using TRAPPIST-1 e, f, and g as examples. However, future investigation is necessary to test the impact of other factors that we either neglected or assumed to be constant. For example, we kept the albedo fixed at α = 0.75 and α = 0, respectively), neglecting albedo variations as the atmosphere evolves (Pluriel et al. 2019). We also assumed in the vertically extended atmosphere immediate reevaporation after condensation, which is a strong simplification, in particular toward the end of the magma ocean stage. Additionally, exploring the evolution of the mantle’s redox state (Katyal et al. 2020) will be important because it may result in outgassing of reduced volatiles, such as H2 and CO (Ortenzi et al. 2020; Deng et al. 2020).

H2 is proposed to be dissolved into the mantle and core during the planetary accretion state, establishing a more direct connection between magma ocean and formation models (Johansen et al. 2023; Young et al. 2023, e.g.,). A significant atmospheric H2 content during the magma ocean phase may lead to even more substantial outgassing feedback effects, as is tackled in this study, because it would significantly reduce the mean molecular weight of the atmosphere. Moreover, H2 is a potent greenhouse gas that can significantly prolong the magma ocean stage (Lichtenberg et al. 2021) even more. At the same time, it can allow for surface liquid water on rocky planets at the outer edge of the habitable zone, such as Mars (Pahlevan et al. 2022). Our versatile outgassing formalism, coupled with the RT atmosphere model, allows for the adoption of other outgassing scenarios.

While our current setup lacks a thermal boundary layer (as in Barth et al. 2021), future investigations are planned that will incorporate now more advanced geophysical VPLanet modules, like ThermInt, for the exploration of an Earth-like planet with a stagnant lid configuration (Driscoll & Bercovici 2014). Additionally, plans include integration of the “mush stage” described by Lebrun et al. (2013) after ψ = 0.4 is reached. Similarly, we recognize the necessity for a modification of the atmospheric equation of state that is in this work based on the ideal gas law at the surface for particularly volatile-rich evolution scenarios; for example, those with an initial volatile mass much larger than 10 TO, which is psur f > 1000 bar.

Still, our work strongly indicates that TRAPPIST-1 e is a particularly interesting planet, the atmospheric composition of which may shed light on key processes of formation and early rocky planet evolution. Furthermore, it will be worthwhile to compare TRAPPIST-1 e with TRAPPIST-1 g and f, which are expected to be more volatile-rich than e. Here, the CO2 atmospheric content will be key, to investigate if it represents the primordial volatile content or if it is modified after the magma ocean phase; for example, by a carbonate-silicate cycle.

In summary, MagmOc2.0 presents a valuable tool to gain insights into the early stages of rocky planet evolution and to understand the impact of various processes on the outgassed secondary atmosphere. These insights may facilitate rocky planet atmosphere characterization with the James Webb Space Telescope and future missions like the Habitable World Observatory and LIFE, and improve our understanding of the formation and evolution history of rocky planets (Bonati et al. 2019; Turbet et al. 2019; Way et al. 2022; Krissansen-Totton & Fortney 2022; Schlecker et al. 2024).

Data availability

Appendix D is available at https://doi.org/10.5281/zenodo.14442985.

Acknowledgements

We thank the anonymous referee for a highly constructive discussion that improved this work substantially. We thank Tim Lichtenberg and his group for excellent discussions about the opacities and atmospheres on rocky planets. We also thank Paul Mollière and Thomas Henning for support during the master thesis of P.B. that was the starting point for the whole MagmOc development. L.C. acknowledges support by the DFG priority programme SP1833 “Building a habitable Earth” Grant CA 1795/3 and the Royal Astronomical Society University Fellowship URF R1 211718 hosted by the University of St Andrews. R.B. acknowledges support from NASA grants numbered 80NSSC20K0229, 80NSSC18K0829 and 80NSSC18K0261. P.B. acknowledges financial support from the Austrian Academy of Science. Ch.H. and L.C. acknowledge funding from the European Union H2020-MSCA-ITN-2019 under Grant Agreement no. 860470 (CHAMELEON). L.N. acknowledges funding from the European Union (ERC, DIVERSE, 101087755).

Appendix A Derivation of the coupled H2O-CO2 outgassing differential equations

We present here an improvement of the volatile model used by Barth et al. (2021) for MagmOcl.O, which originally adopted the magma ocean model including outgassing of Schaefer et al. (2016). To derive the coupled differential equations that drive multi-component outgassing, we use the mass balance in the magma ocean and atmosphere as a starting point as by Barth et al. (2021, Eq. 12) (see also Eq 1).

For clarity, we outline again the setup of the volatile reservoirs. Each volatile i (here H2O and CO2 ) is stored across the magma ocean and atmosphere system (MOA) as follows (Fig. 1): a) in the crystallized portion of the magma ocean (Micrystal )$\left( {M_i^{{\rm{crystal}}}} \right)$, b)in the liquid magma phase (Miliq)$\left( {M_i^{{\rm{liq}}}} \right)$, c) and in the atmosphere (Miatm )$\left( {M_i^{{\rm{atm}}}} \right)$. Additionally, we define, Fi, as the mass fraction of the volatile i solved in the melt. We also assume a constant, mantle-averaged partition coefficient ki between the melt and the crystal phase. The latter is assumed to sink to the bottom and form the solidified mantle. For the partition coefficient, we adopt kH2O${k_{{{\rm{H}}_2}{\rm{O}}}}$ from Schaefer et al. (2016) and feco2 from Lebrun et al. (2013) (Table 1).

Lastly, the mass of volatile i in the atmosphere is calculated via Miatm=4πrp2ɡpi,mass,$M_i^{{\rm{atm}}} = {{4\pi r_{\rm{p}}^2} \over }{p_{i,{\rm{}}mass{\rm{}}}},$(A.1)

where pi,mass is mass weighted pressure with respect to the partial pressure pi,part of volatile i (see e.g., Bower et al. (2019)): pi,mass=μiμ¯atmpi,part,${p_{i,{\rm{mass}}}} = {{{\mu _i}} \over {{{\bar \mu }_{{\rm{atm}}}}}}{p_{i,{\rm{part}}}},$(A.2)

where µi represents the molecular mass of the specific volatile and μ¯atm${{\bar \mu }_{atm}}$ stands for the mean atmospheric molecular mass, considering the combination of all outgassed volatiles. It is worth noting that in the water-dominated steam atmosphere setup of MagmOc1.0 (Barth et al. 2021), the authors assumed pH2O,mass=pH2O${p_{{{\rm{H}}_2}{\rm{O}},mass}} = {p_{{{\rm{H}}_2}{\rm{O}}}}$ because it held true for most of the simulation time that μH2Oμ¯atm${\mu _{{{\rm{H}}_2}{\rm{O}}}} \approx {{\bar \mu }_{atm}}$. This assumption no longer holds in cases where a H2O dominated atmosphere evolves into CO2 dominated atmosphere as simulated in this work. We also neglect here O2 because atmospheric O2 is in this model typically produced after the end of the magma ocean and therefore does not enter the mass balance equations describing balance between outgassing and dissolved volatile mass in the magma ocean phase.

Thus, the volatile mass balance can be summarized as (see Table 2 for an overview of all the components of the volatile model.): Mimoa=Micrystal+Miliq+Miatm=kiFiMcrystal+FiMliq+4πrp2ɡpi, mass, $\matrix{ {M_i^{{\rm{moa}}} = M_i^{{\rm{crystal}}} + M_i^{{\rm{liq}}} + M_i^{{\rm{atm}}}} \cr { = {k_i}{F_i}{M^{{\rm{crystal}}}} + {F_i}{M^{{\rm{liq}}}} + {{4\pi r_{\rm{p}}^2} \over }{p_{i,{\rm{mass,}}}}} \cr } $(A.3)

where Mcrystal denotes the mass of the crystallized part of the magma ocean and Mliq is the mass of the liquid phase of the magma ocean. We further assume that the partial pressure of the volatile at the surface is determined by the fraction of the volatile in the melt Fi (Elkins-Tanton 2008; Schaefer et al. 2016; Lichtenberg et al. 2021).

In Barth et al. (2021), the water outgassing law of Schaefer et al. (2016), based on the laboratory data of Papale (1997), was used. Therefore, we employ for the pure water outgassing the same relation to compare with MagmOc1.0: pH2O=(FH2O3.44×108)1/0.74[Pa].${p_{{{\rm{H}}_2}{\rm{O}}}} = {\left( {{{{F_{{{\rm{H}}_2}{\rm{O}}}}} \over {3.44 \times {{10}^{ - 8}}}}} \right)^{1/0.74}}[{\rm{Pa}}].$(A.4)

For a mixed atmosphere model, Elkins-Tanton (2008) assumed the following solubility laws for H2O and CO2 that were also based Papale (1997) but have a form that suppresses outgassing below a certain volatile melt fraction: pH2O=(FH2O3×1032.08×106)1/0.52[Pa]${{p_{{{\rm{H}}_2}{\rm{O}}}} = {{\left( {{{{F_{{{\rm{H}}_2}{\rm{O}}}} - 3 \times {{10}^{ - 3}}} \over {2.08 \times {{10}^{ - 6}}}}} \right)}^{1/0.52}}[{\rm{Pa}}]}$(A.5) pCO2=(FCO25×1042.08×106)1/0.45[Pa].${{p_{{\rm{C}}{{\rm{O}}_{\rm{2}}}}} = {{\left( {{{{F_{{\rm{C}}{{\rm{O}}_2}}} - 5 \times {{10}^{ - 4}}} \over {2.08 \times {{10}^{ - 6}}}}} \right)}^{1/0.45}}[{\rm{Pa}}].}$(A.6)

Figure A.1 provides an overview of solubility laws used in the literature. The H2O solubility laws agree with each other within one order of magnitude for 260 bar≤ p < lO4 bar. Consequently, it is not surprising that magma oceans with H2O outgassing yield generally similar results (see Appendix C). However, significant deviations are apparent between the CO2 solubility law used by Elkins-Tanton (2008) and those used by Lichtenberg et al. (2021) and Nikolaou et al. (2019). Conversely, the CO2laws used by Lichtenberg et al. (2021) and Nikolaou et al. (2019) generally agree with each other.

In MagmOc1,0 the mass balance equations (Eq. A.3) were solved for FH2O${F_{{{\rm{H}}_2}{\rm{O}}}}$ at every time step using a root finding method. Using this approach for two or more volatiles, however, is cumbersome and numerically costly.

In the new model MagmOc2.0, we opt instead to solve by numerical integration: FH2O(t)=FH2O(0)+dFH2O(t)dtΔt${F_{{{\rm{H}}_2}{\rm{O}}}}(t) = {F_{{{\rm{H}}_2}{\rm{O}}}}(0) + {{d{F_{{{\rm{H}}_2}{\rm{O}}}}(t)} \over {dt}}{\rm{\Delta }}t$(A.7) FCO2(t)=FCO2(0)+dFCO2(t)dtΔt${F_{{\rm{C}}{{\rm{O}}_2}}}(t) = {F_{{\rm{C}}{{\rm{O}}_2}}}(0) + {{d{F_{{\rm{C}}{{\rm{O}}_2}}}(t)} \over {dt}}\Delta t$(A.8)

where the starting points of the integration, FH2O(0)${{F}_{{{\text{H}}_{2}}\text{O}}}(0)$ and FCO2(0)${{F}_{\text{C}{{\text{O}}_{2}}}}(0)$, are input parameters for MagmOc2.0 and are derived via root finding only once when preparing the simulations.

The derivatives of the melt fractions Fi${{F}_{i}}^{\prime }$ can be found by differentiating Eqs. A.3, assuming that the total volatile mass in the system, Mimoa$M_{i}^{moa\text{ }\!\!~\!\!\text{ }}$, has to remain constant when a new equilibrium in mass for H2O and CO2 is established between all reservoirs.

For mixed H2O-CO2 outgassing, we also account for changes in the mean molecular mass of the atmosphere μ¯atm${{\bar{\mu }}_{atm}}$, given by μ¯atm=pH2OμH2O+pCO2μCO2pH2O+pCO2psurf,${\bar \mu _{atm{\rm{}}}} = {{{p_{{{\rm{H}}_2}{\rm{O}}}} \cdot {\mu _{{{\rm{H}}_2}{\rm{O}}}} + {p_{{\rm{C}}{{\rm{O}}_2}}} \cdot {\mu _{{\rm{C}}{{\rm{O}}_2}}}} \over {\mathop {\mathop {{p_{{{\rm{H}}_2}{\rm{O}}}} + {p_{{\rm{C}}{{\rm{O}}_2}}}}\limits_ }\limits_{{p_{surf{\rm{}}}}} }},$(A.9)

where pH2O+pCO2${{p}_{{{\text{H}}_{2}}\text{O}}}+{{p}_{\text{C}{{\text{O}}_{2}}}}$ is the surface pressure of the complete atmosphere psurf.

To make the derivation more manageable, we introduce placeholders for several terms of the mass balance equation, which are listed in Table A.1. The derivatives of the mass balance equations for H2O and CO2 are then expressed by ddt(a1+b1ɡ+hɡ+dmw1(l+uv)exp2(ɡ+ef)exp1+1)=0${d \over {dt}}\left( {{a_1} + {b_1} \cdot + h \cdot + {{d \cdot m} \over {{w_1}{{\left( {{{l + u} \over v}} \right)}^{ex{p_2}}} \cdot {{\left( {{{ + e} \over f}} \right)}^{ - ex{p_1}}} + 1}}} \right) = 0$(A.10) ddt(a2+b2l+hl+dmw2(ɡ+ef)exp1(l+uv)exp2+1)=0${d \over {dt}}\left( {{a_2} + {b_2} \cdot l + h \cdot l + {{d \cdot m} \over {{w_2}{{\left( {{{ + e} \over f}} \right)}^{ex{p_1}}} \cdot {{\left( {{{l + u} \over v}} \right)}^{ - ex{p_2}}} + 1}}} \right) = 0$(A.11)

We point out that m′ is a place holder for the derivative of the total surface pressure psurf of the H2O- CO2 atmosphere and thus depends on both, FH2O=ɡ$F_{{{\text{H}}_{2}}\text{O}}^{\prime }={}'$ and FCO2=l$F_{\text{C}{{\text{O}}_{2}}}^{\prime }={l}'$. Its derivative is (according to the substitution we chose in Table A.1): dmdt=ddt(ɡ+ef)exp1+ddt(l+uv)exp2             =exp1(ɡf)(ɡ+ef)exp11+exp2(lv)(l+uv)exp21$\matrix{ {{{dm} \over {dt}} = {d \over {dt}}{{\left( {{{ + e} \over f}} \right)}^{ex{p_1}}} + {d \over {dt}}{{\left( {{{l + u} \over v}} \right)}^{ex{p_2}}}} \hfill \cr {\,\,\,\,\,\,\,\,\,\,\,\,\, = ex{p_1}\left( {{{'} \over f}} \right){{\left( {{{ + e} \over f}} \right)}^{ex{p_1} - 1}} + ex{p_2}\left( {{{l'} \over v}} \right){{\left( {{{l + u} \over v}} \right)}^{ex{p_2} - 1}}} \hfill \cr } $(A.12)

