License: arXiv.org perpetual non-exclusive license
arXiv:2402.06933v1 [astro-ph.EP] 10 Feb 2024

The Diffusion Limit of Photoevaporation in Primordial Planetary Atmospheres

Darius Modirrousta-Galian Yale University, Department of Earth and Planetary Sciences, 210 Whitney Avenue, New Haven, CT 06511, USA; darius.modirrousta-galian@yale.edu Jun Korenaga Yale University, Department of Earth and Planetary Sciences, 210 Whitney Avenue, New Haven, CT 06511, USA; darius.modirrousta-galian@yale.edu

Accepted: ApJ

Abstract

Photoevaporation is thought to play an important role in the early planetary evolution. In this study, we investigate the diffusion limit of X-ray and ultraviolet induced photoevaporation in primordial atmospheres. We find that compositional fractionation resulting from mass loss is more significant than currently recognized because it is controlled by the conditions at the top of the atmosphere, where particle collisions are less frequent. Such fractionation at the top of the atmosphere develops a compositional gradient that extends downward. Mass outflow eventually reaches a steady state in which hydrogen loss is diffusion limited. We derive new analytic expressions for the diffusion-limited mass loss rate and the crossover mass.

Keywords: Atmospheric dynamics – Atmospheric structure — Aeronomy

1 Introduction

Stars are most luminous in their high-energy bands in their first few hundred millions years after formation (Penz et al. 2008; Penz & Micela 2008; Sanz-Forcada et al. 2011). Newly formed planets with primordial atmospheres efficiently absorb X-ray and ultraviolet (XUV) photons, triggering photoevaporation and the gradual loss of their hydrogen reservoirs. Observations suggest that atmospheric evaporation is prevalent, as evidenced by detections of leaking hydrogen (Joshi et al. 2019) and Xenon isotopic ratios on Earth (Porcelli & Pepin 2014), and by exoplanet population trends such as the bimodal radial distribution and sub-Jovian desert (Fulton et al. 2017). The physics of XUV-induced mass loss is however, less clear, with two major models being suggested: inviscid hydrodynamic outflow (Tian et al. 2005) and diffusion-limited escape (Zahnle et al. 2019). The first model builds on Parker wind theory (Parker 1958) and applies it to planetary atmospheres, suggesting that mass loss occurs through free advection and rapidly erodes hydrogen-rich atmospheres (Kubyshkina et al. 2018; Caldiroli et al. 2021). In contrast, diffusion-limited escape posits that mass loss occurs preferentially for lighter species such as hydrogen. Interactions between the fast-moving hydrogen and the slow-moving heavier species result in an overall decrease in the hydrogen outflow rate, with mass loss being limited by momentum diffusion between different species (Zahnle et al. 2019).

When a hydrogen-rich atmosphere is exposed to X-ray and ultraviolet irradiation, a steep conducting temperature inversion develops in the thermosphere (Gross 1972). Temperatures become sufficiently high for gases to become gravitationally unbound (Öpik 1963) and be lost through hydrodynamic winds (Sekiya et al. 1980, 1981). The question is then whether interactions between different species are indeed significant and mass loss becomes diffusion limited (Hunten et al. 1987; Zahnle et al. 1990, 2019) or whether they are small and all species are lost equally through free advection (Kubyshkina et al. 2018; Caldiroli et al. 2021). If fractionation occurs, the upper regions of the atmosphere become preferentially enriched, imposing a diffusive flux that replenishes lost hydrogen in the enriched background gas. In other words, preferential hydrogen loss requires preferential restocking, which cannot occur through bulk advection alone. Mass transport in the upper atmosphere must therefore involve diffusion, lowering mass loss rates by several orders of magnitude.

In the following, we evaluate the effectiveness of fractionation and the activation of diffusion-limited mass loss in planets with primordial atmospheres. The framework presented in this paper is derived from first principles and it applies to all planets with ideal gas atmospheres, though it is most relevant to those with hydrogen-rich primordial envelopes. We begin with reviewing how the diffusion-limited mass loss rate is defined (section 2.1) and derive a new model for the crossover mass (section 2.2). In section 2.3, we apply these models to the upper atmosphere of Earth. The upper atmosphere of Earth, being hydrogen-rich, is an appropriate analog for exoplanets with hydrogen-rich atmospheres, for which compositional data is limited and uncertain. In section 3.1, we demonstrate that compositional fractionation almost always occurs as a result of X-ray and ultraviolet induced photoevaporation. Advection is shown to be more efficient than diffusion, so a hydrogen depleted layer forms at the top of the atmosphere, which grows rapidly until reaching the bottom of the atmosphere. After equilibrium is reached, mass loss becomes diffusion limited. In section 3.2, we present and discuss representative simulations of an exoplanet undergoing photoevaporation and its dependence on the chosen mass loss model. We conclude with an overview of how our findings compare to other approaches in the literature.

2 Model

2.1 Defining the diffusion-limited flux

Diffusion-limited flux is the maximum rate at which a species can traverse through an atmosphere when advection and diffusion are considered concurrently. Despite its name, diffusion-limited mass loss does not imply a lack of advection. Instead, it describes how momentum diffusion between species acts against bulk flow in an advecting binary gas mixture. At the microscopic scale, the fast moving light species collide with slower moving heavy species, resulting in the light species decelerating and the heavy species accelerating. At the macroscopic scale, it appears as if the heavy species exert drag on the lighter species while the lighter species pull on the heavy species. The diffusion-limited mass loss model extends the Euler equations by incorporating interactions between different species (Chapman & Cowling 1970, p. 107). In the traditional approach of Hunten (1973), the diffusion-limited flux is estimated by evaluating the relative average velocity of a binary gas mixture in a gravitational field (Chapman & Cowling 1970),

v1v2=D[n2n1n2(n1n)+μ2μ1μ¯(lnP)+αT(lnT)μ1μ2μ¯kBT(a1a2)],subscript𝑣1subscript𝑣2𝐷delimited-[]superscript𝑛2subscript𝑛1subscript𝑛2subscript𝑛1𝑛subscript𝜇2subscript𝜇1¯𝜇𝑃subscript𝛼𝑇𝑇subscript𝜇1subscript𝜇2¯𝜇subscript𝑘B𝑇subscript𝑎1subscript𝑎2\vec{v}_{1}-\vec{v}_{2}=-D\left[\frac{n^{2}}{n_{1}n_{2}}\nabla\left(\frac{n_{1% }}{n}\right)+\frac{\mu_{2}-\mu_{1}}{\bar{\mu}}\nabla\left(\ln{P}\right)+\alpha% _{T}\nabla\left(\ln{T}\right)-\frac{\mu_{1}\mu_{2}}{\bar{\mu}k_{\rm B}T}\left(% \vec{a}_{1}-\vec{a}_{2}\right)\right],over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - over→ start_ARG italic_v end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = - italic_D [ divide start_ARG italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG ∇ ( divide start_ARG italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_n end_ARG ) + divide start_ARG italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_μ end_ARG end_ARG ∇ ( roman_ln italic_P ) + italic_α start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT ∇ ( roman_ln italic_T ) - divide start_ARG italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_μ end_ARG italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T end_ARG ( over→ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - over→ start_ARG italic_a end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ] , (1)