We now replace m′ in the derivation of the mass balance equations with this resolved expression and rearrange the two mass balance equations to gather all terms that contain FH2O=ɡ$F_{{{\text{H}}_{2}}\text{O}}^{\prime }={}'$ and FCO2=l$F_{\text{C}{{\text{O}}_{2}}}^{\prime }={l}'$ on the right hand side and all other terms on the left hand side. We thus derive for H2O: ɡ [ b1+h+dmexp1w11f(l+uv)exp2(ɡ+ef)exp1+1(w1(l+uv)exp2(ɡ+ef)exp1+1)2                           +dexp11f(ɡ+ef)exp11w1(l+uv)exp2(ɡ+ef)exp1+1 ]                     +l [ dmexp2w11v(l+uv)exp21(ɡ+ef)exp1(w1(l+uv)exp2(ɡ+ef)exp1+1)2                   +dexp21v(l+uv)exp21w1(l+uv)exp2(ɡ+ef)exp1+1 ]                   =[ a1+ɡb1+ɡh ].$\eqalign{ & {g^\prime }\left[ {{b_1} + h + d \cdot m \cdot {{{{\exp }_1}{w_1}{1 \over f} \cdot {{\left( {{{l + u} \over v}} \right)}^{{{\exp }_2}}}} \over {{{\left( {{{g + e} \over f}} \right)}^{{{\exp }_1} + 1}}}} \cdot {{\left( {{{{w_1}{{\left( {{{l + u} \over v}} \right)}^{{{\exp }_2}}}} \over {{{\left( {{{g + e} \over f}} \right)}^{{{\exp }_1}}}}} + 1} \right)}^{ - 2}}} \right. \cr & \left. {\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, + {{d \cdot {{\exp }_1}{1 \over f} \cdot {{\left( {{{g + e} \over f}} \right)}^{{{\exp }_1} - 1}}} \over {{w_1}{{\left( {{{l + u} \over v}} \right)}^{{{\exp }_2}}}{{\left( {{{g + e} \over f}} \right)}^{ - {{\exp }_1}}} + 1}}} \right] \cr & \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, + {l^\prime }\left[ { - d \cdot m \cdot {{{{\exp }_2}{w_1}{1 \over v} \cdot {{\left( {{{l + u} \over v}} \right)}^{{{\exp }_2} - 1}}} \over {{{\left( {{{g + e} \over f}} \right)}^{{{\exp }_1}}}}} \cdot {{\left( {{{{w_1}{{\left( {{{l + u} \over v}} \right)}^{{{\exp }_2}}}} \over {{{\left( {{{g + e} \over f}} \right)}^{{{\exp }_1}}}}} + 1} \right)}^{ - 2}}} \right. \cr & \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\left. { + {{d \cdot {{\exp }_2}{1 \over v}{{\left( {{{l + u} \over v}} \right)}^{{{\exp }_2} - 1}}} \over {{w_1}{{\left( {{{l + u} \over v}} \right)}^{{{\exp }_2}}}{{\left( {{{g + e} \over f}} \right)}^{ - {{\exp }_1}}} + 1}}} \right] \cr & \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\, = - \left[ {a_1^\prime + g \cdot b_1^\prime + g \cdot {h^\prime }} \right]. \cr} $(A.13)

and accordingly for CO2: l [ b2+h+dmexp2w2(1v)(ɡ+ef)exp1(l+uv)exp2+1(w2(ɡ+ef)exp1(l+uv)exp2+1)2                       +dexp2(1v)(l+uv)exp21w2(ɡ+ef)exp1(l+uv)exp2+1 ]          +ɡ [ dmexp1w21f(ɡ+ef)exp11(l+uv)exp2(w2(ɡ+ef)exp1(l+uv)exp2+1)2            +dexp11f(ɡ+ef)exp11w2(ɡ+ef)exp1(l+uv)exp2+1 ]           =[ a2+lb2+lh ].$\eqalign{ & {l^\prime }\left[ {{b_2} + h + d \cdot m \cdot {{{{\exp }_2}{w_2}\left( {{1 \over v}} \right){{\left( {{{g + e} \over f}} \right)}^{{{\exp }_1}}}} \over {{{\left( {{{l + u} \over v}} \right)}^{{{\exp }_2} + 1}}}} \cdot {{\left( {{{{w_2}{{\left( {{{g + e} \over f}} \right)}^{{{\exp }_1}}}} \over {{{\left( {{{l + u} \over v}} \right)}^{{{\exp }_2}}}}} + 1} \right)}^{ - 2}}} \right. \cr & \,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\,\left. { + {{d \cdot {{\exp }_2}\left( {{1 \over v}} \right){{\left( {{{l + u} \over v}} \right)}^{{{\exp }_2} - 1}}} \over {{w_2}{{\left( {{{g + e} \over f}} \right)}^{{{\exp }_1}}}{{\left( {{{l + u} \over v}} \right)}^{ - {{\exp }_2}}} + 1}}} \right] \cr & \,\,\,\,\,\,\,\,\,\, + {g^\prime }\left[ { - d \cdot m \cdot {{{{\exp }_1}{w_2}{1 \over f}{{\left( {{{g + e} \over f}} \right)}^{{{\exp }_1} - 1}}} \over {{{\left( {{{l + u} \over v}} \right)}^{{{\exp }_2}}}}} \cdot {{\left( {{{{w_2}{{\left( {{{g + e} \over f}} \right)}^{{{\exp }_1}}}} \over {{{\left( {{{l + u} \over v}} \right)}^{{{\exp }_2}}}}} + 1} \right)}^{ - 2}}} \right. \cr & \,\,\,\,\,\,\,\,\,\,\,\left. { + {{d \cdot {{\exp }_1}{1 \over f}{{\left( {{{g + e} \over f}} \right)}^{{{\exp }_1} - 1}}} \over {{w_2}{{\left( {{{g + e} \over f}} \right)}^{{{\exp }_1}}}{{\left( {{{l + u} \over v}} \right)}^{ - {{\exp }_2}}} + 1}}} \right] \cr & \,\,\,\,\,\,\,\,\,\,\, = - \left[ {a_2^\prime + l \cdot b_2^\prime + l \cdot {h^\prime }} \right]. \cr} $(A.14)

These two equations establish a set of linear equations with respect to ɡ=FH2O${}'=F_{{{\text{H}}_{2}}\text{O}}^{\prime }$ and l=FCO2$\[{l}'=F_{\text{C}{{\text{O}}_{2}}}^{\prime }\]$ of the form: AH2Ol+BH2Oɡ=CH2O${{A_{{{\rm{H}}_2}{\rm{O}}}}l' + {B_{{{\rm{H}}_2}{\rm{O}}}}g' = - {C_{{{\rm{H}}_2}{\rm{O}}}}}$(A.15) ACO2l+BCO2ɡ=CCO2.${{A_{{\rm{C}}{{\rm{O}}_2}}}l' + {B_{{\rm{C}}{{\rm{O}}_2}}}g' = - {C_{{\rm{C}}{{\rm{O}}_2}}}.}$(A.16)

Therefore, we can use Cramer’s rule to derive unique solutions for 𝑔′ and l′ as long as AH2OBCO2ACO2BH2O${{A}_{{{\text{H}}_{2}}\text{O}}}{{B}_{\text{C}{{\text{O}}_{2}}}}\ne {{A}_{\text{C}{{\text{O}}_{2}}}}{{B}_{{{\text{H}}_{2}}\text{O}}}$: ɡ=dFH2Odt=CH2OACO2CCO2AH2OAH2OBCO2ACO2BH2O${g' = {{d{F_{{{\rm{H}}_2}{\rm{O}}}}} \over {dt}} = {{{C_{{{\rm{H}}_2}{\rm{O}}}}{A_{{\rm{C}}{{\rm{O}}_2}}} - {C_{{\rm{C}}{{\rm{O}}_2}}}{A_{{{\rm{H}}_2}{\rm{O}}}}} \over {{A_{{{\rm{H}}_2}{\rm{O}}}}{B_{{\rm{C}}{{\rm{O}}_2}}} - {A_{{\rm{C}}{{\rm{O}}_2}}}{B_{{{\rm{H}}_2}{\rm{O}}}}}}}$(A.17) l=dFCO2dt=CCO2BH2OCH2OBCO2AH2OBCO2ACO2BH2O.${l' = {{d{F_{{\rm{C}}{{\rm{O}}_2}}}} \over {dt}} = {{{C_{{\rm{C}}{{\rm{O}}_2}}}{B_{{{\rm{H}}_2}{\rm{O}}}} - {C_{{{\rm{H}}_2}{\rm{O}}}}{B_{{\rm{C}}{{\rm{O}}_2}}}} \over {{A_{{{\rm{H}}_2}{\rm{O}}}}{B_{{\rm{C}}{{\rm{O}}_2}}} - {A_{{\rm{C}}{{\rm{O}}_2}}}{B_{{{\rm{H}}_2}{\rm{O}}}}}}.}$(A.18)

Thus, we demonstrate that while the mass balance equation with the parametric equation for volatile outgassing cannot be solved for analytically with respect to FH2O${{F}_{{{\text{H}}_{2}}\text{O}}}$ and FCO2${{F}_{\text{C}{{\text{O}}_{2}}}}$, their derivatives can. Therefore in MagmOc2.0, we add the two differential equations, Eqs. (A.7) and (A.8), with Eqs. A.17 and A.18. We note that the last term of the sum comprising AH2O,,BH2O${{A}_{{{\text{H}}_{2}}\text{O}}},\ldots ,{{B}_{{{\text{H}}_{2}}\text{O}}}$ encapsulates changes in surface pressure and also in mean molecular weight (see also Eq. A.2).

We further emphasize, similar to Bower et al. (2019), that the derivation outlined in this section can in principle be generalized to more than two volatiles as long as it is possible to derive a set of n independent linear equations for n volatiles that can be solved for time derivatives of all relevant volatile mass fractions F1… n.

Table A.1

Placeholder for differentiation of mass balance equations.

Appendix B A corrected gray atmosphere model for a two component H2O-CO2 atmosphere

As is outlined in Sect. 2.5, the thermal evolution of the planet is determined by the net outgoing flux Fnet; that is, the difference between ASR and the thermal flux emitted by the planet, also known as OLR, on top of the atmosphere. In MagmOc1.0 (Barth et al. 2021), the OLR was calculated with a gray atmosphere model: FOLR=ϵH2OσTsurf 4,${F_{OLR}} = {_{{{\rm{H}}_2}{\rm{O}}}}\sigma T_{{\rm{surf}}}^4,$(B.1)

where σ is the Stefan-Boltzmann constant, and ϵH2O${{\epsilon }_{{{\text{H}}_{2}}\text{O}}}$ is the emissivity of the H2O atmosphere.

The emissivity of a volatile i can be derived from the optical depth of an atmosphere composed of volatile i with the Rosse- land mean opacity over the infrared wavelength range, κ0(i), at the reference pressure, p0. The optical depth, τi, further depends on surface gravity, 𝑔, and partial pressure, pi, at surface as (see e.g., Carone et al. 2014; Elkins-Tanton 2008; Catling & Kasting 2017; Pierrehumbert 2010): τi=pi0.75κ0(i)/(ɡp0).${\tau _i} = {p_i}\sqrt {0.75 \cdot {\kappa _0}(i)/\left( { \cdot {p_0}} \right)} .$(B.2)

From this relation, the emissivity, єi, follows as ϵi=2τi+2.${_i} = {2 \over {{\tau _i} + 2}}.$(B.3)

We used this gray formalism as a basis to compare to calculations with full RT in a vertically extended atmosphere and to derive a parametric fit.

B.1 Pure H2O atmosphere for Earth gravity

For a pure H2O atmosphere, the gray atmosphere analytical model with κ0(H2O) = 0.25 m2/kg at p0 = 1.013 bar and the runaway greenhouse radiation limit of OLRlim(H2O) = 282 W/m2 by prescribing (Fig. B.1, left) FOLR(H2O,cP=const)=max(OLRlim(H2O),ϵH2OσTsurf 4)${F_{OLR}}\left( {{{\rm{H}}_2}{\rm{O}},{c_P} = {\rm{const}}} \right) = \max \left( {{\rm{OL}}{{\rm{R}}_{lim{\rm{}}}}\left( {{{\rm{H}}_2}{\rm{O}}} \right),{_{{{\rm{H}}_2}{\rm{O}}}}\sigma T_{{\rm{surf}}}^4} \right)$(B.4)

yields an agreement for Tsurf = 500 − 4000 K and psurf = 1 − 26000 bar within one order of magnitude when compared to thermal emission from full RT (Sect. 2.5) in a vertically extended atmosphere (Sect. 2.4) with a constant dry adiabatic lapse rate of cp,H2O=37 J${{c}_{p,{{\text{H}}_{2}}\text{O}}}=37 J$ J mole−1 K−1.8 The heat capacity of water is, however, not constant, but increases strongly for T > 1000 K, which results in a steeper gradient for temperature profiles for the early stages of the magma ocean evolution (Fig. B.2). The steeper temperature-pressure gradient results in hotter upper atmospheric layers and consequently higher outgoing thermal fluxes (Fig. B.1, left). We can account to first order for increased thermal emission with larger surface temperatures and pressures (Fig. B.1, right) by adopting an emission correction term: E*(H2O)=(Tsurf1500K)loɡ10(pH2O1bar).${E^*}\left( {{{\rm{H}}_2}{\rm{O}}} \right) = {\left( {{{{T_{surf{\rm{}}}}} \over {1500K}}} \right)^{lo{_{10}}\left( {{{{p_{{{\rm{H}}_{\rm{2}}}{\rm{O}}}}} \over {1bar}}} \right)}}.$(B.5)

The corrected gray atmosphere model for a pure H2O is then described with FOLR(H2O)=max(OLRlim(H2O),ϵH2OσTsurf4E*(H2O)).${F_{OLR{\rm{}}}}\left( {{{\rm{H}}_2}{\rm{O}}} \right) = \max \left( {{\rm{OL}}{{\rm{R}}_{lim{\rm{}}}}\left( {{{\rm{H}}_2}{\rm{O}}} \right),{_{{{\rm{H}}_2}{\rm{O}}}}\sigma T_{surf{\rm{}}}^4{E^*}\left( {{{\rm{H}}_2}{\rm{O}}} \right)} \right).$(B.6)

B.2 Pure CO2 atmosphere for Earth gravity

For a pure CO2 atmosphere, we find even larger disagreement between thermal emission based on the gray atmosphere model and the full RT calculations (Fig. B.3, left). The gray atmosphere model (κ0(CO2) = 0.001 m2/kg at p0 = 1.013 bar) yields only good agreement for psurf = 1 bar and underestimates the OLR by orders of magnitudes for higher surface pressures. Assuming κ0(CO2) = 0.005 m2/kg at p0 = 1.013 bar, as used in Lebrun et al. (2013); Elkins-Tanton (2008), yields even worse results and underestimates the OLR by at least one order of magnitude also for psurf = 1 bar.

The following analytical prescription is, however, able to mimic the OLR from the full RT calculations in a vertically extended pure CO2 atmosphere to good accuracy for surface pressures between 1 − 26000 bar and Tsurf = 500 − 4000 K. Here, we set the CO2 radiation limit OLRlim(CO2) = 64 W/m2 (Fig. B.3, right): FOLR(CO2)=max(OLRlim(CO2),ϵCO2*σTsurf4E*),${F_{OLR}}\left( {{\rm{C}}{{\rm{O}}_2}} \right) = \max \left( {OL{R_{lim{\rm{}}}}\left( {{\rm{C}}{{\rm{O}}_2}} \right),_{{\rm{C}}{{\rm{O}}_2}}^*\sigma T_{surf{\rm{}}}^4 \cdot {E^*}} \right),$(B.7)

and used an adapted optical depth for CO2 and the partial pressure, PCO2${{P}_{\text{C}{{\text{O}}_{2}}}}$: τCO2*=(pCO2pref(CO2))0.75pref(CO2)0.75κ0(CO2)/(ɡp0),$\tau _{{\rm{C}}{{\rm{O}}_2}}^* = {\left( {{{{p_{{\rm{C}}{{\rm{O}}_2}}}} \over {{p_{ref{\rm{}}}}\left( {{\rm{C}}{{\rm{O}}_2}} \right)}}} \right)^{0.75}} \cdot {p_{ref{\rm{}}}}\left( {{\rm{C}}{{\rm{O}}_2}} \right)\sqrt {0.75 \cdot {\kappa _0}\left( {{\rm{C}}{{\rm{O}}_2}} \right)/\left( { \cdot {p_0}} \right)} ,$(B.8)

where pref (CO2)=1 bar is the reference pressure for the fit. We then derived ϵCO2$\epsilon _{\text{C}{{\text{O}}_{2}}}^{*}$via Eq. (B.2) by substituting τCO2${{\tau }_{\text{C}{{\text{O}}_{2}}}}$ with τCO2$\tau _{\text{C}{{\text{O}}_{2}}}^{*}$.

thumbnail Fig. B.1

Outgoing thermal flux for a pure H2O atmosphere on an Earth gravity planet (𝑔 = 9.81 m/s2) based on the analytic gray atmosphere and full RT calculations under different assumptions. The results of the full RT calculations in a vertically extended atmosphere (Full RT) with heat capacity varying with temperature, cP(T), are used for benchmarking (solid lines). The models are compared for the same surface temperatures and pressures, where different surface pressures are indicated by variations in color. Left panel: Results of full RT calculations with temperature-dependent cP (solid lines), full RT with constant cP (dashed lines), and the gray atmosphere model (dash-dotted lines). Right panel: Results of the corrected gray atmosphere (dash-dotted lines) compared to the full RT.

thumbnail Fig. B.2

Left: Specific heat capacity versus temperature for constant cP (blue) and cP(T) (dark orange). Right: H2O pressure-temperature profiles for psurf = 260 bar and Tsurf = 1000 K (dash-dotted) and 2000 K (solid). Profiles assuming constant cP are shown in blue and profiles assuming temperature dependent cP are shown in dark orange. Note that for a 1000 K surface, the difference between the profiles is negligible.

As in the pure H2O-atmosphere case, the steepness of the pressure temperature profile for large surface temperatures in a CO2 atmosphere requires a correction term for the emission, which we fit as E*(CO2)=(Tsurf1200K)loɡ10(pCO21bar).${E^*}\left( {{\rm{C}}{{\rm{O}}_2}} \right) = {\left( {{{{T_{surf{\rm{}}}}} \over {1200K}}} \right)^{lo{_{10}}\left( {{{{p_{{\rm{C}}{{\rm{O}}_{\rm{2}}}}}} \over {{\rm{1}}bar}}} \right)}}.$(B.9)

B.3 Mixed H2O-CO2 atmosphere for Earth gravity

For mixed H2O-CO2 atmospheres, the following parametric approximation yields thermal emission within one order of magnitude compared to the full RT model (Fig. B.4) for Tsurf = 500 K - 4000 K and psurf = 1 bar - 2600 bar using: FOLR(xH2O)=max(OLRlim(xH2O),ϵMix*σTsurf4EMix*),${F_{{\rm{OLR}}}}\left( {{x_{{{\rm{H}}_2}{\rm{O}}}}} \right) = \max \left( {{\rm{OL}}{{\rm{R}}_{\lim }}\left( {{{\rm{x}}_{{{\rm{H}}_2}{\rm{O}}}}} \right),_{{\rm{Mix}}}^*\sigma {\rm{T}}_{{\rm{surf}}}^4 \cdot {\rm{E}}_{{\rm{Mix}}}^*} \right),$(B.10)

with ϵMix*=2τMix+2,$_{Mix}^* = {2 \over {{\tau _{Mix}} + 2}},$(B.11)

where τMix is the sum of the modified optical depths derived in the previous two subsections: τMix=τH2O+τCO2*.${\tau _{Mix}} = {\tau _{{{\rm{H}}_2}{\rm{O}}}} + \tau _{{\rm{C}}{{\rm{O}}_2}}^*.$(B.12)

thumbnail Fig. B.3

Outgoing thermal flux for a pure CO2 atmosphere on an Earth gravity planet (𝑔 = 9.81 m/s2) are shown for gray atmosphere model with different assumptions. The results of the full RT in a vertically extended atmosphere (“Full RT”) is used for benchmarking (solid lines). The models are compared for the same surface temperatures and pressures, where different surface pressures are indicated by variations in color. Left panel: Results of the gray atmosphere model is shown for ĸ0(CO2) = 0.001 m2/kg (dashed lines) and ĸ0(CO2) = 0.05 m2/kg (dash- dotted lines). Right panel: Results of the corrected gray atmosphere is shown (dash-dotted lines) in comparison to the Full RT calculations.

The runaway greenhouse radiation radiation, or H2O OLR limit, decreases when CO2 is added to a H2O-dominated atmosphere, as already pointed out by Goldblatt et al. (2013). We systemti- cally investigate this effect for a vertically extended H2O-CO2 atmosphere. We find that the following fit to the H2O OLR limit OLRLim(xH2O)$OL{{R}_{Lim}}\left( {{x}_{{{\text{H}}_{2}}\text{O}}} \right)$ with respect to volume mixing ratio (xH2O)$\left( {{x}_{{{\text{H}}_{\text{2}}}\text{O}}} \right)$ and for Earth surface gravity (𝑔 = 9.81 m/s) yields agreement to first order (Fig. B.5): OLRLim(xH2O)=4.9log10(xH2O)2+66.9log10(xH2O)+282,${\rm{OL}}{{\rm{R}}_{{\rm{Lim}}}}\left( {{{\rm{x}}_{{{\rm{H}}_2}{\rm{O}}}}} \right) = 4.9 \cdot {\log _{10}}{\left( {{{\rm{x}}_{{{\rm{H}}_2}{\rm{O}}}}} \right)^2} + 66.9 \cdot {\log _{10}}\left( {{{\rm{x}}_{{{\rm{H}}_2}{\rm{O}}}}} \right) + 282,$(B.13)

where xH2O106${{x}_{{{\text{H}}_{2}}\text{O}}}\le {{10}^{-6}}$ was set to be identical to a pure CO2 atmosphere. That is, we assume for such low volume mixing ratios of water in a CO2-dominated atmosphere that the water contribution can be neglected (within 10% accuracy) when simulating the thermal evolution of the magma ocean (see also Fig. 2, which shows that both the pressure-temperature profile and the emission are very close to that of a 100% CO2 atmosphere for xH2O=105${{x}_{{{\text{H}}_{2}}\text{O}}}={{10}^{-5}}$).

We adopt the following correction in emission, using the volume fraction of H2O, xH2O=pH2O/psurf${{x}_{{{\text{H}}_{2}}\text{O}}}={{p}_{{{\text{H}}_{2}}\text{O}}}/{{p}_{surf}}$, in a two-component atmosphere setup with pH2O+pCO2=psurf${{p}_{{{\text{H}}_{2}}\text{O}}}+{{p}_{\text{C}{{\text{O}}_{2}}}}={{p}_{surf}}$: EMix*=(Tsurf1200K)loɡ10((psurf1bar)(1xH2O)+xH2O)(1xH2O)+              (Tsurf1500K)loɡ10( (psurf1bar)xH2O+(1xH2O))xH2O$\eqalign{ & E_{Mix}^* = {\left( {{{{T_{surf{\rm{ }}}}} \over {1200K}}} \right)^{lo{_{10}}\left( {\left( {{{{p_{surf{\rm{ }}}}} \over {1{\rm{ }}bar{\rm{ }}}}} \right) \cdot \left( {1 - {x_{{{\rm{H}}_2}{\rm{O}}}}} \right) + {x_{{{\rm{H}}_2}{\rm{O}}}}} \right)}} \cdot \left( {1 - {x_{{{\rm{H}}_2}{\rm{O}}}}} \right) + \cr & \,\,\,\,\,\,\,\,\,\,\,\,\,\,{\left( {{{{T_{surf{\rm{ }}}}} \over {1500K}}} \right)^{lo{_{10}}\left( {\left( {{{{p_{surf{\rm{ }}}}} \over {1bar}}} \right){x_{{{\rm{H}}_2}{\rm{O}}}} + \left( {1 - {x_{{{\rm{H}}_2}{\rm{O}}}})} \right)} \right.}} \cdot {x_{{{\rm{H}}_2}{\rm{O}}}} \cr} $(B.14)

Two examples of the resulting grid emission and corrected gray emission are shown in (Fig. B.4), using κ0(CO2) = 0.0008 m2/kg and κ0(H2O) = 0.25 m2/kg at p0 = 1.013 bar. These show that the corrected gray atmosphere model yields agreement to first order to the full RT calculations even for very different atmosphere mixing. We note that this two-component H2O-CO2 atmosphere model also encapsulates 100% H2O and CO2 compositions by setting xH2O${{x}_{{{\text{H}}_{\text{2}}}\text{O}}}$ equal to 1 or 0, respectively.

thumbnail Fig. B.4

Outgoing thermal flux for a mixed H2O-CO2 atmosphere on an Earth gravity planet (𝑔 = 9.81 m/s2) based on the corrected gray atmosphere (dashed lines) and full RT calculations with vertically extended atmosphere (solid lines) for the same surface temperatures and pressures, respectively. Different surface pressures are indicated by variations in color. Results are shown for two examples of a two component atmosphere with water volume mixing ratio xH2O=0.75${{x}_{{{\text{H}}_{\text{2}}}\text{O}}}=0.75$ (left panel) and xH2O=1014${{x}_{{{\text{H}}_{\text{2}}}\text{O}}}={{10}^{-14}}$(right panel), respectively.

thumbnail Fig. B.5

OLR limit for Earth with a two component H2O-CO2 atmosphere for different H2O volume mixing ratios using full RT (circles) vs. the parametric fit defined in Eq. (B.13) (solid line).

B.4 Generalization of gray atmosphere for rocky exoplanets

So far, we have applied the RT and corrected gray atmosphere model to Earth. In principle, the latter can also be applied to any rocky planet with a different surface gravity 𝑔.

The pressure-temperature profiles will not change with different surface gravities for a given atmospheric composition and surface temperature and pressure set by the hot magma ocean, because the atmospheric lapse rate is determined either by R/cp,i when the vertical coordinate is chosen, where R is the ideal gas constant and cp (T) is the combined specific heat capacity of the volatiles in the atmosphere or by latent heat release Li. Thus, pressure-temperature profiles for rocky planets with the same atmospheric composition, surface temperature, and surface pressure are in this setup identical for different surface gravities. The emission arising from the top of the atmosphere, however, will still change because the planet’s opacity depends on surface gravity (Eq. B.2). Consequently, also the runaway greenhouse limit changes with surface gravity (Fig. B.6, left).

thumbnail Fig. B.6

Left panel: Emission versus wavelength for a pure H2O atmosphere in the runaway greenhouse limit for different surface gravities. Right panel: OLR limit for a two component H2O-CO2 atmosphere and different surface gravities (circles) including the adjusted parametric fits (solid lines) using Eq. (B.15).

We find that the runaway greenhouse radiation limit for different surface gravities can be parameterized with a simple adjustment to Eq. (B.13) (Fig. B.6, right): OLRLim(xH2O, g)=OLRLim(xH2O)(1+ln(ggEarth )0.1).${\rm{OL}}{{\rm{R}}_{{\rm{Lim}}}}\left( {{x_{{{\rm{H}}_2}{\rm{O}}}},{\rm{g}}} \right) = {\rm{OL}}{{\rm{R}}_{{\rm{Lim}}}}\left( {{{\rm{x}}_{{{\rm{H}}_2}{\rm{O}}}}} \right) \cdot \left( {1 + \ln \left( {{{\rm{g}} \over {{{\rm{g}}_{{\rm{Earth}}}}}}} \right) \cdot 0.1} \right).$(B.15)

Appendix C Validation of MagmOc2.0 for an oxidized Earth

We applied our magma ocean to the Earth regime to make a comparison with the previous results of Barth et al. (2021), who also benchmarked MagmOc1.0 for Earth-like conditions for comparison with Hamano et al. (2013) and Elkins-Tanton (2008). We further exploited the versatility of MagmOc2.0 to test three different atmosphere models and their impact on the thermal evolution of the mag, a ocean: the gray formalism as used in MagmOc1.0, an improved gray formalism, and a model based on full RT calculations in a vertically extended atmosphere. Henceforth, we refer to the first model as the “gray” model, to the second as the “corrected gray” model, and to the third as the “RT” model. Moreover, we investigate the impact of various H2O and CO2 outgassing laws. Specifically, we compare the laws used by Elkins-Tanton (2008) and Nikolaou et al. (2019), respectively. Finally, we assess the impact of adding CO2 to a water-dominated atmosphere for our simulations of the oxidized Earth magma ocean.

In the following, we begin by evaluation the role of the different atmosphere models for pure H2O and CO2 atmospheres, respectively. Subsequently, we compare the outcome of simulations using the different atmosphere models for mixed H2O-CO2 atmospheres. For these mixed scenarios, we adopt as in Barth et al. (2021), an initial magma ocean depth of 2000 km, consistent with Elkins-Tanton (2008), one of the first models to investigate magma oceans with outgassing of both, H2O and CO2. Next, we assess the feedback between H2O and CO2 outgassing and compare to the results of Bower et al. (2019).

C.1 Pure H2O atmosphere

For the pure H2O atmosphere, we follow Barth et al. (2021) by simulating the evolution of the Earth magma ocean with an initial water content of 5 TO for comparison with Hamano et al. (2013). We use the H2O solubility law of Schaefer et al. (2016) and begin with fully molten mantle down to the core (RCore = 3400 km), which gives an initial magma ocean depth of 2978 km. We test MagmOc2.0 with the gray atmosphere model and ĸ0(H2O) = 0.01 m2/kg, the corrected gray atmosphere model (Appendix B and ĸ0(H2O) = 0.25 m2/kg), and the RT atmosphere model (Sect. 2.5).

Figure C.1 (top) illustrates that the gray atmosphere model yields solidification after 0.9 Myrs, as was expected9 from Barth et al. (2021). The solidification time of 0.9 Myrs is notably shorter than the Earth magma ocean duration of 4 Myrs reported by Hamano et al. (2013) for the same initial water mass. The simulations with the corrected gray and full RT atmosphere model yield longer solidification times of 1.5 Myrs, aligning more closely to Hamano et al. (2013). The magma ocean simulations reach across all three atmosphere models the runaway greenhouse radiation limit, OLRlim(1) = 282 W/m2, toward the end of the magma ocean evolution. As the simulations progress further, the net flux increases because the Sun’s bolometric flux, and thus the ASR, has substantially decreased after 1 Myrs, whereas the planet’s OLR does not change significantly. We note that the gray atmosphere model consistently assumes OLRlim(1) = 282 W/m2 for Tsurf ≤ 1800 K, irrespective of surface pressure and prior OLR evolution. This assumption can result in an abrupt drop in net flux, as is evident in Fig. C.1 (bottom panel).

We further compare our results to the Earth magma ocean simulations of Nikolaou et al. (2019), who also investigated differences between a gray atmosphere model and a full RT model for the pure H2O atmosphere. However, these authors adopted an initial water content of about 1.0 TO H2O. For better comparison with this work and that of Elkins-Tanton (2008), we adopt an initial H2O melt fraction of 0.05 in a 2000 km deep magma ocean, which is equivalent to 1.02 TO initial H2O mass and aligns with the magma ocean depth of the model by Elkins-Tanton (2008). These simulations yield shorter solidification times ranging between 0.8 and 1 Myrs compared to the 5 TO H2O simulation, where once again the gray model results again in the shortest solidification time (Fig. C.1, bottom).