where subscripts 1 and 2 are for the light and heavy constituents, respectively, D𝐷Ditalic_D is the diffusion coefficient, n𝑛nitalic_n is the particle number density, μ¯¯𝜇\bar{\mu}over¯ start_ARG italic_μ end_ARG is the mean molecular mass, P𝑃Pitalic_P is the pressure, αTsubscript𝛼𝑇\alpha_{T}italic_α start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT is the thermal diffusion factor (not to be confused with the coefficient of thermal diffusivity; Leuenberger & Lang 2002), T𝑇Titalic_T is the temperature, kBsubscript𝑘Bk_{\rm B}italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT is the Boltzmann constant, and a𝑎aitalic_a is the acceleration acting on the particles from external forces. The first term is the concentration gradient, the second is the mass gradient, the third is the temperature gradient, and the fourth is the external force gradient. Equation 1 is evaluated for Earth at the homopause, which is very close to the mesopause, where there is a temperature inversion and (lnT)=0𝑇0\nabla\left(\ln{T}\right){=}0∇ ( roman_ln italic_T ) = 0. The homopause is the level below which an atmosphere is well-mixed whereas the mesopause is the boundary between the mesosphere and the thermosphere; the homopause and the mesopause are both located at approximately 85km85km85~{}{\rm km}85 roman_km, though the homopause is known to vary in altitude from 80120km80120km80{-}120~{}{\rm km}80 - 120 roman_km (Salinas et al. 2016; Liu 2021; Swenson et al. 2021).

The planetary magnetic field is small, so the external acceleration for both particles is set by gravity, which is g=GMp/r2𝑔𝐺subscript𝑀psuperscript𝑟2g{=}GM_{\rm p}/r^{2}italic_g = italic_G italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT / italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (G𝐺Gitalic_G is the gravitational constant, Mpsubscript𝑀pM_{\rm p}italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT is the planetary mass, and r𝑟ritalic_r is the radial distance) and therefore cancels out. In its current form, Equation 1 only includes molecular diffusion, which involves the random movement of individual molecules from areas of high concentration to low concentration. In contrast, eddy diffusion involves mixing by turbulence and acts on the compositional gradient term (Catling & Kasting 2017), thus it has to be included in its numerator. Introducing the mole fraction χ𝜒\chiitalic_χ so that n1=χnsubscript𝑛1𝜒𝑛n_{1}{=}\chi nitalic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_χ italic_n and n2=(1χ)nsubscript𝑛21𝜒𝑛n_{2}{=}\left(1{-}\chi\right)nitalic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = ( 1 - italic_χ ) italic_n, and considering only the radial profile, we thus have

v1v2=D[1+KzzDχ(1χ)dχdr+μ2μ1μ¯1PdPdr].subscript𝑣1subscript𝑣2𝐷delimited-[]1subscript𝐾zz𝐷𝜒1𝜒d𝜒d𝑟subscript𝜇2subscript𝜇1¯𝜇1𝑃d𝑃d𝑟v_{1}-v_{2}=-D\left[\frac{1+\frac{K_{\rm zz}}{D}}{\chi\left(1-\chi\right)}% \frac{{\rm d}\chi}{{\rm d}r}+\frac{\mu_{2}-\mu_{1}}{\bar{\mu}}\frac{1}{P}\frac% {{\rm d}P}{{\rm d}r}\right].italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = - italic_D [ divide start_ARG 1 + divide start_ARG italic_K start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT end_ARG start_ARG italic_D end_ARG end_ARG start_ARG italic_χ ( 1 - italic_χ ) end_ARG divide start_ARG roman_d italic_χ end_ARG start_ARG roman_d italic_r end_ARG + divide start_ARG italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_μ end_ARG end_ARG divide start_ARG 1 end_ARG start_ARG italic_P end_ARG divide start_ARG roman_d italic_P end_ARG start_ARG roman_d italic_r end_ARG ] . (2)

According to Hunten (1973), the compositional gradient term is negligible. Moreover, gas is assumed to be in hydrostatic equilibrium at the homopause, so 1/P(dP/dr)1𝑃d𝑃d𝑟1/P\left({\rm d}P/{\rm d}r\right)1 / italic_P ( roman_d italic_P / roman_d italic_r ) is equal to 1/H1𝐻{-}1/H- 1 / italic_H, where H𝐻Hitalic_H is the scale height defined as H=kBT/(gμ¯)𝐻subscript𝑘B𝑇𝑔¯𝜇H{=}k_{\rm B}T/\left(g\bar{\mu}\right)italic_H = italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T / ( italic_g over¯ start_ARG italic_μ end_ARG ),

v1v2=DgkBT(μ2μ1).subscript𝑣1subscript𝑣2𝐷𝑔subscript𝑘B𝑇subscript𝜇2subscript𝜇1v_{1}-v_{2}=\frac{Dg}{k_{\rm B}T}\left(\mu_{2}-\mu_{1}\right).italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = divide start_ARG italic_D italic_g end_ARG start_ARG italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T end_ARG ( italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) . (3)

In the limit when μ1μ2much-less-thansubscript𝜇1subscript𝜇2\mu_{1}{\ll}\mu_{2}italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≪ italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, Graham’s law of diffusion suggests that v1v2much-greater-thansubscript𝑣1subscript𝑣2v_{1}{\gg}v_{2}italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≫ italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, and v1v2v1subscript𝑣1subscript𝑣2subscript𝑣1v_{1}{-}v_{2}{\approx}v_{1}italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≈ italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. Instead of employing Chapman-Enskog theory (Chapman & Cowling 1970) for calculating the molecular diffusion coefficient, Hunten (1973) substitutes D𝐷Ditalic_D with the binary diffusion coefficient, bij=nDsubscript𝑏ij𝑛𝐷b_{\rm ij}{=}nDitalic_b start_POSTSUBSCRIPT roman_ij end_POSTSUBSCRIPT = italic_n italic_D, which is determined experimentally (Marrero & Mason 1972) and takes the form bij=ATBsubscript𝑏ij𝐴superscript𝑇𝐵b_{\rm ij}{=}AT^{B}italic_b start_POSTSUBSCRIPT roman_ij end_POSTSUBSCRIPT = italic_A italic_T start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT, with A𝐴Aitalic_A and B𝐵Bitalic_B being constants. Multiplying both sides of Equation 3 by the number density of the light species, n1subscript𝑛1n_{1}italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, yields the diffusion-limited flux,

ϕ1=n1v1=χbijgkBT(μ2μ1),subscriptitalic-ϕ1subscript𝑛1subscript𝑣1𝜒subscript𝑏ij𝑔subscript𝑘B𝑇subscript𝜇2subscript𝜇1\phi_{1}=n_{1}v_{1}=\frac{\chi b_{\rm ij}g}{k_{\rm B}T}\left(\mu_{2}-\mu_{1}% \right),italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = divide start_ARG italic_χ italic_b start_POSTSUBSCRIPT roman_ij end_POSTSUBSCRIPT italic_g end_ARG start_ARG italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T end_ARG ( italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) , (4)

where ϕ1subscriptitalic-ϕ1\phi_{1}italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is the escaping particle flux. Mass conservation ensures equal mass flow through all atmospheric spherical shells, thus the global mass loss rate is set by mass flow at the homopause (subscript H),

M˙1=χH4πGMpμ1bij,HkBTH(μ2μ1).subscript˙𝑀1subscript𝜒H4𝜋𝐺subscript𝑀psubscript𝜇1subscript𝑏ijHsubscript𝑘Bsubscript𝑇Hsubscript𝜇2subscript𝜇1\dot{M}_{1}=\chi_{\rm H}\frac{4\pi GM_{\rm p}\mu_{1}b_{\rm ij,H}}{k_{\rm B}T_{% \rm H}}\left(\mu_{2}-\mu_{1}\right).over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_χ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT divide start_ARG 4 italic_π italic_G italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT roman_ij , roman_H end_POSTSUBSCRIPT end_ARG start_ARG italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG ( italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) . (5)

Equation 5 assumes a binary gas mixture, with a light major component (atomic hydrogen) and a heavy minor component. Atmospheres are, however, composed of various species, and it is therefore necessary to define the mean molecular mass of the average effective heavy component,

μ2=μ¯χμ11χ,subscript𝜇2¯𝜇𝜒subscript𝜇11𝜒\mu_{2}=\frac{\bar{\mu}-\chi\mu_{1}}{1-\chi},italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = divide start_ARG over¯ start_ARG italic_μ end_ARG - italic_χ italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 1 - italic_χ end_ARG , (6)

where μ¯¯𝜇\bar{\mu}over¯ start_ARG italic_μ end_ARG is the local mean molecular mass. Collectively, one gets,

M˙1=χH1χH4πGMpμ1bij,HkBTH(μ¯μ1).subscript˙𝑀1subscript𝜒H1subscript𝜒H4𝜋𝐺subscript𝑀psubscript𝜇1subscript𝑏ijHsubscript𝑘Bsubscript𝑇H¯𝜇subscript𝜇1\dot{M}_{1}=\frac{\chi_{\rm H}}{1{-}\chi_{\rm H}}\frac{4\pi GM_{\rm p}\mu_{1}b% _{\rm ij,H}}{k_{\rm B}T_{\rm H}}\left(\bar{\mu}-\mu_{1}\right).over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = divide start_ARG italic_χ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG start_ARG 1 - italic_χ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG divide start_ARG 4 italic_π italic_G italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT roman_ij , roman_H end_POSTSUBSCRIPT end_ARG start_ARG italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG ( over¯ start_ARG italic_μ end_ARG - italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) . (7)

Equations 5 and 7 apply to planets that have achieved a diffusion-limited steady state (section 3.1), such as Earth (Joshi et al. 2019). These equations are independent of X-ray and ultraviolet irradiation because they describe the maximum rate at which hydrogen can be transported within the atmosphere, thus setting the limit for how much hydrogen can be lost at the top of the atmosphere. As mass loss is evaluated at the homopause, and as the homopause is neutral, the effects of photochemistry are unimportant.

2.2 Defining the crossover mass

The crossover mass is defined as the threshold mass above which particles are too heavy to be dragged along with other escaping species. Fractionation occurs when the crossover mass is low enough for a gas mixture to undergo differential separation. Hunten et al. (1987) finds the crossover mass by solving the diffusion-limited flux for μcrsubscript𝜇cr\mu_{\rm cr}italic_μ start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT (=μ2absentsubscript𝜇2{=}\mu_{2}= italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT in Equations 4 and 5),

μcr=μ1+kBTϕ1χbijg,subscript𝜇crsubscript𝜇1subscript𝑘B𝑇subscriptitalic-ϕ1𝜒subscript𝑏ij𝑔\mu_{\rm cr}=\mu_{1}+\frac{k_{\rm B}T\phi_{1}}{\chi b_{\rm ij}g},italic_μ start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT = italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + divide start_ARG italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_χ italic_b start_POSTSUBSCRIPT roman_ij end_POSTSUBSCRIPT italic_g end_ARG , (8)

or, equivalently,

μcr=μ1+kBTM˙14πχbijGMpμ1.subscript𝜇crsubscript𝜇1subscript𝑘B𝑇subscript˙𝑀14𝜋𝜒subscript𝑏ij𝐺subscript𝑀psubscript𝜇1\mu_{\rm cr}=\mu_{1}+\frac{k_{\rm B}T\dot{M}_{1}}{4\pi\chi b_{\rm ij}GM_{\rm p% }\mu_{1}}.italic_μ start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT = italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + divide start_ARG italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_π italic_χ italic_b start_POSTSUBSCRIPT roman_ij end_POSTSUBSCRIPT italic_G italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG . (9)

Equations 8 and 9 derive from Equation 1, which assumes that both gas species experience the same acceleration as the bulk fluid, that is, dv1/dt=dv2/dt=dv/dtdsubscript𝑣1d𝑡dsubscript𝑣2d𝑡d𝑣d𝑡{\rm d}v_{1}/{\rm d}t{=}{\rm d}v_{2}/{\rm d}t{=}{\rm d}v/{\rm d}troman_d italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / roman_d italic_t = roman_d italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_d italic_t = roman_d italic_v / roman_d italic_t (Chapman & Cowling 1970, p. 107). However, this assumption does not apply when evaluating fractionation because, under such circumstances, dv2/dt=0dsubscript𝑣2d𝑡0{\rm d}v_{2}/{\rm d}t{=}0roman_d italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_d italic_t = 0. To address this, we use Equation 6.62, 9 of Chapman & Cowling (1970) instead, with dv2/dtdsubscript𝑣2d𝑡{\rm d}v_{2}/{\rm d}troman_d italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_d italic_t in place of dv/dtd𝑣d𝑡{\rm d}v/{\rm d}troman_d italic_v / roman_d italic_t:

ρ2dv2dt=ρ2gdP2dr+ρ1ρ2ρτ1,2(v1v2),subscript𝜌2dsubscript𝑣2d𝑡subscript𝜌2𝑔dsubscript𝑃2d𝑟subscript𝜌1subscript𝜌2𝜌subscript𝜏12subscript𝑣1subscript𝑣2\rho_{2}\frac{{\rm d}v_{2}}{{\rm d}t}=-\rho_{2}g-\frac{{\rm d}P_{2}}{{\rm d}r}% +\frac{\rho_{1}\rho_{2}}{\rho\tau_{1,2}}\left(v_{1}-v_{2}\right),italic_ρ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT divide start_ARG roman_d italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_t end_ARG = - italic_ρ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_g - divide start_ARG roman_d italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_r end_ARG + divide start_ARG italic_ρ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ italic_τ start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT end_ARG ( italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) , (10)

where τ1,2subscript𝜏12\tau_{1,2}italic_τ start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT is the time between particle collisions. From Equations 6.62, 6 and 6.62, 7 of Chapman & Cowling (1970),