In both the 5 TO and 1.02 TO H2O cases, we consistently find that the OLR of the gray model is higher during the earliest evolution stages (within the range of 103 – 104 years) compared to the corrected gray and RT atmosphere model. The higher ini- tal OLR of the gray atmosphere model leads to faster overall cooling, thereby explaining the shorter solidification times. A similar deviation in OLR evolution between a gray and RT atmosphere model was reported by Nikolaou et al. (2019, Fig. 3). Consequently, they found that in their simulation with a gray atmosphere model that the magma ocean solidified after several 0.1 Myrs, whereas the simulation using the line-by-line RT model of Katyal et al. (2019) yielded a solidification time of 1 Myrs, in good agreement with our 1.02 TO H2O simulations. These results confirm that the magma ocean lifetime for a pure H2O atmosphere can differ by several 0.1 million years across different models and atmosphere assumptions (Nikolaou et al. 2019).

However, our simulations show here that the corrected gray and full RT yield evolution tracks that are in good agreement with each other in our magma ocean model. We further find agreement between the OLR of our RT model and the RT calculations of Lichtenberg et al. (2021) for a pure H2O atmosphere with Tsurf = 500 - 3000 K and psurf = 1 bar and 260 bar. These authors similarly reproduce a magma ocean lifetime of about 1 Myrs with a pure H2O atmosphere, where we note different definitions for solidification times (see Sect. 3). Furthermore, Lichtenberg et al. (2021) have a cooler initial temperature of 3000 K compared to our initial temperature of 4000 K (Hamano et al. 2013; Schaefer et al. 2016; Nikolaou et al. 2019; Barth et al. 2021). The general agreement between our pure H2O simulations and that of Lichtenberg et al. (2021) is not surprising, given the usage of very similar H2O opacities (Table 3) and vertically extended atmospheres for the RT calculations.

C.2 Pure CO2 atmosphere

We simulate the pure CO2 outgassing scenario for Earth following Elkins-Tanton (2008). Consequently, we assume an initial magma ocean depth of 2000 km and a CO2 outgassing law with initial saturation in the melt (Sect. 2.2). The simulation assumes an initial CO2 melt fraction of 0.6, which corresponds to 14.7 TO CO2.

Figure C.2 demonstrates the thermal evolution of our Earth magma ocean simulations with a dense CO2 atmosphere. We find that solidification occurs after 0.5 Myrs when the gray model is used, in qualitative agreement with the solidification time of 0.8 Myrs reported by Elkins-Tanton (2008). Table C.1 also illustrates general agreement between the final volatile budget and atmospheric pressure in MagmOc2.0 compared to the results of Elkins-Tanton (2008). The agreement between both models in volatile reservoirs and solidification times with the gray atmosphere model, provides once again confidence that our model, despite its simplification, captures the magma ocean accurately enough for comparison with more complex magma ocean models.

thumbnail Fig. C.1

MagmOc2.0 simulations for Earth and an initial water content of 5 TO (top panel) and 1.02 TO (bottom panel). Depicted from left to right in each panel are: Surface temperature, H2O atmosphere surface pressure and net flux at the top of the atmosphere (FOLRFASR). Three different atmosphere models are used: The gray atmosphere model (solid red line), the corrected gray model (dotted red line), and the RT model (solid purple line).

thumbnail Fig. C.2

Simulation of MagmOcV2.0 for Earth and 14.7 TO CO2 content for a 2000 km deep magma ocean. Depicted from left to right: Surface temperature, H2O atmosphere surface pressure, and net flux at the top of the atmosphere (FOLRFASR) for three different atmosphere models. These models are the gray model without any correction (solid red line), the corrected gray model (dotted red line), and the RT adaptation (solid purple line).

However, significant differences in solidification time compared to the gray atmosphere model arise in simulations that use the corrected gray and RT atmosphere model. More precisely, the solidification time decreases by one order of magnitude to approximately 20 000 - 30 000 years. This discrepancy stems from the fact that the nominal gray model neglects the vertical extent of the atmosphere. The large extension of a hot, dense CO2 atmosphere is evident from Fig. B.2 and also supported by Lichtenberg et al. (2021, Fig. 3). A pure CO2 atmosphere has a much steeper temperature gradient in the troposphere compared to H2O, resulting in hotter atmosphere layers that contribute to the thermal emission on top of the atmosphere. Therefore, a particularly high thermal flux occurs on top of a CO2-dominated atmosphere for the initial magma ocean stage with high surface temperatures (Tsurf ≥ 2000 K) and pressures (psurf > 1 bar), which are not captured with a gray model (Appendix B).

thumbnail Fig. C.3

Pure CO2 Earth magma ocean simulation with 14.7 TO of initial CO2 mass following Elkins-Tanton (2008) and using the RT atmosphere model (solid black line). Two RT thermal emission grid lines for psurf = 655 bar (blue solid line) and psurf = 4122 bar (orange solid line) are shown, respectively, as specified in Sect. 2.5. The corresponding thermal emission for the gray atmosphere model as used in Elkins-Tanton (2008) is shown for comparison (dashed lines). The thermal emission deviates between the RT and gray atmosphere model by up to two orders of magnitude during the magma ocean stage in this case with Tsurf = 1500-3000 K.

Figure C.3 provides a detailed illustration of the strong deviation in emission between the gray model and the full RT grid. The RT grid model yields OLR > 105 W/m2 for Tsurf > 2500 K, whereas the gray model yields for the same temperature and pressure range OLR < 105 W/m2. Figure C.2 also demonstrates how the magma ocean simulation with 14.7 TO initial CO2 mass evolves with the RT grid model (black line) evolves within the emission grid, build with full RT calculations. The comparison reveals that the OLR differs by more than one order of magnitude already at the start of the simulation (Fig. C.2, panel right).

Similarly, Lichtenberg et al. (2021) demonstrate that a magma ocean with a pure CO2 atmosphere and full RT yields emits large amounts of flux (OLR > 105 W/m2) for Tsurf > 2500 K and psurf = 260 bar. When we further compare our CO2 magma ocean simulation to that of Lichtenberg et al. (2021), we likewise find that their magma ocean with a 200 - 300 bar dense CO2 atmosphere reaches surface temperatures between 1500-1600 K, which corresponds to our solidification criterion, after 10 000 years in the same order of magnitude than our CO2 magma ocean simulation.

While the gray atmosphere model fails to accurately capture the thermal evolution of a magma ocean with a dense CO2 atmosphere, the agreement between the RT grid simulation and the simulation with the corrected gray atmosphere model is much better (Fig. C.3). The latter yields a solidification time of the same order of magnitude (40 000 years). The results of the corrected gray model shows that our analytical approximation generally reproduce thermal emission even for a hot, vertically extended CO2 atmosphere (Appendix B).

C.3 Mixed CO2-H2O atmospheres - Different atmosphere models

To validate MagmOc2.0 for a mixed H2O-CO2 atmosphere, we follow again Barth et al. (2021) and compare our simulations first to results of the Earth simulations of Elkins-Tanton (2008).

In Barth et al. (2021), the pure H2O atmosphere already showed promising agreement with Elkins-Tanton (2008).

We test here two scenarios: one with melt fractions 0.05 and 0.01 for H2O and CO2, respectively, which we call henceforth the dry scenario, the other with melt fractions 0.5 and 0.1 for H2O and CO2, respectively, which we call henceforth the wet scenario. With an initial magma ocean depth of 2000 km, these melt fractions correspond for the dry case to an initial water mass 1.02 TO of H2O and for the wet case to an initial mass of 10.2 TO H2O. The initial CO2 mass is 0.21 and 2.12 TO, respectively. These melt fractions (Table C.1) also correspond to a 5:1 ratio between H2O and CO2 that is assumed by Elkins-Tanton (2008). We further adopt the same solubility laws used by Elkins-Tanton (2008) that differ from the H2O outgassing law used in the previous section, where instead the law by Schaefer et al. (2016) was used (see Sect. 2.2 for an overview of outgassing laws used in this work).

thumbnail Fig. C.4

Earth magma ocean simulation setup for comparison with Elkins-Tanton (2008) with the traditional gray model but without greenhouse limit and the full RT grid model. Scenarios are as listed in Table C.1. The solid lines denote results of simulations with the gray atmosphere model. The dotted lines denote results of simulations with full RT. We note again that we show the net flux, which is equal ASR- OLR.

First, we investigate the impact of the gray atmosphere and the full RT atmosphere model for mixed H2O-CO2 outgassing. We emphasize, however, that we “switched off” the runaway greenhouse radiation limit in the gray atmosphere model for these specific simulations. This choice is made because Elkins-Tanton (2008) did not include this limiting factor for the OLR. In all other gray simulations in this work, we consider the radiation limit. Figure C.4 (solid lines) and Table C.1 illustrate that our gray simulations without the runaway greenhouse radiation limit are in general agreement with the solidification times reported by Elkins-Tanton (2008): The magma ocean in the wet case solidifies after 1.7 Myrs, and in the dry case after only 0.1 Myrs.

In the wet case, the differences between simulations with the gray atmosphere model and the RT model do not significantly impact the magma ocean solidification times. However, the RT model simulation yields a hotter surface during the early evolution due to a lower initial net flux, resulting in less outgassing of volatiles initially. The hotter magma ocean persists until solidification at 1.7 Myrs, after which the gray and RT evolution tracks converge. This convergence occurs because the net flux in the RT atmosphere simulation is larger compared to the gray simulation in the end, thereby offsetting initial differences.

In the dry case, the magma ocean begins its evolution essentially without a significant atmosphere, when the outgassing laws of Elkins-Tanton (2008) are used. This dry, hot initial state once again leads to a hotter magma ocean in the simulation with the RT atmosphere model compared to the gray simulation. Unlike in the wet case, the evolution tracks do not converge at the end. This is because the net flux of the RT simulation mostly remains below that of the gray simulation until solidification. Consequently, the magma ocean stage is prolonged with the RT atmosphere model compared to simulations using the gray model.

Table C.1

MagmOc2.0 Results for Earth: Comparison between Elkins-Tanton (2008, Tab. 3, Earth (2000 km)) and MagmOc2.0 at the end of their respective magma ocean stages.

While the lifetime of the Earth magma ocean lifetime may vary with different atmosphere models, this variability does not appear to affect the final volatile content. Simulations using the RT and gray atmosphere model show virtually identical quantities of volatiles in the end (Table C.1).

We further compare the RT grid simulations for H2O-CO2 atmospheres with simulations using the corrected gray atmosphere model. Our analysis reveals that in the mixed H2O-CO2 setup, the corrected gray atmosphere model underestimates the OLR in the initial evolution stage, during which surface temperatures are very high and surface pressures are particularly low ( < 10 bar). This tendency is illustrated in Fig. B.4 (left): The corrected gray atmosphere model consistently underestimates the OLR compared to the RT grid for high surface temperatures (Tsurf ≥ 2000 K) and low pressures (psurf ≤ 26 bar).

However, these discrepancies in flux are limited to the first thousand years of simulation. More importantly, they do not lead to significant differences in later stages of the magma ocean evolution. Throughout the majority of the magma ocean evolution, in both the dry and wet cases, the surface temperature and atmosphere surface pressures agree well between the RT and corrected gray simulation.

Finally, we present RT atmosphere simulations for the dry and wet cases that exclude CO2 (Fig. C.5 dash-dotted line). We find that the absence of CO2 has a minimal influence on the magma ocean evolution, when the outgassing laws of Elkins-Tanton (2008) are used. However, this setup severely suppresses outgassing of CO2, such that it remains a minor constituent throughout the entire magma ocean evolution. In the following, we thus tackle the impact of alternative solubility laws.

thumbnail Fig. C.5

Earth magma ocean simulation for comparison with Elkins-Tanton (2008) with the corrected gray and full RT atmosphere model. Scenarios are as listed in Table C.1. The dotted lines denote results of simulations with the corrected gray atmosphere model. Solid lines denote results of simulations with full RT. The dash-dotted line denote the results of simulations with full RT only with H2O to assess the impact of CO2.

In summary, we generally demonstrate with MagmOc2.O agreement in H2O and CO2 distribution with Elkins-Tanton (2008), and also in solidification times, when the gray atmosphere model without runaway greenhouse limit is used. Different atmosphere models may yield variations in solidification time, confirming Nikolaou et al. (2019). Generally, we find that the differences between atmosphere models are largest during the initial stage and we confirm again that the runaway greenhouse radiation limit has a strong impact on the later magma oecean evolution and thus the solidification times.

C.4 Mixed CO2-H2O atmospheres - Different outgassing laws

More recent solubility laws (e.g., Nikolaou et al. 2019; Lichtenberg et al. 2021) suggest that CO2 should dominate the atmosphere during the initial magma ocean stage due its significantly lower solubility in the magma compared to H2O (Sect. 2.2). To test the impact of such a scenario, we implement in MagmOc2.0 the outgassing laws of Nikolaou et al. (2019). In the following, we denote simulations using the outgassing laws of Elkins-Tanton (2008) as “ET” and simulations using outgassing laws of Nikolaou et al. (2019) as “Ni.”

Table C.2

Parameters for comparing outgassing laws.

To facilitate a comparison of the impact of different outgassing laws with results from previous sections, we once again perform dry and wet case simulations (Table C.2), maintaining an initial magma ocean depth of 2000 km. In the dry-case simulation, volatile inventories are broadly consistent with the setup of the simulations outlined in Nikolaou et al. (2019).

Figure C.6 illustrates the impact of different outgassing laws on the dry case, where major differences are evident for CO2 surface pressures. In the ET-simulation, CO2 is outgassed relatively late and never becomes a dominant component of the atmosphere, consistent with results from the previous section. Conversely, in the Niko-simulations, a significant portion of CO2 is already outgassed at the onset of the magma ocean stage, in agreement with magma ocean simulations of Nikolaou et al. (2019) and Lichtenberg et al. (2021). Consequently, CO2 initially dominates the atmosphere, with substantial H2O outgassing occurring only after 10 000 years of simulation time. Eventually H2O becomes the dominant atmospheric species.

Because the magma ocean evolution in the Ni-simulation begins in the dry case with a thick atmosphere rather than a “bare rock scenario,” the magma ocean cooling is less efficient (Fig. C.6 dash-dotted red line). Consequently, the Ni-simulation reaches solidification later, at 0.8 Myrs, compared to the solidification time of 0.5 Myrs in the ET-simulation (Fig. C.6 solid red line).

We conduct a similar comparison between different outgassing laws and atmosphere models for a volatile-rich wet-case Earth scenario. Here, all initial masses are scaled from the dry case by a factor of ten (Fig. C.7, Table C.2). In this scenario, the thermal evolution of the magma ocean across diverse scenarios is quite similar, converging at around 2 Myrs. Significant differences are again obtained for the H2O and CO2 partial pressures. In the ET-simulations, CO2 consistently remains a minor constituent in the atmosphere. Conversely, in the Ni-simulations, the atmospheric composition initially CO2-dominated and eventually becomes H2O-dominated.