τ1,2=ρ1ρ2PP1P2ρD,subscript𝜏12subscript𝜌1subscript𝜌2𝑃subscript𝑃1subscript𝑃2𝜌𝐷\tau_{1,2}=\frac{\rho_{1}\rho_{2}P}{P_{1}P_{2}\rho}D,italic_τ start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT = divide start_ARG italic_ρ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ρ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_P end_ARG start_ARG italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_ρ end_ARG italic_D , (11)

which upon inserting into Equation 10 and simplifying with the ideal gas equation,

dv2dt=gc22P2dP2dr+χc22D(v1v2),dsubscript𝑣2d𝑡𝑔superscriptsubscript𝑐22subscript𝑃2dsubscript𝑃2d𝑟𝜒superscriptsubscript𝑐22𝐷subscript𝑣1subscript𝑣2\frac{{\rm d}v_{2}}{{\rm d}t}=-g-\frac{c_{2}^{2}}{P_{2}}\frac{{\rm d}P_{2}}{{% \rm d}r}+\frac{\chi c_{2}^{2}}{D}\left(v_{1}-v_{2}\right),divide start_ARG roman_d italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_t end_ARG = - italic_g - divide start_ARG italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG divide start_ARG roman_d italic_P start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_r end_ARG + divide start_ARG italic_χ italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_D end_ARG ( italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) , (12)

where c2=kBT/μ2subscript𝑐2subscript𝑘B𝑇subscript𝜇2c_{2}{=}\sqrt{k_{\rm B}T/\mu_{2}}italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = square-root start_ARG italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T / italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG is the isothermal sound speed of the heavy species. Applying chain rule and dividing through by c22superscriptsubscript𝑐22c_{2}^{2}italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT,

1c22dv2dt=GMpμ2kBTr2+11χdχdr1PdPdr+χD(v1v2).1superscriptsubscript𝑐22dsubscript𝑣2d𝑡𝐺subscript𝑀psubscript𝜇2subscript𝑘B𝑇superscript𝑟211𝜒d𝜒d𝑟1𝑃d𝑃d𝑟𝜒𝐷subscript𝑣1subscript𝑣2\frac{1}{c_{2}^{2}}\frac{{\rm d}v_{2}}{{\rm d}t}=-\frac{GM_{\rm p}\mu_{2}}{k_{% \rm B}Tr^{2}}+\frac{1}{1-\chi}\frac{{\rm d}\chi}{{\rm d}r}-\frac{1}{P}\frac{{% \rm d}P}{{\rm d}r}+\frac{\chi}{D}\left(v_{1}-v_{2}\right).divide start_ARG 1 end_ARG start_ARG italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG roman_d italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_t end_ARG = - divide start_ARG italic_G italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG 1 end_ARG start_ARG 1 - italic_χ end_ARG divide start_ARG roman_d italic_χ end_ARG start_ARG roman_d italic_r end_ARG - divide start_ARG 1 end_ARG start_ARG italic_P end_ARG divide start_ARG roman_d italic_P end_ARG start_ARG roman_d italic_r end_ARG + divide start_ARG italic_χ end_ARG start_ARG italic_D end_ARG ( italic_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) . (13)

Hunten (1973) evaluates the diffusion-limited flux at the homopause, but it is unclear if this is a suitable location for evaluating the crossover mass, which depends on local system properties. In fact, the properties of the upper atmosphere differ substantially from those in the deeper regions, and, if we are interested in knowing the compositional evolution of the bulk atmosphere, the most relevant place would be the upper edge of the atmosphere where atmospheric particles escape to space. It is thus preferable to evaluate the crossover mass at the exobase (subscript x), which is defined as the location above which gas becomes rarefied and collisions no longer dominate particle dynamics. The exobase height varies from 5001000km5001000km500{-}1000~{}{\rm km}500 - 1000 roman_km depending on solar activity (Emmert 2015); beyond the exobase, a lower crossover mass is less significant because the reduced particle density gives rise to greater statistical noise. This statistical noise invalidates the continuity assumption because individual particle motions are too significant for gas to be treated as a coherent flow (Oran et al. 1998; Shematovich et al. 2015). Evaluating Equation 13 at the exobase where hydrostatic equilibrium applies (Modirrousta-Galian & Korenaga submitted) reduces the (1/P)dP/dr1𝑃d𝑃d𝑟(1/P){\rm d}P/{\rm d}r( 1 / italic_P ) roman_d italic_P / roman_d italic_r term to the negative inverse of the scale height of the bulk fluid,

1c22dv2dt=GMp(μ¯μ2)kBTxRx2+11χx(dχdr)x+χxDx(v1,xv2,x).1superscriptsubscript𝑐22dsubscript𝑣2d𝑡𝐺subscript𝑀p¯𝜇subscript𝜇2subscript𝑘Bsubscript𝑇xsuperscriptsubscript𝑅x211subscript𝜒xsubscriptd𝜒d𝑟xsubscript𝜒xsubscript𝐷xsubscript𝑣1xsubscript𝑣2x\frac{1}{c_{2}^{2}}\frac{{\rm d}v_{2}}{{\rm d}t}=\frac{GM_{\rm p}\left(\bar{% \mu}-\mu_{2}\right)}{k_{\rm B}T_{\rm x}R_{\rm x}^{2}}+\frac{1}{1-\chi_{\rm x}}% \left(\frac{{\rm d}\chi}{{\rm d}r}\right)_{\rm x}+\frac{\chi_{\rm x}}{D_{\rm x% }}\left(v_{\rm 1,x}-v_{\rm 2,x}\right).divide start_ARG 1 end_ARG start_ARG italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG roman_d italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_t end_ARG = divide start_ARG italic_G italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT ( over¯ start_ARG italic_μ end_ARG - italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_ARG start_ARG italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG 1 end_ARG start_ARG 1 - italic_χ start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT end_ARG ( divide start_ARG roman_d italic_χ end_ARG start_ARG roman_d italic_r end_ARG ) start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT + divide start_ARG italic_χ start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT end_ARG start_ARG italic_D start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT end_ARG ( italic_v start_POSTSUBSCRIPT 1 , roman_x end_POSTSUBSCRIPT - italic_v start_POSTSUBSCRIPT 2 , roman_x end_POSTSUBSCRIPT ) . (14)

Fractionation occurs for species heavier than the crossover mass (μ2=μcrsubscript𝜇2subscript𝜇cr\mu_{2}{=}\mu_{\rm cr}italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_μ start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT) when v2=dv2/dt=0subscript𝑣2dsubscript𝑣2d𝑡0v_{2}{=}{\rm d}v_{2}/{\rm d}t{=}0italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = roman_d italic_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT / roman_d italic_t = 0 so that

GMp(μ¯μcr)kBTxRx2+11χx(dχdr)x+χxv1,xDx=0.𝐺subscript𝑀p¯𝜇subscript𝜇crsubscript𝑘Bsubscript𝑇xsuperscriptsubscript𝑅x211subscript𝜒xsubscriptd𝜒d𝑟xsubscript𝜒xsubscript𝑣1xsubscript𝐷x0\frac{GM_{\rm p}\left(\bar{\mu}-\mu_{\rm cr}\right)}{k_{\rm B}T_{\rm x}R_{\rm x% }^{2}}+\frac{1}{1-\chi_{\rm x}}\left(\frac{{\rm d}\chi}{{\rm d}r}\right)_{\rm x% }+\frac{\chi_{\rm x}v_{\rm 1,x}}{D_{\rm x}}=0.divide start_ARG italic_G italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT ( over¯ start_ARG italic_μ end_ARG - italic_μ start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT ) end_ARG start_ARG italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG 1 end_ARG start_ARG 1 - italic_χ start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT end_ARG ( divide start_ARG roman_d italic_χ end_ARG start_ARG roman_d italic_r end_ARG ) start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT + divide start_ARG italic_χ start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT 1 , roman_x end_POSTSUBSCRIPT end_ARG start_ARG italic_D start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT end_ARG = 0 . (15)

Multiplying the numerator and denominator of the rightmost term by n𝑛nitalic_n yields,

GMp(μ¯μcr)kBTxRx2+11χx(dχdr)x+ϕxbij,x=0,𝐺subscript𝑀p¯𝜇subscript𝜇crsubscript𝑘Bsubscript𝑇xsuperscriptsubscript𝑅x211subscript𝜒xsubscriptd𝜒d𝑟xsubscriptitalic-ϕxsubscript𝑏ijx0\frac{GM_{\rm p}\left(\bar{\mu}-\mu_{\rm cr}\right)}{k_{\rm B}T_{\rm x}R_{\rm x% }^{2}}+\frac{1}{1-\chi_{\rm x}}\left(\frac{{\rm d}\chi}{{\rm d}r}\right)_{\rm x% }+\frac{\phi_{\rm x}}{b_{\rm ij,x}}=0,divide start_ARG italic_G italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT ( over¯ start_ARG italic_μ end_ARG - italic_μ start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT ) end_ARG start_ARG italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG 1 end_ARG start_ARG 1 - italic_χ start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT end_ARG ( divide start_ARG roman_d italic_χ end_ARG start_ARG roman_d italic_r end_ARG ) start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT + divide start_ARG italic_ϕ start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT end_ARG start_ARG italic_b start_POSTSUBSCRIPT roman_ij , roman_x end_POSTSUBSCRIPT end_ARG = 0 , (16)

which can also be expressed in terms of the mass flow rate,

GMp(μ¯μcr)kBTxRx2+11χx(dχdr)x+M˙14πRx2μ1bij,x=0.𝐺subscript𝑀p¯𝜇subscript𝜇crsubscript𝑘Bsubscript𝑇xsuperscriptsubscript𝑅x211subscript𝜒xsubscriptd𝜒d𝑟xsubscript˙𝑀14𝜋superscriptsubscript𝑅x2subscript𝜇1subscript𝑏ijx0\frac{GM_{\rm p}\left(\bar{\mu}-\mu_{\rm cr}\right)}{k_{\rm B}T_{\rm x}R_{\rm x% }^{2}}+\frac{1}{1-\chi_{\rm x}}\left(\frac{{\rm d}\chi}{{\rm d}r}\right)_{\rm x% }+\frac{\dot{M}_{1}}{4\pi R_{\rm x}^{2}\mu_{1}b_{\rm ij,x}}=0.divide start_ARG italic_G italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT ( over¯ start_ARG italic_μ end_ARG - italic_μ start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT ) end_ARG start_ARG italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG 1 end_ARG start_ARG 1 - italic_χ start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT end_ARG ( divide start_ARG roman_d italic_χ end_ARG start_ARG roman_d italic_r end_ARG ) start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT + divide start_ARG over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_π italic_R start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_b start_POSTSUBSCRIPT roman_ij , roman_x end_POSTSUBSCRIPT end_ARG = 0 . (17)

The compositional term is small and can be discarded (Zahnle & Kasting 1986). Solving for μcrsubscript𝜇cr\mu_{\rm cr}italic_μ start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT,

μcr=μ¯+kBTxM˙14πbij,xGMpμ1,subscript𝜇cr¯𝜇subscript𝑘Bsubscript𝑇xsubscript˙𝑀14𝜋subscript𝑏ijx𝐺subscript𝑀psubscript𝜇1\mu_{\rm cr}=\bar{\mu}+\frac{k_{\rm B}T_{\rm x}\dot{M}_{1}}{4\pi b_{\rm ij,x}% GM_{\rm p}\mu_{1}},italic_μ start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT = over¯ start_ARG italic_μ end_ARG + divide start_ARG italic_k start_POSTSUBSCRIPT roman_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_π italic_b start_POSTSUBSCRIPT roman_ij , roman_x end_POSTSUBSCRIPT italic_G italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG , (18)

where we see that Equations 9 and 18 share a similar functional form but with different parameter values, and χ𝜒\chiitalic_χ not being in the denominator. When a diffusion-limited steady state applies (section 3.1), the hydrogen mass flow rate, M˙1subscript˙𝑀1\dot{M}_{1}over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, is set by Equation 7 and cannot be replaced by alternative models such as the energy-limited (Watson et al. 1981) or hydro-based (Kubyshkina et al. 2018) approximations (discussed in section 4). As mentioned previously, diffusion-limited mass loss combines advection and momentum diffusion. This interaction creates a negative feedback mechanism in multicomponent gas mixtures where extreme mass loss becomes unsustainable because heavy species slow down escaping hydrogen, reducing the crossover mass, and leading to increased fractionation and further decreased hydrogen outflow. It would therefore be misleading to use an extreme mass loss model that assumes no fractionation because this is the very reason why they exhibit extreme mass loss in the first place. Such models inherently preclude fractionation because they do not incorporate the drag effect imposed by heavy species on escaping hydrogen. To self-consistently quantify fractionation in a multicomponent system, we therefore use Equation 7 and combine it with bijT3/4subscript𝑏ijsuperscript𝑇34b_{\rm ij}{\mathchoice{\mathrel{\raise 1.07639pt\hbox{\hbox to 0.0pt{\hbox{$% \displaystyle\propto$}\hss}\lower 4.03563pt\hbox{$\displaystyle\sim$}}}}{% \mathrel{\raise 1.07639pt\hbox{\hbox to 0.0pt{\hbox{$\textstyle\propto$}\hss}% \lower 4.03563pt\hbox{$\textstyle\sim$}}}}{\mathrel{\raise 0.75346pt\hbox{% \hbox to 0.0pt{\hbox{$\scriptstyle\propto$}\hss}\lower 2.82494pt\hbox{$% \scriptstyle\sim$}}}}{\mathrel{\raise 0.5382pt\hbox{\hbox to 0.0pt{\hbox{$% \scriptscriptstyle\propto$}\hss}\lower 2.0178pt\hbox{$\scriptscriptstyle\sim$}% }}}}T^{3/4}italic_b start_POSTSUBSCRIPT roman_ij end_POSTSUBSCRIPT ∝∼ italic_T start_POSTSUPERSCRIPT 3 / 4 end_POSTSUPERSCRIPT (Mason & Marrero 1970; Marrero & Mason 1972),