Evaluating again the performance of the corrected gray atmosphere model against the RT model in the Ni-simulations, we find that the former overestimates the OLR of a hot, CO2- dominated atmosphere by a factor of two to three (Figs. C.6 and C.7, dash-dotted purple line). This discrepancy has been already previously noted (Fig. B.4, right panel and Fig. C.2), where it was found that the corrected gray model can deviate from the RT model by up to half an order of magnitude for thick CO2 atmospheres with high surface temperatures (Tsurf > 2500 K). In any case, also here we find that the surface temperature evolution tracks with the corrected gray atmosphere model eventually converge with those with the RT atmosphere. In the dry case, convergence occurs after 10,000 years of simulation time, while in the wet case, convergence occurs after 0.1 Myrs. Despite these discrepancies in surface temperatures, the outgassed H2O and CO2 volatile content generally agrees throughout the entire magma ocean evolution between the Ni-simulations with the corrected gray and RT atmosphere model.

thumbnail Fig. C.6

Earth magma ocean simulations with mixed H2O-CO2 atmospheres to compare the outgassing laws of Elkins-Tanton (2008) (ET) and Nikolaou et al. (2019) (Ni). Here, the results of the dry-case scenarios are shown (see Table C.2). For most scenarios the full RT atmosphere model is used (red). The solid lines denote the dry case scenario of Elkins-Tanton (2008) with 0.21 TO CO2 and the same outgassing laws. The dash-dotted lines denote results of a dry simulation with the outgassing laws from Nikolaou et al. (2019). For the latter scenario, an additional simulation with the corrected gray atmosphere model is shown (purple line). The dotted red line denotes a simulation with only H2O, using the outgassing laws of Nikolaou et al. (2019) and the RT model.

thumbnail Fig. C.7

Earth magma ocean simulation that compares the mixed H2O- CO2 magma oceans of Elkins-Tanton (2008) (ET) and Nikolaou et al. (2019) (Ni) and with a wet case volatile budget (Table C.2). For most scenarios the full RT atmosphere model is used (red). The solid lines denote ET-simulations. The dash-dotted lines denote a wet simulation with the outgassing laws from Nikolaou et al. (2019). For the latter scenario, an additional simulation with the corrected gray atmosphere model is shown (purple lines). The dotted red line denotes a simulation with only H2O using the outgassing laws of Nikolaou et al. (2019) and the RT atmosphere model.

Correctly capturing the amount of outgassed volatiles is crucial for assessing the impact of atmospheric erosion on magma oceans around M dwarf stars. For the Earth simulations, the corrected gray atmosphere model, benchmarked against the numerically more costly RT atmosphere model (Sect. D), appears to adequately capture the volatile evolution in Earth magma oceans. This good performance also holds true with outgassing laws that result in atmospheric composition changing from CO2-dominated to H2O-dominated.

C.5 Mixed CO2-H2O atmospheres - Impact of CO2

In the ET-simulations, CO2 consistently remains a minor atmospheric constituent with a negligible influence on the magma ocean evolution. However, in simulations using the outgassing laws used of Nikolaou et al. (2019), CO2 is initially the dominant atmospheric species that can affect the initial magma ocean stage, as is evidenced by the prolongation of the magma ocean stage in the Ni-simulations compared to the ET-simulations in the dry case. To assess the impact of CO2 in the Ni-simulations more coherently, we conducted several simulations without CO2 (Figs. C.6 and C.7, dotted red curves).

A detailed inspection of the thermal evolution of the dry case shows that a mixed CO2 -H2 O atmosphere exhibits higher thermal emission during the first 0.3 Myrs of the evolution compared to a pure H2O atmosphere. This difference arises because a CO2-rich atmosphere emits more flux than an equivalent H2O- atmosophere for high surface temperatures (Tsurf > 2000 K, see Sections C.2 and B). However, beyond 0.3 Myrs, the system reaches the runaway greenhouse limit that is set primarly determined by the water content in the atmosphere.

As is discussed in Sect. 2.4, the addition of CO2 tends to cool the emitting atmosphere layers, thereby reducing the OLR limit and diminishing thermal cooling in a mixed atmosphere compared to a pure H2O atmosphere. In other words, the addition of CO2 for a sufficiently oxidized mantle (Ortenzi et al. 2020) tends to shorten the magma ocean lifetime outside of the runaway greenhouse limit and tends to extend the solidification time within it. In the dry case, both effects counterbalance each other, resulting in a slightly longer solidification time for the mixed atmosphere (by a few tens of thousands of years or a few percent).

The higher thermal emission of a mixed atmosphere outside of the runaway greenhouse limit is more pronounced when more volatiles are in the system as in the wet case (Fig. C.7, dotted red line). However, even in this scenario, the runaway greenhouse limit reached toward the end of the evolution results after 2 Myrs in a convergence of solidification time compared to a magma ocean model with a pure H2O atmosphere. Upon close examination, the mixed atmosphere exhibits a slightly shorter solidification time, albeit by only a few percent.

Our assessment of a minimal impact of CO2 on the solidification times of magma oceans disagrees with the results of Nikolaou et al. (2019) who report an extension of magma ocean lifetime with additional CO2. These authors used, however, a gray atmosphere model. As is discussed in Appendix B, a gray atmosphere model consistently overestimates thermal emission for thick H2O atmospheres (Fig. B.1) and underestimates thermal emission for thick CO2 atmospheres (Fig. B.3) because it does not adequately capture the thermal emission in a vertically extended atmosphere, as already outlined by Lichtenberg et al. (2021).

C.6 Mixed CO2-H2O atmospheres - Importance of initial conditions

We compare, similarly to Table C.1, the volatile budget at the end of the magma ocean between the ET and Ni-simulations (Table C.3). The most notable difference is the very low remaining CO2 content in the latter cases. This discrepancy can be explained by the different initial conditions, driven by the different CO2 outgassing laws.

thumbnail Fig. C.8

Volatile content in the magma ocean and the atmosphere for an Earth magma ocean simulation with the outgassing laws of Nikolaou et al. (2019), using the RT grid and for the dry case (top panels) and wet case (bottom panels). See Table C.2 for the respective volatile content of the simulations. Shown are melt fraction (left) and volume mixing ratio (VMR, right) for H2O (in blue) and CO2 (in dark red) for the mixed case and the cases with only H2O (dotted blue) and only CO2 (dotted red lines), respectively.

In the model of Elkins-Tanton (2008), CO2 is mostly dissolved in the melt initially, resulting in a relatively large initial melt fraction, FCO , compared to the outgassing laws used by Nikolaou et al. (2019). As the magma ocean solidifies and the thickness of the magma ocean decreases, as a result significant enrichment of volatile mass fraction occurs, assuming no substantial sink terms. Only a very small fraction of CO2 is partitioned from the melt into the solid mantle (Table 1) and atmospheric erosion for Earth is negligible during the magma ocean evolution. Thus, the volatile mass fraction at the end of the magma ocean stage can only be larger than the initial value.

Because CO2 outgassing is suppressed in the ET-simulations, initial values of FH2O${F_{{{\rm{H}}_2}{\rm{O}}}}$ are relatively high already. Consequently, the magma ocean ends with large CO2 mass fractions of 0.2 - 0.7 wt%. Conversely, the Ni-simulations begin with a substantially lower mass fraction in the magma ocean and consequently end the magma ocean stage with CO2 mass fractions of 0.003 - 0.3 wt% , which is up to two orders of magnitudes lower compared to the ET-simulations. The differences are much smaller for H2O, because the ET and Ni-outgassing laws assume both a high solubility of H2O, resulting in similar initial values of FH2O${F_{{{\rm{H}}_2}{\rm{O}}}}$.

We thus conclude that the initial conditions, which determine the initial volatile mass fraction in the magma ocean, are critical for determining the overall volatile budget at the end of the magma ocean phase and how much of the volatiles can be retained in the mantle.

C.7 Mixed CO2-H2O atmospheres - Outgassing feedback

We confirm here that the evolution of the atmospheric CO2 and H2O exhibit a feedback effect, when the mean molecular weight of the atmosphere changes during the magma ocean evolution (Bower et al. 2019). A change in atmospheric composition occurs during the magma ocean evolution because CO2 is less soluble than H2O and thus dominates outgassing at the beginning of the magma ocean stage, as captured by the outgassing laws of Nikolaou et al. (2019) (Sect. C.4). The model of Elkins-Tanton (2008) avoids this effect by suppressing CO2 outgassing and hence these simulations show no such feedback.

Table C.3

MagmOc2.0 results for Earth to compare with Elkins-Tanton (2008, Tab. 3, Earth (2000 km)).

Figure C.8 provides more detailed insights about the outgassing feedback in the Nicolaou (short: Niko) simulations. In the mixed volatile setup (solid lines), the atmospheric composition undergoes drastic changes from CO2-dominated (molecular weight of CO2: µ = 44 g/mol) to H2O-dominated (molecular weight of H2O: µ = 18 g/mol), as also reported by Nikolaou et al. (2019); Bower et al. (2019). Consequently, the mean molecular weight of the atmosphere decreases as the magma ocean solidifies. Additionally, the melt fraction of CO2 decreases as H2O begins to accumulate in the atmosphere (Fig. C.8, left panel, dotted red lines). In the absence of H2O, CO2 outgassing would continue to increase steadily as the magma ocean solidifies. In contrast, the melt fraction of H2O, and thus the partial pressure of H2O, increases in the mixed atmosphere case. In the absence of CO2 , the melt fraction still increases as the mantle solidifies, albeit not as strongly (Fig. C.8, left panel, dotted blue line).

The feedback between atmosphere composition and outgassing affects the volatile distribution at the end of the magma ocean. For the dry case, the final CO2 partial pressure decreases from pCO2=100${p_{{\rm{C}}{{\rm{O}}_2}}} = 100$ bar in a pure CO2 atmosphere to pCO2=67${p_{{\rm{C}}{{\rm{O}}_2}}} = 67$ bar in a mixed atmosphere as H2O becomes the dominant species. The interplay between atmospheric composition and outgassing results in higher H2O outgassing (pH2O=206bar${p_{{{\rm{H}}_2}{\rm{O}}}} = 206{\rm{bar}}$) compared to the scenario with no CO2 in the system, where ’only’ pH2O=206${p_{{{\rm{H}}_2}{\rm{O}}}} = 206}$ bar is outgassed. The changes in H2O and CO2 pressures for the dry case are in first order agreement with Bower et al. (see 2019, their Fig. 7a) for a similar volatile content. Nikolaou et al. (2019) did not account for the outgassing feedback effect and thus report an increase in both, CO2 and H2O partial pressures as the magma ocean solidifies.

We further find, for the wet-case magma ocean, that volatile outgassing is modified similarly to the dry case. We also emphasize again that a drastic change in atmosphere composition, and thus in outgassing, does not occur when the outgassing laws of Elkins-Tanton (2008) are used, where H2O is always the dominant volatile in the atmosphere.

We conclude that MagmOc2.0 adequately accounts for outgassing feedback between H2O and CO2, and that the corrected gray and RT atmosphere model capture thermal emission in vertically extended thick atmosphere. Even more, MagmOc2.0 is well-equipped to also tackle planets in the habitable zone around M dwarf stars, where atmospheric erosion of H2O mainly by XUV photolysis (see e.g., Luger & Barnes 2015) has the potential to further change the composition of a mixed H2O-CO2 atmosphere during the magma ocean stage.

Appendix D Numerical stability and runtime

The code MagmOc2.0 is designed to efficiently explore diverse planetary and stellar scenarios. To achieve this goal, the code solves various sets of ordinary differential equations (ODEs), including equations that describe outgassing (Sect. 2.2). In this appendix, we provide details on numerical stability, convergence tests, and the typical wall clock times of the simulations discussed earlier. We focus here on the albedo=0.75 simulations.

One parameter in these specific ODEs is the decrease in liquid magma ocean dMliqdt${{d{M^{liq}}} \over {dt}}$ as the solidification radius increases drsdt${{d{r_s}} \over {dt}}$, which undergoes, however, phase state transitions for the critical mantle melt fraction ψc = 0.4 (Barth et al. 2021, Sections 2.1.1 and 2.1.2). To tackle numerical instabilities introduced by non-continuous changes of melt fractions FCO2(t)${F_{{\rm{C}}{{\rm{O}}_2}}}(t)$ and FH2O(t)${F_{{{\rm{H}}_2}{\rm{O}}}}(t)$ in time t, we carefully monitor the volatile masses in the different reservoirs of the coupled MOA system (Micystal +Miliq+Miatm $M_i^{{\rm{cystal}}} + M_i^{liq} + M_i^{{\rm{atm}}}$) to ensure that their sum is equal to Mimoa$M_i^{{\rm{moa}}}$ within reasonable limits (differences of less than 5%). Mimoa$M_i^{{\rm{moa}}}$ also has to be initially equal to the prescribed initial volatile mass, Miini$M_i^{{\rm{ini}}}$. To avoid numerical “leakage of mass,” we added two terms, ΔMH2OCorr $\Delta M_{{{\rm{H}}_2}{\rm{O}}}^{{\rm{Corr}}}$ and ΔMCO2Corr $\Delta M_{{\rm{C}}{{\rm{O}}_2}}^{{\rm{Corr}}}$, that provide source or sink terms to the partial derivative of MH2Omoa$M_{{{\rm{H}}_2}{\rm{O}}}^{{\rm{moa}}}$ to ensure mass conservation. These two terms take the form ΔMH2OCorr =MH2OiniMH2Omoa Δtcurr ΔMCO2Corr =MCO2iniMCO2moa Δtcurr $\eqalign{ & \Delta M_{{{\rm{H}}_2}{\rm{O}}}^{{\rm{Corr }}} = {{M_{{{\rm{H}}_2}{\rm{O}}}^{{\rm{ini}}} - M_{{{\rm{H}}_2}{\rm{O}}}^{{\rm{moa }}}} \over {\Delta {t_{{\rm{curr }}}}}} \cr & \Delta M_{{\rm{C}}{{\rm{O}}_2}}^{{\rm{Corr }}} = {{M_{{\rm{C}}{{\rm{O}}_2}}^{{\rm{ini}}} - M_{{\rm{C}}{{\rm{O}}_2}}^{{\rm{moa }}}} \over {\Delta {t_{{\rm{curr }}}}}} \cr} $(D.1)

where ∆tcurr is the current time step during the magma ocean evolution that ensures that the volatile mass budget during runtime does not deviate more than 5% from the initial mass budget. We note that ΔMH2OCorr $\Delta M_{{{\rm{H}}_2}{\rm{O}}}^{{\rm{Corr}}}$ and ΔMCO2$\Delta {M_{{\rm{C}}{{\rm{O}}_2}}}$ are rates in units of [kg/s] to counterbalance numerical mass loss.

thumbnail Fig. D.1

TRAPPIST-1 g magma ocean evolution with different relative accuracies ϵ in the integration during runtime for no additional CO2 in the system (red: ϵ = 10–1, green: ϵ = 10–2, purple: ϵ = 10–3). Top panels show surface temperatures, middle panels show volatile partial pressures (solid: H2O, dotted: O2), bottom panels show net flux for initial water masses of 1 TO, 5 TO, and 100 TO from left to right.

That is, the differentiation of the mass balance equations for volatile i is modified during simulation run time such that: dMimoa dt+ΔMiCorr =dMicrystal dt+dMiliqdt+dMiatmdt.${{dM_i^{{\rm{moa }}}} \over {dt}} + \Delta M_i^{{\rm{Corr }}} = {{dM_i^{{\rm{crystal }}}} \over {dt}} + {{dM_i^{{\rm{liq}}}} \over {dt}} + {{dM_i^{{\rm{atm}}}} \over {dt}}$(D.2)

In terms of the substitution framework (Table A.1), we add ΔMiCorr $\Delta M_i^{{\rm{Corr}}}$ to a1$a_1^\prime $ (for H2O) and a1$a_1^\prime $ (for CO2), respectively. To avoid overcompensation, we limit the correction term to values smaller than 1% of the total volatile budget. Despite these measures, sometimes noticeable fluctuations in the outgassed volatiles can occur due to fluctuations in Fi.

We demonstrate such fluctuations for the TRAPPIST-1 g magma ocean evolution tracks for 1 TO, 5 TO, and 100 TO initial mass H2O with CO2 initial mass scaled by H2O mass with 0x, 0.3x and 1x the inital H2O mass simulated with different relative accuracies ϵ of the Runge-Kutta integrator (Figs. D.1, D.2, D.3).

It is evident that atmospheric pressures and net flux can strongly fluctuate for the lowest integrator accuracy of ϵ = 10–1. In all investigated cases, however, these fluctuations have a negligible impact on the surface temperature and overall volatile evolution. This is confirmed by comparison with simulations of higher numerical accuracy. More precisely, a relative accuracy of ϵ = 10–3 in the Runge-Kutta integrator results in stable simulations that can be used to benchmark a specific evolution track in case of numerical issues.

thumbnail Fig. D.2

TRAPPIST-1 g magma ocean evolution with different relative accuracies ϵ in the integration during runtime for additional CO2, the mass of which is scaled by a factor of 0.3 with the initial H2O masses in the system (red: ϵ = 10–1, green: ϵ = 10–2, purple: ϵ = 10–3). Top panels show surface temperatures, middle panels show volatile partial pressures (solid: H2O, dashed: CO2 dotted: O2), bottom panels show net flux for initial water masses of 1 TO, 5 TO, and 100 TO from left to right.

thumbnail Fig. D.3

TRAPPIST-1 g magma ocean evolution with different relative accuracies ϵ in the integration during runtime for additional Co2, the mass of which is scaled by a factor of 1 with the initial H2O masses in the system (red: ϵ = 10–1, green: ϵ = 10–2, purple: ϵ = 10–3). Top panels show surface temperatures, middle panels show volatile partial pressures (solid: H2O, dashed: CO2 dotted: O2), bottom panels show net flux for initial water masses of 1 TO, 5 TO, and 100 TO from left to right.

We further show the run times of the 18 TRAPPIST-1 g simulations with 1 TO, 5 TO, and 100 TO initial mass of H2O and an equivalent quantity of CO2 mass with different numerical accuracies (Table D.1). The simulations were carried out on a single CPU on a AMD Ryzen Threadripper PRO 5955WX with 16 cores. The magma ocean simulations using the full RT atmosphere model may require up to 28 mins to complete. Simulations with the corrected gray atmosphere model, benchmarked with full RT calculations, are much more efficient, with runtimes of 1 minute and less, depending on numerical accuracy. The difference in runtime between ϵ = 10–3 and 10–1 is just 15 to 20 seconds. Thus, with the corrected gray atmosphere model, the simulations can be carried out with highest numerical stability without excessive numerical costs. MagmOc2.0 as part of the VPLanet-software framework is thus ideally suited to explore the magma ocean evolution for rocky exoplanets in the habitable zone of their host stars, spanning a wide range in fundamental parameters and exploring different processes to account for the large diversities in exoplanet evolution.

Table D.1

Run times of TRAPPIST-1 g simulations with initial CO2 mass equal to the initial H2O mass.

Appendix E Chemical composition

Since stars and their planetary accretion disks are formed by the collapse of the same interstellar dust cloud, the composition of a star can be used as a first estimate for the upper limit of the composition of the accretion disk. However, the stellar composition of TRAPPIST-1 has not yet been determined. We therefore derive elemental abundances for the main planet-forming elements from the stellar metallicity. For this, large-scale astronomical surveys play a crucial role in providing the necessary data for understanding the chemical composition of stars. One such survey is the Galactic Archaeology with HERMES (GALAH) project. The GALAH survey focuses on determining the detailed chemical composition of Milky Way stars, contributing to a better understanding of Galactic chemical evolution. In this study, the third release of the GALAH survey by Buder et al. (2021) serves as the main resource, offering a rich dataset that facilitates the calculation of metallicity-dependent compositional variations. However, the GALAH survey has limitations in providing data for all chemical elements of interest to our study. To address this gap, the Hypatia catalog was incorporated specifically for elements such as nitrogen and sulfur. The Hypatia Catalog, compiled from 84 literature sources, presents spectroscopic abundance data for 50 elements across 3058 stars in the solar neighborhood (within 150 pc of the Sun). Employing a binning strategy with intervals set at [Fe/H] = 0.05, stars were systematically organized to facilitate thorough data analysis. The mean metallicity and chemical composition for each bin was computed, providing a representative value within that specific metallicity range. For TRAPPIST-1, we use a metallic- ity value of 0.04±0.08 (Gillon et al. 2017), which leads to the predicted stellar composition reported in Sect. 3.3.

Table E.1

50% condensation temperatures and volume mixing ration for different molecules in the accretion disk.

Bitsch & Battistini (2020) suggested a stoichiometric model to obtain a first-order estimate on the compositional variation in planetary building material depending on the local temperature within an accretion disk. In this approach, the gas within the accretion disk is assumed to have achieved the state of chemical equilibrium before condensation, with the complete set of molecules preexisting in the gas. Consequently, the relative abundance of molecules can be calculated stochiometrically and based on their condensation temperature Lodders (2003).

For a first estimate on the temperature profile within the accretion disk, we employ a simple power law Williams & Cieza (2011): Tdisk(r)=TS(rxRi)3/4,${T_{disk}}(r) = {T_S}{\left( {{r \over {x{R_i}}}} \right)^{ - 3/4}},$(E.1)

where TS represents the sublimation temperature and is set to 1500 K at the inner edge of the disk Ri (following observations by Dullemond & Monnier 2010). x is a free scaling parameter which we set here to x = 2 for a good agreement of our profile compared to Jorge et al. (2022). Using this estimate for the disk temperature profile allows us to determine the composition of the planetary building blocks at different distances of the star. For this, we extended the model of Bitsch & Battistini (2020) to include nitrogen, aluminum and calcium species. The full set of molecules, their condensation temperatures and stoichiometric calculations are listed in table E.1. Due to the overabundance of hydrogen, the elemental abundances are expressed as X/H.

Appendix F Interior structure models

We integrate the compositional information in our interiorstructure models for all three planets. It is important to note here that while we directly employed the refractory elements condensing out of the accretion disk in our interior-structure model, we allowed for the core-mass fraction (considering here for simplicity a pure iron core) as well as the volatile fraction (considering here for simplicity only water, since it is the most abundant volatile molecule) to vary, to explore the range of potential interior structures in TRAPPIST-1 e, f, and g. For a first justification of our model, we applied our interior structure model without any volatile material to the three innermost planets, and obtain modeled planet radii consistent with observations (see Sect. 3.3). Figure F.1 displays the full range of water content within error bars of planet radii and mass measurements.

thumbnail Fig. F.1

Ranges of possible water fractions of TRAPPIST-1 e, f, and g for different iron mass fractions within error bars of the mass-radius values of Agol et al. (2021).

The interior structure model follows Noack et al. (2016), where we dissect the planet into 1000 shells, which we fill with core, mantle or ice/water material for given planet mass as well as core and water mass fractions. We assume an adiabatic temperature profile from surface to the center of the planet, while adopting a temperature increase at the core-mantle boundary following Stixrude (2014) and Noack & Lasbleis (2020). The surface temperature is set at 300 K, which means that we consider here for simplicity at the surface a liquid water layer. High- pressure ice forms for all three planets for water fractions starting at a few wt-%.

For the core and water thermodynamic properties (such as density or heat capacity), we employ equations of state as outlined in Noack et al. (2016). For the silicate mantle, we improved the approach of that study by developing lookup-tables for all relevant thermodynamic properties for silicate mantles of varying composition calculated with Perple_X (Connolly 2009) and interpolated for our predicted mantle composition as well as local temperature and pressure conditions in each shell. We use a different resolution for our look-up tables for low, intermediate, and high pressures, where we vary temperature in steps of 25 K between 200 and 5000 K and pressure in steps of 1250 bar for pressures below 25 GPa, 1.9 GPa for pressures between 25 and 400 GPa, and 5.5 GPa for pressures above 400 GPa going up to 1500 GPa (which ensures coverage of the entire pressure range encountered in the planetary interior in this study).