μcr=μ¯x+χH1χH(TxTH)14(μ¯Hμ1).subscript𝜇crsubscript¯𝜇xsubscript𝜒H1subscript𝜒Hsuperscriptsubscript𝑇xsubscript𝑇H14subscript¯𝜇Hsubscript𝜇1\mu_{\rm cr}=\bar{\mu}_{\rm x}+\frac{\chi_{\rm H}}{1{-}\chi_{\rm H}}\left(% \frac{T_{\rm x}}{T_{\rm H}}\right)^{\frac{1}{4}}\left(\bar{\mu}_{\rm H}-\mu_{1% }\right).italic_μ start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT = over¯ start_ARG italic_μ end_ARG start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT + divide start_ARG italic_χ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG start_ARG 1 - italic_χ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG ( divide start_ARG italic_T start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 4 end_ARG end_POSTSUPERSCRIPT ( over¯ start_ARG italic_μ end_ARG start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT - italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) . (19)

For well-mixed gas, such as exoplanets with primordial atmospheres (i.e., KzzDmuch-greater-thansubscript𝐾zz𝐷K_{\rm zz}{\gg}Ditalic_K start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT ≫ italic_D; Modirrousta-Galian & Korenaga 2023), μ¯xμ¯Hμ¯subscript¯𝜇xsubscript¯𝜇H¯𝜇\bar{\mu}_{\rm x}{\approx}\bar{\mu}_{\rm H}{\approx}\bar{\mu}over¯ start_ARG italic_μ end_ARG start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT ≈ over¯ start_ARG italic_μ end_ARG start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT ≈ over¯ start_ARG italic_μ end_ARG and,

μcr=μ¯[1+χH1χH(TxTH)14(1μ1μ¯)].subscript𝜇cr¯𝜇delimited-[]1subscript𝜒H1subscript𝜒Hsuperscriptsubscript𝑇xsubscript𝑇H141subscript𝜇1¯𝜇\mu_{\rm cr}=\bar{\mu}\left[1+\frac{\chi_{\rm H}}{1{-}\chi_{\rm H}}\left(\frac% {T_{\rm x}}{T_{\rm H}}\right)^{\frac{1}{4}}\left(1-\frac{\mu_{1}}{\bar{\mu}}% \right)\right].italic_μ start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT = over¯ start_ARG italic_μ end_ARG [ 1 + divide start_ARG italic_χ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG start_ARG 1 - italic_χ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG ( divide start_ARG italic_T start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT end_ARG start_ARG italic_T start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 4 end_ARG end_POSTSUPERSCRIPT ( 1 - divide start_ARG italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_μ end_ARG end_ARG ) ] . (20)

Equations 19 and 20 apply to all planets with compositionally stratified and well-mixed ideal gas atmospheres, respectively, that have achieved a diffusion-limited steady state (section 3.1). They suggest that in the diffusion-limited mass loss framework, the crossover mass is independent of the hydrogen outflow rate but depends on the hydrogen mole fraction, the temperature ratio to the one-fourth power, and the atmospheric composition. The photosphere (where optical depth τ=2/3𝜏23\tau{=}2/3italic_τ = 2 / 3) is a suitable alternative to the homopause for exoplanets with hydrogen-rich atmospheres because it is also a temperature minimum (i.e., dT/dr=0d𝑇d𝑟0{\rm d}T/{\rm d}r{=}0roman_d italic_T / roman_d italic_r = 0; Modirrousta-Galian & Korenaga 2023). In the following section, we compare our approach to that of Hunten et al. (1987).

2.3 Model comparison

Our model for the crossover mass (Equations 19 and 20) and that of Hunten et al. (1987) (Equations 8 and 9) are based on different assumptions. We compare their predictions with the observations of Earth’s atmosphere, using the NRLMSIS 2.0 code (Emmert et al. 2021), which provides the average observed behavior of an atmospheric column at a given latitude, longitude, and elevation. Athens, Greece (37.9838{}^{\circ}start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPT N, 23.7275{}^{\circ}start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPT E) is chosen as the reference location because of its proximity to multiple continents and its mean annual temperature being similar to Earth’s average temperature (288Ksimilar-toabsent288K{\sim}288~{}{\rm K}∼ 288 roman_K). From the output data, we extract the mole fraction of hydrogen (H), helium (He), oxygen (O and O22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT), nitrogen (N and N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT), and argon (Ar), along with the mean molecular mass, pressure, and temperature for elevations of 01000km01000km0{-}1000~{}{\rm km}0 - 1000 roman_km. Figure 1 shows the parameter and compositional profiles adopted in this study.

Refer to caption
Figure 1: (a) Normalized temperature (solid line), mean molecular mass (dotted line), and pressure (dashed line). The temperature, mean molecular mass, and pressure are normalized with T/(923.5K)𝑇923.5KT/\left(923.5~{}{\rm K}\right)italic_T / ( 923.5 roman_K ), μ¯/28.96¯𝜇28.96\bar{\mu}/28.96over¯ start_ARG italic_μ end_ARG / 28.96, and log10(P/Pmin)/log10(Pmax/Pmin)subscript10𝑃subscript𝑃minsubscript10subscript𝑃maxsubscript𝑃min\log_{10}\left(P/P_{\rm min}\right)/\log_{10}\left(P_{\rm max}/P_{\rm min}\right)roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_P / italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ) / roman_log start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT ( italic_P start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT / italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ), respectively, where Pmax=1.002×105Pasubscript𝑃max1.002superscript105PaP_{\rm max}{=}1.002{\times}10^{5}~{}{\rm Pa}italic_P start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 1.002 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT roman_Pa and Pmin=2.969×109Pasubscript𝑃min2.969superscript109PaP_{\rm min}{=}2.969{\times}10^{-9}~{}{\rm Pa}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = 2.969 × 10 start_POSTSUPERSCRIPT - 9 end_POSTSUPERSCRIPT roman_Pa are the maximum and minimum pressures in our atmospheric grid. (b) The terrestrial compositional profile for hydrogen (H, solid orange line), helium (He, solid green line), oxygen (blue dashed line for O and blue solid line for O22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT), nitrogen (red dashed line for N and red solid line for N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT), and argon (Ar, solid purple line). Data from the NRLMSIS 2.0 code (Emmert et al. 2021).