References

  1. Acuña, L., Deleuil, M., Mousis, O., et al. 2021, A&A, 647, A53 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  2. Acuña, L., Deleuil, M., & Mousis, O. 2023, A&A, 677, A14 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  3. Agol, E., Dorn, C., Grimm, S. L., et al. 2021, Planet. Sci. J., 2, 1 [NASA ADS] [CrossRef] [Google Scholar]
  4. Aguichine, A., Mousis, O., Devouard, B., & Ronnet, T. 2020, ApJ, 901, 97 [NASA ADS] [CrossRef] [Google Scholar]
  5. Alduchov, O. A., & Eskridge, R. E. 1996, J. Appl. Meteorol., 35, 601 [NASA ADS] [CrossRef] [Google Scholar]
  6. Anisman, L. O., Chubb, K. L., Elsey, J., et al. 2022, J. Quant. Spectrosc. Radiat. Transfer, 278, 108013 [NASA ADS] [CrossRef] [Google Scholar]
  7. Baraffe, I., Homeier, D., Allard, F., & Chabrier, G. 2015, A&A, 577, A42 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  8. Baranov, Y., Lafferty, W., Ma, Q., & Tipping, R. 2008, J. Quant. Spectrosc. Radiat. Transfer, 109, 2291 [NASA ADS] [CrossRef] [Google Scholar]
  9. Barnes, R., Mullins, K., Goldblatt, C., et al. 2013, Astrobiology, 13, 225 [Google Scholar]
  10. Barnes, R., Luger, R., Deitrick, R., et al. 2020, PASP, 132, 024502 [NASA ADS] [CrossRef] [Google Scholar]
  11. Barth, P., Carone, L., Barnes, R., et al. 2021, Astrobiology, 21, 1325 [NASA ADS] [CrossRef] [Google Scholar]
  12. Bitsch, B., & Battistini, C. 2020, A&A, 633, A10 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  13. Bitsch, B., Raymond, S. N., & Izidoro, A. 2019, A&A, 624, A109 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  14. Blank, J. G., Stloper, E. M., & Carroll, M. R. 1993, Earth Planet. Sci. Lett., 119, 27 [CrossRef] [Google Scholar]
  15. Bonati, I., Lichtenberg, T., Bower, D. J., Timpe, M. L., & Quanz, S. P. 2019, A&A, 621, A125 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  16. Bond, J. C., O’Brien, D. P., & Lauretta, D. S. 2010, ApJ, 715, 1050 [Google Scholar]
  17. Bottinga, Y., & Javoy, M. 1990, Chemical Geol., 81, 255 [NASA ADS] [CrossRef] [Google Scholar]
  18. Boukrouche, R., Lichtenberg, T., & Pierrehumbert, R. T. 2021, ApJ, 919, 130 [NASA ADS] [CrossRef] [Google Scholar]
  19. Bower, D. J., Kitzmann, D., Wolf, A. S., et al. 2019, A&A, 631, A103 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  20. Bower, D. J., Hakim, K., Sossi, P. A., & Sanan, P. 2022, Planet. Sci. J., 3, 93 [NASA ADS] [CrossRef] [Google Scholar]
  21. Buder, S., Sharma, S., Kos, J., et al. 2021, MNRAS, 506, 150 [NASA ADS] [CrossRef] [Google Scholar]
  22. Cabral, N., Guilbert-Lepoutre, A., Bitsch, B., Lagarde, N., & Diakite, S. 2023, A&A, 673, A117 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  23. Carone, L., Keppens, R., & Decin, L. 2014, MNRAS, 445, 930 [NASA ADS] [CrossRef] [Google Scholar]
  24. Carroll, M. R., & Holloway, J. R. 1994, Volatiles in Magmas (Berlin, Boston: De Gruyter) [CrossRef] [Google Scholar]
  25. Catling, D. C., & Kasting, J. F. 2017, Atmospheric Evolution on Inhabited and Lifeless Worlds (Cambridge: Cambridge University Press) [CrossRef] [Google Scholar]
  26. Chao, K.-H., deGraffenried, R., Lach, M., et al. 2021, Chemie der Erde / Geochem., 81, 125735 [CrossRef] [Google Scholar]
  27. Chase, M. 1998, J. Phys. Chem. Ref. Data Monograph, 9, 1 [Google Scholar]
  28. Chen, J., & Kipping, D. 2017, ApJ, 834, 17 [Google Scholar]
  29. Chubb, K. L., Rocchetto, M., Al-Refaie, A. F., et al. 2021, A&A, 646, A21 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  30. Cohen, O., Glocer, A., Garraffo, C., et al. 2024, ApJ, 962, 157 [NASA ADS] [CrossRef] [Google Scholar]
  31. Coleman, G. A. L., Leleu, A., Alibert, Y., & Benz, W. 2019, A&A, 631, A7 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  32. Connolly, J. 2009, Geochem. Geophys. Geosyst., 10, 1121 [NASA ADS] [CrossRef] [Google Scholar]
  33. Cox, J., Wagman, D., & Medvedev, V. 1984, CODATA Key Values for Thermodynamics (New York: Hemisphere Publishing Corp.) [Google Scholar]
  34. Debaille, V., Brandon, A. D., O’Neill, C., Yin, Q. Z., & Jacobsen, B. 2009, Nat. Geosci., 2, 548 [NASA ADS] [CrossRef] [Google Scholar]
  35. Deng, J., Du, Z., Karki, B. B., Ghosh, D. B., & Lee, K. K. M. 2020, Nat. Commun., 11, 2007 [NASA ADS] [CrossRef] [Google Scholar]
  36. Dixon, J. E., Stolper, E. M., & Holloway, J. R. 1995, J. Petrol., 36, 1607 [Google Scholar]
  37. Dorn, C., Mosegaard, K., Grimm, S. L., & Alibert, Y. 2018, ApJ, 865, 20 [NASA ADS] [CrossRef] [Google Scholar]
  38. Driscoll, P., & Bercovici, D. 2014, Phys. Earth Planet. Interiors, 236, 36 [CrossRef] [Google Scholar]
  39. Dullemond, C., & Monnier, J. 2010, Ann. Rev. Astron. Astrophys., 48, 205 [Google Scholar]
  40. Elkins-Tanton, L. T. 2008, Earth Planet. Sci. Lett., 271, 181 [Google Scholar]
  41. Ferraz-Mello, S., Rodríguez, A., & Hussmann, H. 2008, Celest. Mech. Dyn. Astron., 101, 171 [Google Scholar]
  42. Gardner, J. E., Hilton, M., & Carroll, M. R. 1999, Earth Planet. Sci. Lett., 168, 201 [CrossRef] [Google Scholar]
  43. Gillon, M., Triaud, A. H., Demory, B.-O., et al. 2017, Nature, 542, 456 [NASA ADS] [CrossRef] [Google Scholar]
  44. Godolt, M., Tosi, N., Stracke, B., et al. 2019, A&A, 625, A12 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  45. Goldblatt, C., Robinson, T. D., Zahnle, K. J., & Crisp, D. 2013, Nat. Geosci., 6, 661 [NASA ADS] [CrossRef] [Google Scholar]
  46. Gordon, I. E., Rothman, L. S., Hargreaves, R. J., et al. 2022, J. Quant. Spectrosc. Radiat. Transfer, 277, 107949 [NASA ADS] [CrossRef] [Google Scholar]
  47. Graham, R. J., Lichtenberg, T., Boukrouche, R., & Pierrehumbert, R. T. 2021, Planet. Sci. J., 2, 207 [NASA ADS] [CrossRef] [Google Scholar]
  48. Greene, T. P., Bell, T. J., Ducrot, E., et al. 2023, Nature, 618, 39 [NASA ADS] [CrossRef] [Google Scholar]
  49. Grimm, S. L., Demory, B.-O., Gillon, M., et al. 2018, A&A, 613, A68 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  50. Hamano, K., Abe, Y., & Genda, H. 2013, Nature, 497, 607 [NASA ADS] [CrossRef] [Google Scholar]
  51. Herbort, O., Woitke, P., Helling, C., & Zerkle, A. 2020, A&A, 636, A71 [EDP Sciences] [Google Scholar]
  52. Hier-Majumder, S., & Hirschmann, M. M. 2017, Geochem. Geophys. Geosyst., 18, 3078 [Google Scholar]
  53. Holtz, F., Behrens, H., Dingwell, D. B., & Johannes, W. 1995, Am. Mineral., 80, 94 [NASA ADS] [CrossRef] [Google Scholar]
  54. Hunten, D. M., Pepin, R. O., & Walker, J. C. G. 1987, Icarus, 69, 532 [NASA ADS] [CrossRef] [Google Scholar]
  55. Ikoma, M., Elkins-Tanton, L., Hamano, K., & Suckale, J. 2018, Space Sci. Rev., 214, 76 [NASA ADS] [CrossRef] [Google Scholar]
  56. Johansen, A., Ronnet, T., Schiller, M., Deng, Z., & Bizzarro, M. 2023, A&A, 671, A76 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  57. Johnson, K. T. M., & Dick, H. J. B. 1992, J. Geophys. Res., 97, 9219 [NASA ADS] [CrossRef] [Google Scholar]
  58. Johnstone, C. P., Bartel, M., & Güdel, M. 2021, A&A, 649, A96 [EDP Sciences] [Google Scholar]
  59. Jorge, D. M., Kamp, I., Waters, L., Woitke, P., & Spaargaren, R. J. 2022, A&A, 660, A85 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  60. Katyal, N., Nikolaou, A., Godolt, M., et al. 2019, ApJ, 875, 31 [NASA ADS] [CrossRef] [Google Scholar]
  61. Katyal, N., Ortenzi, G., Lee Grenfell, J., et al. 2020, A&A, 643, A81 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  62. Kopparapu, R. K., Ramirez, R., Kasting, J. F., et al. 2013, ApJ, 765, 131 [NASA ADS] [CrossRef] [Google Scholar]
  63. Krissansen-Totton, J. 2023, ApJ, 951, L39 [NASA ADS] [CrossRef] [Google Scholar]
  64. Krissansen-Totton, J., & Fortney, J. J. 2022, ApJ, 933, 115 [NASA ADS] [CrossRef] [Google Scholar]
  65. Lammer, H., Zerkle, A. L., Gebauer, S., et al. 2018, A&A Rev., 26, 2 [NASA ADS] [CrossRef] [Google Scholar]
  66. Lebrun, T., Massol, H., ChassefièRe, E., et al. 2013, J. Geophys. Res. Planets, 118, 1155 [NASA ADS] [CrossRef] [Google Scholar]
  67. Leconte, J., Chabrier, G., Baraffe, I., & Levrard, B. 2010, A&A, 516, A64 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  68. Lensky, N. G., Niebo, R. W., Holloway, J. R., Lyakhovsky, V., & Navon, O. 2006, Earth Planet. Sci. Lett., 245, 278 [CrossRef] [Google Scholar]
  69. Lichtenberg, T., Bower, D. J., Hammond, M., et al. 2021, J. Geophys. Res. Planets, 126, e06711 [NASA ADS] [CrossRef] [Google Scholar]
  70. Lim, O., Benneke, B., Doyon, R., et al. 2023, ApJ, 955, L22 [NASA ADS] [CrossRef] [Google Scholar]
  71. Lincowski, A. P., Meadows, V. S., Zieba, S., et al. 2023, ApJ, 955, L7 [NASA ADS] [CrossRef] [Google Scholar]
  72. Liu, Y., Zhang, Y., & Behrens, H. 2005, J. Volcanol. Geothermal Res., 143, 219 [NASA ADS] [CrossRef] [Google Scholar]
  73. Lodders, K. 2003, ApJ, 591, 1220 [Google Scholar]
  74. Luger, R., & Barnes, R. 2015, Astrobiology, 15, 119 [Google Scholar]
  75. Marcq, E., Salvador, A., Massol, H., & Davaille, A. 2017, J. Geophys. Res. Planets, 122, 1539 [Google Scholar]
  76. Mason, E. A., & Marrero, T. R. 1970, Adv. Atomic Mol. Phys., 6, 155 [Google Scholar]
  77. Maurice, M., Tosi, N., Samuel, H., et al. 2017, J. Geophys. Res. Planets, 122, 577 [NASA ADS] [CrossRef] [Google Scholar]
  78. Miguel, Y., Cridland, A., Ormel, C. W., Fortney, J. J., & Ida, S. 2020, MNRAS, 491, 1998 [NASA ADS] [Google Scholar]
  79. Mlawer, E. J., Payne, V. H., Moncet, J.-L., et al. 2012, Phil. Trans. R. Soc. A Math. Phys. Eng. Sci., 370, 2520 [CrossRef] [Google Scholar]
  80. Mollière, P., Wardenier, J. P., van Boekel, R., et al. 2019, A&A, 627, A67 [Google Scholar]
  81. Moore, G., & Carmichael, I. S. E. 1998, Contrib. Mineral. Petrol., 130, 304 [NASA ADS] [CrossRef] [Google Scholar]
  82. Moore, K., Cowan, N. B., & Boukaré, C.-É. 2023, MNRAS, 526, 6235 [NASA ADS] [CrossRef] [Google Scholar]
  83. Mousis, O., Aguichine, A., Bouquet, A., et al. 2021, Planet. Sci. J., 2, 72 [NASA ADS] [CrossRef] [Google Scholar]
  84. Müller, J., Bitsch, B., & Schneider, A. D. 2024a, A&A, 688, A139 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  85. Müller, S., Baron, J., Helled, R., Bouchy, F., & Parc, L. 2024b, A&A, 686, A296 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  86. Mysen, B. O., Arculus, R. J., & Eggler, D. H. 1975, Contrib. Mineral. Petrol., 53, 227 [NASA ADS] [CrossRef] [Google Scholar]
  87. Nikolaou, A., Katyal, N., Tosi, N., et al. 2019, ApJ, 875, 11 [Google Scholar]
  88. Noack, L., & Lasbleis, M. 2020, A&A, 638, A129 [EDP Sciences] [Google Scholar]
  89. Noack, L., Höning, D., Rivoldini, A., et al. 2016, Icarus, 277, 215 [NASA ADS] [CrossRef] [Google Scholar]
  90. Noack, L., Rivoldini, A., & Van Hoolst, T. 2017, Phys. Earth Planet. Int., 269, 40 [CrossRef] [Google Scholar]
  91. Odintsova, T. A., Tretyakov, M. Y., Simonova, A. A., et al. 2020, J. Mol. Struc., 1210, 128046 [NASA ADS] [CrossRef] [Google Scholar]
  92. O’Neill, H., & Palme, H. 1998, The Earth’s Mantle: Composition, Structure, and Evolution, ed. I. Jackson (Cambridge: Cambridge University Press) [Google Scholar]
  93. Ormel, C. W., Liu, B., & Schoonenberg, D. 2017, A&A, 604, A1 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  94. O’Rourke, J. G. 2020, Geophys. Res. Lett., 47, e86126 [Google Scholar]
  95. Ortenzi, G., Noack, L., Sohl, F., et al. 2020, Sci. Rep., 10, 10907 [NASA ADS] [CrossRef] [Google Scholar]
  96. Pahlevan, K., Schaefer, L., Elkins-Tanton, L. T., Desch, S. J., & Buseck, P. R. 2022, Earth Planet. Sci. Lett., 595, 117772 [CrossRef] [Google Scholar]
  97. Pan, V., Holloway, J. R., & Hervig, R. L. 1991, Geochim. Cosmochim. Acta, 55, 1587 [NASA ADS] [CrossRef] [Google Scholar]
  98. Papale, P. 1997, Contrib. Mineral. Petrol., 126, 237 [NASA ADS] [CrossRef] [Google Scholar]
  99. Paynter, D. J., Ptashnik, I. V., Shine, K. P., et al. 2009, J. Geophys. Res. Atmos., 114, D21301 [CrossRef] [Google Scholar]
  100. Pichierri, G., Morbidelli, A., Batygin, K., & Brasser, R. 2024, Nat. Astron., 8, 1408 [NASA ADS] [CrossRef] [Google Scholar]
  101. Pierrehumbert, R. T. 2010, Principles of Planetary Climate (Cambridge: Cambridge University Press) [CrossRef] [Google Scholar]
  102. Piso, A.-M. A., Öberg, K. I., Birnstiel, T., & Murray-Clay, R. A. 2015, ApJ, 815, 109 [Google Scholar]
  103. Pluriel, W., Marcq, E., & Turbet, M. 2019, Icarus, 317, 583 [NASA ADS] [CrossRef] [Google Scholar]
  104. Ptashnik, I. V., McPheat, R. A., Shine, K. P., Smith, K. M., & Williams, R. G. 2011, J. Geophys. Res. Atmos., 116, D16305 [CrossRef] [Google Scholar]
  105. Raymond, S. N., Izidoro, A., Bolmont, E., et al. 2022, Nat. Astron., 6, 80 [NASA ADS] [CrossRef] [Google Scholar]
  106. Ribas, I., Guinan, E. F., Güdel, M., & Audard, M. 2005, ApJ, 622, 680 [Google Scholar]
  107. Richey-Yowell, T., Shkolnik, E. L., Loyd, R. O. P., et al. 2022, ApJ, 929, 169 [Google Scholar]
  108. Ros, K., & Johansen, A. 2013, A&A, 552, A137 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  109. Salvador, A., & Samuel, H. 2023, Icarus, 390, 115265 [NASA ADS] [CrossRef] [Google Scholar]
  110. Samuel, H., Drilleau, M., Rivoldini, A., et al. 2023, Nature, 622, 712 [NASA ADS] [CrossRef] [Google Scholar]
  111. Schaefer, L., Wordsworth, R. D., Berta-Thompson, Z., & Sasselov, D. 2016, ApJ, 829, 63 [NASA ADS] [CrossRef] [Google Scholar]
  112. Schlecker, M., Apai, D., Lichtenberg, T., et al. 2024, Planet. Sci. J., 5, 3 [NASA ADS] [CrossRef] [Google Scholar]
  113. Schneider, A. D., & Bitsch, B. 2021, A&A, 654, A72 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  114. Schoonenberg, D., Liu, B., Ormel, C. W., & Dorn, C. 2019, A&A, 627, A149 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  115. Sergeev, D. E., Fauchez, T. J., Turbet, M., et al. 2022, Planet. Sci. J., 3, 212 [NASA ADS] [CrossRef] [Google Scholar]
  116. Sergeev, D. E., Boutle, I. A., Lambert, F. H., et al. 2024, ApJ, 970, 7 [Google Scholar]
  117. Shine, K. P., Campargue, A., Mondelain, D., et al. 2016, J. Mol. Spectr., 327, 193 [NASA ADS] [CrossRef] [Google Scholar]
  118. Silver, L. A., Ihinger, P. D., & Stolper, E. 1990, Contrib. Mineral. Petrol., 104, 142 [NASA ADS] [CrossRef] [Google Scholar]
  119. Stixrude, L. 2014, Phil. Trans. R. Soc. A Math. Phys. Eng. Sci., 372, 20130076 [CrossRef] [Google Scholar]
  120. Stolper, E., & Holloway, J. R. 1988, Earth Planet. Sci. Lett., 87, 397 [CrossRef] [Google Scholar]
  121. Stüeken, E. E., Som, S. M., Claire, M., et al. 2020, Space Sci. Rev., 216, 31 [CrossRef] [Google Scholar]
  122. Tian, F., & Ida, S. 2015, Nat. Geosci., 8, 177 [NASA ADS] [CrossRef] [Google Scholar]
  123. Turbet, M., Ehrenreich, D., Lovis, C., Bolmont, E., & Fauchez, T. 2019, A&A, 628, A12 [NASA ADS] [CrossRef] [EDP Sciences] [Google Scholar]
  124. Turbet, M., Bolmont, E., Chaverot, G., et al. 2021, Nature, 598, 276 [CrossRef] [Google Scholar]
  125. Unterborn, C. T., Hinkel, N. R., & Desch, S. J. 2018, Res. Notes Am. Astron. Soc., 2, 116 [Google Scholar]
  126. Watson, A. J., Donahue, T. M., & Walker, J. C. G. 1981, Icarus, 48, 150 [NASA ADS] [CrossRef] [Google Scholar]
  127. Way, M. J., Ernst, R. E., & Scargle, J. D. 2022, Planet. Sci. J., 3, 92 [NASA ADS] [CrossRef] [Google Scholar]
  128. Williams, J. P., & Cieza, L. A. 2011, Ann. Rev. Astron. Astrophys., 49, 67 [CrossRef] [Google Scholar]
  129. Wordsworth, R. D., Schaefer, L. K., & Fischer, R. A. 2018, AJ, 155, 195 [NASA ADS] [CrossRef] [Google Scholar]
  130. Yamashita, S. 1999, J. Petrol., 40, 1497 [NASA ADS] [CrossRef] [Google Scholar]
  131. Yang, J., Boué, G., Fabrycky, D. C., & Abbot, D. S. 2014, ApJ, 787, L2 [NASA ADS] [CrossRef] [Google Scholar]
  132. Young, E. D., Shahar, A., & Schlichting, H. E. 2023, Nature, 616, 306 [NASA ADS] [CrossRef] [Google Scholar]
  133. Zahnle, K. J., & Kasting, J. F. 1986, Icarus, 68, 462 [NASA ADS] [CrossRef] [Google Scholar]
  134. Zahnle, K. J., Lupu, R., Dobrovolskis, A., & Sleep, N. H. 2015, Earth Planet. Sci. Lett., 427, 74 [CrossRef] [Google Scholar]
  135. Zahnle, K. J., Lupu, R., Catling, D. C., & Wogan, N. 2020, Planet. Sci. J., 1, 11 [NASA ADS] [CrossRef] [Google Scholar]
  136. Zieba, S., Kreidberg, L., Ducrot, E., et al. 2023, Nature, 620, 746 [NASA ADS] [CrossRef] [Google Scholar]

1

The transition between bona-fide rocky planets and volatile-rich planets is predicted to occur at 2 (Chen & Kipping 2017) or 4 Earth masses (Müller et al. 2024b). In this work we discuss planets well below 2 Earth masses.

2

1 TO of water is 1.39 × 1021 kg of H2O.

3

More specifically, H2O is supercritical for T > 646 K and p > 221 bar and CO2 is supercritical for T > 304 K and p > 74 bar.

4

Note that the assumption of a dry lapse rate is no longer valid beyond the triple point of H2O and CO2. However, for consistency, we maintain the dry adiabatic assumption even for extreme temperatures and pressures. We acknowledge the crossing of the triple points as a limitation and area for potential improvement in Sect. 7.

5

A more detailed grid displaying critical properties at the end of the magma ocean can be found on Zenodo.

6

However, it’s important to consider that we are exploring a large range of initial water masses and that the small percentages of remaining water in the mantle has to be scaled with initial water mass to diagnose how much total water mass is in the mantle.

7

According to Kopparapu et al. (2013).

8

Using κ0(H2O) = 0.01 m2/kg as in Elkins-Tanton (2008); Barth et al. (2021) yields larger deviations (not shown).

9

We have verified that MagmOc1.0 and MagmOc2.0 always yield the same results for the pure H2O outgassing mode and the gray atmosphere model.

All Tables

Table 1

Parameters of the geophysical model for MagmOcV2.0

Table 2

Parameters of the volatile model for MagmOcV2.0.

Table 3

Opacity sources used in this work.

Table 4

Definition of the thermal emission grid.

Table 5

Physical and run parameters for TRAPPIST-1 e, f, and g used by MagmOc2.0.

Table 6

Combinations of core mass fractions and water mass fractions for TRAPPIST-1 e, f, and g that match the observed radii from Agol et al. (2021).

Table A.1

Placeholder for differentiation of mass balance equations.

Table C.1

MagmOc2.0 Results for Earth: Comparison between Elkins-Tanton (2008, Tab. 3, Earth (2000 km)) and MagmOc2.0 at the end of their respective magma ocean stages.

Table C.2

Parameters for comparing outgassing laws.

Table C.3

MagmOc2.0 results for Earth to compare with Elkins-Tanton (2008, Tab. 3, Earth (2000 km)).

Table D.1

Run times of TRAPPIST-1 g simulations with initial CO2 mass equal to the initial H2O mass.

Table E.1

50% condensation temperatures and volume mixing ration for different molecules in the accretion disk.

All Figures

thumbnail Fig. 1

Schematic depicting the setup of the volatile exchange during initialization (t=0) and run time. For initialization, a surface temperature of 4000 K and a completely molten magma ocean was assumed, where the dissolved volatiles are in balance with the outgassed volatile content, set by the volatile melt fraction, Fi. As the magma ocean solidifies, part of the volatile budget is deposited in the solid mantle. Furthermore, atmospheric escape can remove H2O. These two sink terms thus reduce the amount of a volatile available in the fully coupled MOA system, Mimoa$M_i^{moa{\rm{}}}$. The full overview of all included processes, including radiogenic heating, is shown in Barth et al. (2021, Fig. 1).

In the text
thumbnail Fig. 2

Vertically extended pressure-temperature (pgas, Tgas) profiles for psur f = 260 bar and Tsur f = 1000 K and different water mixing ratios, xH2O${x_{{{\rm{H}}_2}{\rm{O}}}}$ (left). The solid black and solid red line denote 100% CO2 and 100% H2O atmosphere composition, respectively. The profiles converge in the upper atmosphere to the condensation curve, where we assume equilibrium between condensation and evaporation of H2O and CO2 (supersaturation ratio S = 1). Change in emission (right) from a pure H2O atmosphere (red line) to CO2-dominated (black line).

In the text
thumbnail Fig. 3

Integrated OLR for psurf = 260 bar and Tsurf = 1000 K and different H2O to CO2 content. Mixing CO2 into the steam atmosphere mainly acts to cool the upper atmosphere layers toward the CO2 condensation curve. Only for xH2O<103${x_{{{\rm{H}}_2}{\rm{O}}}}\, < \,{10^{ - 3}}$ (that is, xH2O>0.999${x_{{{\rm{H}}_2}{\rm{O}}}}\, > \,0.999$) will the overall thermal emission increase again as the steep adiabat of a CO2- dominated atmosphere extends into the upper thermally emitting part of the atmosphere.

In the text
thumbnail Fig. 4

Magma ocean evolution for initial H2O of 1 TO, 5 TO, and 100 TO , respectively. All scenarios are for albedo=0. Initial CO2 mass content is scaled relative to the H2O content by a factor of 0, 0.3, and 1, denoted by red, green, and purple lines, respectively. Surface temperature (left) and volatile content evolution (right) are shown.H2O, CO2, and O2 are denoted by solid, dashed, and dotted lines, respectively.

In the text
thumbnail Fig. 5

TRAPPIST-1 g: Net flux (OLR-ASR) evolution for the magma ocean scenarios shown in Fig. 4; that is, for 1 TO, 5 TO, and 100 TO initial H2O and various initial CO2 mass fractions for an albedo of 0. We note simulations with an assumed albedo of 0.75 (not shown) are qualitatively similar but have a shorter regime with net flux limited by the runaway greenhouse radiation limit.

In the text
thumbnail Fig. 6

Overview of total atmospheric desiccation times of TRAPPIST-1 e (red), f (green) and g (purple) for a dry composition with 1–10 TO of initial H2O, assuming an albedo of 0.75 (left) and 0 (right). Colored shaded regions denote the time, when the respective planet enters the habitable zone. Solid lines represent scenario with pure H2O atmospheres, dash-dotted lines denote scenarios with added CO2 scaled by 0.3x compared to the initial H2O mass, and dotted lines show scenarios with added CO2 equal to the initial water mass. The solid circles denote for a given H2O-CO2 mass fraction and planet the maximum initial water mass, for which total atmosphere desiccation occurs before the respective planet enters the habitable zone. Please note the change in y-axis scale between the plots.

In the text
thumbnail Fig. 7

Overview of magma ocean solidification time (left) and the remaining water in the solid mantle compared to initial water mass in percent (right) for TRAPPIST-1 e (red), f (green), and g (purple), considering various CO2 contents for albedo 0.75 (top panels) and albedo 0 (bottom panels). Solid lines represent scenarios with pure H2O atmospheres, dash-dotted lines denote scenarios with added CO2 scaled by 0.3x compared to the initial H2O mass, and dotted lines show scenarios with added CO2 equal to the initial water mass. Colored shaded regions denote the time, when the respective planet enters the habitable zone.

In the text
thumbnail Fig. 8

Predicted composition of the building blocks of the different TRAPPIST-1 planets based on the stellar metallicity and the assumption that the planets did not strongly migrate during accretion. Due to secondary processes including collisions and stripping of material, as well as melting and evaporation processes, especially for the outer planets, the final planetary composition is expected to be considerably less volatile-rich than predicted here for the planetary building blocks.

In the text
thumbnail Fig. 9

Predicted increase in planet radius (in km and %) as well as resulting increase in density in % calculated for the sub-atmospheric planet layers (i.e., core, mantle, and water/ice layers) when considering CO2 or H2O atmospheres of variable average atmospheric temperature and atmospheric pressures up to 1000 bar.

In the text
thumbnail Fig. B.1

Outgoing thermal flux for a pure H2O atmosphere on an Earth gravity planet (𝑔 = 9.81 m/s2) based on the analytic gray atmosphere and full RT calculations under different assumptions. The results of the full RT calculations in a vertically extended atmosphere (Full RT) with heat capacity varying with temperature, cP(T), are used for benchmarking (solid lines). The models are compared for the same surface temperatures and pressures, where different surface pressures are indicated by variations in color. Left panel: Results of full RT calculations with temperature-dependent cP (solid lines), full RT with constant cP (dashed lines), and the gray atmosphere model (dash-dotted lines). Right panel: Results of the corrected gray atmosphere (dash-dotted lines) compared to the full RT.

In the text
thumbnail Fig. B.2

Left: Specific heat capacity versus temperature for constant cP (blue) and cP(T) (dark orange). Right: H2O pressure-temperature profiles for psurf = 260 bar and Tsurf = 1000 K (dash-dotted) and 2000 K (solid). Profiles assuming constant cP are shown in blue and profiles assuming temperature dependent cP are shown in dark orange. Note that for a 1000 K surface, the difference between the profiles is negligible.

In the text
thumbnail Fig. B.3

Outgoing thermal flux for a pure CO2 atmosphere on an Earth gravity planet (𝑔 = 9.81 m/s2) are shown for gray atmosphere model with different assumptions. The results of the full RT in a vertically extended atmosphere (“Full RT”) is used for benchmarking (solid lines). The models are compared for the same surface temperatures and pressures, where different surface pressures are indicated by variations in color. Left panel: Results of the gray atmosphere model is shown for ĸ0(CO2) = 0.001 m2/kg (dashed lines) and ĸ0(CO2) = 0.05 m2/kg (dash- dotted lines). Right panel: Results of the corrected gray atmosphere is shown (dash-dotted lines) in comparison to the Full RT calculations.

In the text
thumbnail Fig. B.4

Outgoing thermal flux for a mixed H2O-CO2 atmosphere on an Earth gravity planet (𝑔 = 9.81 m/s2) based on the corrected gray atmosphere (dashed lines) and full RT calculations with vertically extended atmosphere (solid lines) for the same surface temperatures and pressures, respectively. Different surface pressures are indicated by variations in color. Results are shown for two examples of a two component atmosphere with water volume mixing ratio xH2O=0.75${{x}_{{{\text{H}}_{\text{2}}}\text{O}}}=0.75$ (left panel) and xH2O=1014${{x}_{{{\text{H}}_{\text{2}}}\text{O}}}={{10}^{-14}}$(right panel), respectively.

In the text
thumbnail Fig. B.5

OLR limit for Earth with a two component H2O-CO2 atmosphere for different H2O volume mixing ratios using full RT (circles) vs. the parametric fit defined in Eq. (B.13) (solid line).

In the text
thumbnail Fig. B.6

Left panel: Emission versus wavelength for a pure H2O atmosphere in the runaway greenhouse limit for different surface gravities. Right panel: OLR limit for a two component H2O-CO2 atmosphere and different surface gravities (circles) including the adjusted parametric fits (solid lines) using Eq. (B.15).

In the text
thumbnail Fig. C.1

MagmOc2.0 simulations for Earth and an initial water content of 5 TO (top panel) and 1.02 TO (bottom panel). Depicted from left to right in each panel are: Surface temperature, H2O atmosphere surface pressure and net flux at the top of the atmosphere (FOLRFASR). Three different atmosphere models are used: The gray atmosphere model (solid red line), the corrected gray model (dotted red line), and the RT model (solid purple line).

In the text
thumbnail Fig. C.2

Simulation of MagmOcV2.0 for Earth and 14.7 TO CO2 content for a 2000 km deep magma ocean. Depicted from left to right: Surface temperature, H2O atmosphere surface pressure, and net flux at the top of the atmosphere (FOLRFASR) for three different atmosphere models. These models are the gray model without any correction (solid red line), the corrected gray model (dotted red line), and the RT adaptation (solid purple line).

In the text
thumbnail Fig. C.3

Pure CO2 Earth magma ocean simulation with 14.7 TO of initial CO2 mass following Elkins-Tanton (2008) and using the RT atmosphere model (solid black line). Two RT thermal emission grid lines for psurf = 655 bar (blue solid line) and psurf = 4122 bar (orange solid line) are shown, respectively, as specified in Sect. 2.5. The corresponding thermal emission for the gray atmosphere model as used in Elkins-Tanton (2008) is shown for comparison (dashed lines). The thermal emission deviates between the RT and gray atmosphere model by up to two orders of magnitude during the magma ocean stage in this case with Tsurf = 1500-3000 K.

In the text
thumbnail Fig. C.4

Earth magma ocean simulation setup for comparison with Elkins-Tanton (2008) with the traditional gray model but without greenhouse limit and the full RT grid model. Scenarios are as listed in Table C.1. The solid lines denote results of simulations with the gray atmosphere model. The dotted lines denote results of simulations with full RT. We note again that we show the net flux, which is equal ASR- OLR.

In the text
thumbnail Fig. C.5

Earth magma ocean simulation for comparison with Elkins-Tanton (2008) with the corrected gray and full RT atmosphere model. Scenarios are as listed in Table C.1. The dotted lines denote results of simulations with the corrected gray atmosphere model. Solid lines denote results of simulations with full RT. The dash-dotted line denote the results of simulations with full RT only with H2O to assess the impact of CO2.

In the text
thumbnail Fig. C.6

Earth magma ocean simulations with mixed H2O-CO2 atmospheres to compare the outgassing laws of Elkins-Tanton (2008) (ET) and Nikolaou et al. (2019) (Ni). Here, the results of the dry-case scenarios are shown (see Table C.2). For most scenarios the full RT atmosphere model is used (red). The solid lines denote the dry case scenario of Elkins-Tanton (2008) with 0.21 TO CO2 and the same outgassing laws. The dash-dotted lines denote results of a dry simulation with the outgassing laws from Nikolaou et al. (2019). For the latter scenario, an additional simulation with the corrected gray atmosphere model is shown (purple line). The dotted red line denotes a simulation with only H2O, using the outgassing laws of Nikolaou et al. (2019) and the RT model.

In the text
thumbnail Fig. C.7

Earth magma ocean simulation that compares the mixed H2O- CO2 magma oceans of Elkins-Tanton (2008) (ET) and Nikolaou et al. (2019) (Ni) and with a wet case volatile budget (Table C.2). For most scenarios the full RT atmosphere model is used (red). The solid lines denote ET-simulations. The dash-dotted lines denote a wet simulation with the outgassing laws from Nikolaou et al. (2019). For the latter scenario, an additional simulation with the corrected gray atmosphere model is shown (purple lines). The dotted red line denotes a simulation with only H2O using the outgassing laws of Nikolaou et al. (2019) and the RT atmosphere model.

In the text
thumbnail Fig. C.8

Volatile content in the magma ocean and the atmosphere for an Earth magma ocean simulation with the outgassing laws of Nikolaou et al. (2019), using the RT grid and for the dry case (top panels) and wet case (bottom panels). See Table C.2 for the respective volatile content of the simulations. Shown are melt fraction (left) and volume mixing ratio (VMR, right) for H2O (in blue) and CO2 (in dark red) for the mixed case and the cases with only H2O (dotted blue) and only CO2 (dotted red lines), respectively.

In the text
thumbnail Fig. D.1

TRAPPIST-1 g magma ocean evolution with different relative accuracies ϵ in the integration during runtime for no additional CO2 in the system (red: ϵ = 10–1, green: ϵ = 10–2, purple: ϵ = 10–3). Top panels show surface temperatures, middle panels show volatile partial pressures (solid: H2O, dotted: O2), bottom panels show net flux for initial water masses of 1 TO, 5 TO, and 100 TO from left to right.

In the text
thumbnail Fig. D.2

TRAPPIST-1 g magma ocean evolution with different relative accuracies ϵ in the integration during runtime for additional CO2, the mass of which is scaled by a factor of 0.3 with the initial H2O masses in the system (red: ϵ = 10–1, green: ϵ = 10–2, purple: ϵ = 10–3). Top panels show surface temperatures, middle panels show volatile partial pressures (solid: H2O, dashed: CO2 dotted: O2), bottom panels show net flux for initial water masses of 1 TO, 5 TO, and 100 TO from left to right.

In the text
thumbnail Fig. D.3

TRAPPIST-1 g magma ocean evolution with different relative accuracies ϵ in the integration during runtime for additional Co2, the mass of which is scaled by a factor of 1 with the initial H2O masses in the system (red: ϵ = 10–1, green: ϵ = 10–2, purple: ϵ = 10–3). Top panels show surface temperatures, middle panels show volatile partial pressures (solid: H2O, dashed: CO2 dotted: O2), bottom panels show net flux for initial water masses of 1 TO, 5 TO, and 100 TO from left to right.

In the text
thumbnail Fig. F.1

Ranges of possible water fractions of TRAPPIST-1 e, f, and g for different iron mass fractions within error bars of the mass-radius values of Agol et al. (2021).

In the text

Current usage metrics show cumulative count of Article Views (full-text article views including HTML views, PDF and ePub downloads, according to the available data) and Abstracts Views on Vision4Press platform.

Data correspond to usage on the plateform after 2015. The current usage metrics is available 48-96 hours after online publication and is updated daily on week days.

Initial download of the metrics may take a while.