Equations 8 and 19 are applied to the exobase (5001000km5001000km500{-}1000~{}{\rm km}500 - 1000 roman_km), with the former using the binary diffusion coefficients of Hunten (1973), Hunten et al. (1987), and Catling & Kasting (2017). We obtain the required particle flux, ϕ1subscriptitalic-ϕ1\phi_{1}italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, using the parameterization ϕ1=1.2×1012(r/6837km)2m2s1subscriptitalic-ϕ11.2superscript1012superscript𝑟6837km2superscriptm2superscripts1\phi_{1}{=}1.2{\times}10^{12}\left(r/6837~{}{\rm km}\right)^{-2}~{}{\rm m^{-2}% ~{}s^{-1}}italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 1.2 × 10 start_POSTSUPERSCRIPT 12 end_POSTSUPERSCRIPT ( italic_r / 6837 roman_km ) start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_m start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, which derives from observations of Earth’s thermosphere (Joshi et al. 2019). Figure 2 shows that the crossover mass of Equation 8 remains relatively constant at around 1amu1amu1~{}{\rm amu}1 roman_amu, with little variation across different binary diffusion coefficient values. In contrast, our crossover mass formulation (Equation 19) yields a value of 9±6amuplus-or-minus96amu9{\pm}6~{}{\rm amu}9 ± 6 roman_amu, and varies with altitude because of the changing mean molecular mass across the heterosphere, suggesting that fractionation occurs primarily at higher altitudes.

The crossover mass value we obtain when using the approach of Hunten et al. (1987), μcr1amusubscript𝜇cr1amu\mu_{\rm cr}{\approx}1~{}{\rm amu}italic_μ start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT ≈ 1 roman_amu, differs from the value given in their study, μcr=2.5amusubscript𝜇cr2.5amu\mu_{\rm cr}{=}2.5~{}{\rm amu}italic_μ start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT = 2.5 roman_amu, and in other iterations of their work (e.g., μcr=2.25amusubscript𝜇cr2.25amu\mu_{\rm cr}{=}2.25~{}{\rm amu}italic_μ start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT = 2.25 roman_amu; Catling & Kasting 2017, Table 5.3). This discrepancy can be understood through differences in the assumptions and parameter values adopted when evaluating the crossover mass. First, these studies derive the crossover mass from Equation 1, which assumes that all gas species have equal acceleration (Chapman & Cowling 1970, p. 107). This is physically inconsistent because fractionation can occur only when different species accelerate at different rates and, thus, Equation 7 should be used instead. Second, the energy-limited approximation (Watson et al. 1981) is used to estimate hydrogen flux whereas we use present-day observations of leaking hydrogen (Joshi et al. 2019). The use of the energy-limited approximation is inadequate because it assumes free advection, making it inherently incapable of evaluating fractionation. Fractionation involves differential separation resulting from momentum diffusion exchange between different species, which is not incorporated in the energy-limited approximation. Third, they assume a hydrogen mole fraction close to one (χ1𝜒1\chi{\approx}1italic_χ ≈ 1), which is valid only for hydrogen-rich atmospheres and not for the present-day Earth (Figure 1).

Refer to caption
Figure 2: The crossover mass at various altitudes at the exobase. The black line is our suggested crossover mass function (Equation 19) while the colored lines correspond to Equation 8 with the following binary diffusion coefficients: bij=2.67×1019T0.75m1s1subscript𝑏ij2.67superscript1019superscript𝑇0.75superscriptm1superscripts1b_{\rm ij}{=}2.67{\times}10^{19}T^{0.75}~{}{\rm m^{-1}~{}s^{-1}}italic_b start_POSTSUBSCRIPT roman_ij end_POSTSUBSCRIPT = 2.67 × 10 start_POSTSUPERSCRIPT 19 end_POSTSUPERSCRIPT italic_T start_POSTSUPERSCRIPT 0.75 end_POSTSUPERSCRIPT roman_m start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (blue solid line; Hunten 1973), bij=2.2×1021m1s1subscript𝑏ij2.2superscript1021superscriptm1superscripts1b_{\rm ij}{=}2.2{\times}10^{21}~{}{\rm m^{-1}~{}s^{-1}}italic_b start_POSTSUBSCRIPT roman_ij end_POSTSUBSCRIPT = 2.2 × 10 start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT roman_m start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (blue dashed line; Hunten et al. 1987), and bij=1.8×1021m1s1subscript𝑏ij1.8superscript1021superscriptm1superscripts1b_{\rm ij}{=}1.8{\times}10^{21}~{}{\rm m^{-1}~{}s^{-1}}italic_b start_POSTSUBSCRIPT roman_ij end_POSTSUBSCRIPT = 1.8 × 10 start_POSTSUPERSCRIPT 21 end_POSTSUPERSCRIPT roman_m start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (orange line; Catling & Kasting 2017, p. 146).

In the context of exoplanets with primordial atmospheres, Equation 20 can be used because the hydrogen mole fraction is high across the entire atmosphere (χ0.9𝜒0.9\chi{\approx}0.9italic_χ ≈ 0.9) and mixing from eddy diffusion is efficient throughout (Parmentier et al. 2013; Charnay et al. 2015). If compositional stratification occurs, Equation 19 should be employed at the top of the atmosphere (i.e., the exobase) instead. Equation 19 is independent of the chemical gradient in the deeper sections of the atmosphere because the crossover mass at the exobase sets the fractionation bottleneck for the entire atmosphere. Figure 3 shows that the crossover mass increases with χ𝜒\chiitalic_χ, suggesting that the preferential loss of hydrogen leads to a progressively more stable (i.e., more fractionating) and hydrostatic atmosphere composed of heavier species.

Refer to caption
Figure 3: The crossover mass (Equation 20) as a function of the mean molecular mass and the hydrogen mole fraction for Tx/TH=10subscript𝑇xsubscript𝑇H10T_{\rm x}/T_{\rm H}{=}10italic_T start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT / italic_T start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT = 10. The contours show the regions at which various elements are fractionated.

3 Application to exoplanets

3.1 Rapid onset of steady state

In the previous section we discussed how planets with hydrogen-rich upper-atmospheres are prone to compositional fractionation because the low crossover mass prevents heavy species from escaping, leading to local hydrogen depletion. However, information about this depletion is not felt instantaneously across the atmosphere because only the region immediately below this depleted layer can feel the effect of the compositional gradient. It is for this reason that the diffusion limit applies only when hydrogen flow is uniform across all atmospheric spherical shells, and only through achieving such a steady state does mass flow become limited by the maximum rate at which it can be transported within the atmosphere.

In this section, we explore and apply this concept to exoplanets with primordial envelopes experiencing photoevaporation. For mass loss to continue, there must be a preferential transport of hydrogen from deeper layers; otherwise the mole fraction of hydrogen, χ𝜒\chiitalic_χ, approaches zero, and mass loss ceases. This transport can occur either through advection or diffusion. If the mechanism is advection, the upper atmosphere will become exceedingly more enriched as more hydrogen is lost and more heavier species are left behind. In other words, the combination of preferential hydrogen removal at the top of the atmosphere and advection with the average composition would lead to further enrichment, preventing a steady state. Therefore, if advection were the restocking mechanism, the concentration gradient would keep growing until diffusive transport restores equilibrium. Consequently, a steady state can be reached only by balancing the preferential loss of hydrogen with the preferential restocking of hydrogen, which can occur only through diffusion. The balance between diffusion and advection determines whether chemical inhomogeneities form. If diffusion is efficient, these inhomogeneities will disperse rapidly. To evaluate whether this scenario is possible, we consider a hydrogen-rich hydrodynamic atmosphere that has just experienced preferential hydrogen loss. We follow the framework of Modirrousta-Galian & Korenaga (submitted), where it was shown that the exobase and quasi-sonic point coincide for hydrodynamic atmospheres, and we therefore refer to the highest point in the atmosphere as the quasi-sonic point (subscript QS) because we assume the atmosphere is initially hydrodynamic. This point represents the velocity maximum in the atmosphere, which, in the limit of v=cs𝑣subscript𝑐sv{=}c_{\rm s}italic_v = italic_c start_POSTSUBSCRIPT roman_s end_POSTSUBSCRIPT, corresponds to the sonic point. Figure 4 is a schematic diagram of the configuration being examined.

After experiencing mass loss, diffusion will act on the locally formed compositional gradient within a timescale of,

tdf=δDL22(Kzz+D),subscript𝑡dfsuperscriptsubscript𝛿DL22subscript𝐾zz𝐷t_{\rm df}=\frac{\delta_{\rm DL}^{2}}{2\left(K_{\rm zz}{+}D\right)},italic_t start_POSTSUBSCRIPT roman_df end_POSTSUBSCRIPT = divide start_ARG italic_δ start_POSTSUBSCRIPT roman_DL end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 ( italic_K start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT + italic_D ) end_ARG , (21)

where δDLsubscript𝛿DL\delta_{\rm DL}italic_δ start_POSTSUBSCRIPT roman_DL end_POSTSUBSCRIPT is the depleted layer depth. Assuming the atmosphere begins in a highly hydrodynamic state, eddy diffusion greatly exceeds molecular diffusion (i.e., KzzDmuch-greater-thansubscript𝐾zz𝐷K_{\rm zz}{\gg}Ditalic_K start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT ≫ italic_D). The eddy diffusion coefficient is defined as Kzz=vLmsubscript𝐾zzdelimited-⟨⟩𝑣subscript𝐿mK_{\rm zz}{=}\langle vL_{\rm m}\rangleitalic_K start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT = ⟨ italic_v italic_L start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT ⟩ where v=vQS𝑣subscript𝑣QSv{=}v_{\rm QS}italic_v = italic_v start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT is the bulk vertical wind velocity at the quasi-sonic point and Lmsubscript𝐿mL_{\rm m}italic_L start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT is the mixing length. The mixing length is not well understood because it is subject to highly nonlinear dynamical processes, such as gravitational wave breaking (e.g., Lindzen 1971, 1981). Experiments suggest that LmCvKLcsimilar-tosubscript𝐿msubscript𝐶vKsubscript𝐿cL_{\rm m}{\sim}C_{\rm vK}L_{\rm c}italic_L start_POSTSUBSCRIPT roman_m end_POSTSUBSCRIPT ∼ italic_C start_POSTSUBSCRIPT roman_vK end_POSTSUBSCRIPT italic_L start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT, where CvK0.4subscript𝐶vK0.4C_{\rm vK}{\approx}0.4italic_C start_POSTSUBSCRIPT roman_vK end_POSTSUBSCRIPT ≈ 0.4 is the von Kármán constant (Pope 2000) and Lcsubscript𝐿cL_{\rm c}italic_L start_POSTSUBSCRIPT roman_c end_POSTSUBSCRIPT is the characteristic length, which is sensitive to local system properties. In this context, the characteristic length is the depleted layer size, obtaining KzzCvKvQSδDLsimilar-tosubscript𝐾zzdelimited-⟨⟩subscript𝐶vKsubscript𝑣QSsubscript𝛿DLK_{\rm zz}{\sim}\langle C_{\rm vK}v_{\rm QS}\delta_{\rm DL}\rangleitalic_K start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT ∼ ⟨ italic_C start_POSTSUBSCRIPT roman_vK end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT roman_DL end_POSTSUBSCRIPT ⟩. Evaluating the above

tdf>5δDL4vQS=54ta,subscript𝑡df5subscript𝛿DL4subscript𝑣QS54subscript𝑡at_{\rm df}>\frac{5\delta_{\rm DL}}{4v_{\rm QS}}=\frac{5}{4}t_{\rm a},italic_t start_POSTSUBSCRIPT roman_df end_POSTSUBSCRIPT > divide start_ARG 5 italic_δ start_POSTSUBSCRIPT roman_DL end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_v start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT end_ARG = divide start_ARG 5 end_ARG start_ARG 4 end_ARG italic_t start_POSTSUBSCRIPT roman_a end_POSTSUBSCRIPT , (22)

where tasubscript𝑡at_{\rm a}italic_t start_POSTSUBSCRIPT roman_a end_POSTSUBSCRIPT is the advection timescale, demonstrating that diffusion is less efficient than advection. Under such circumstances, a locally depleted layer must form because advection removes hydrogen faster than it is restocked. This will cause the hydrogen mole fraction at the quasi-sonic point to decrease, creating a depleted layer through which hydrogen diffuses. As mass loss continues, the hydrogen content in the region immediately below the depleted layer diffuses through it and is subsequently lost at the quasi-sonic point. The size of the depleted layer will therefore increase with further mass loss (Figure 4).

Refer to caption
Figure 4: Cartoon showing the assumed structure of the upper atmosphere after developing a depleted layer. Continuity requires the hydrogen flux at the quasi-sonic point, ϕQSsubscriptitalic-ϕQS\phi_{\rm QS}italic_ϕ start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT, to equal that through the depleted layer, ϕDLsubscriptitalic-ϕDL\phi_{\rm DL}italic_ϕ start_POSTSUBSCRIPT roman_DL end_POSTSUBSCRIPT, and the growing compositional boundary ϕCBsubscriptitalic-ϕCB\phi_{\rm CB}italic_ϕ start_POSTSUBSCRIPT roman_CB end_POSTSUBSCRIPT. Diagram not to scale.

This gives rise to three equations for mass loss: one for advection at the quasi-sonic point,

ϕQS=χQSnQSvQS,subscriptitalic-ϕQSsubscript𝜒QSsubscript𝑛QSsubscript𝑣QS\phi_{\rm QS}=\chi_{\rm QS}n_{\rm QS}v_{\rm QS},italic_ϕ start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT = italic_χ start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT italic_v start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT , (23)

one for hydrogen diffusion through the depleted layer,

ϕDL=(K¯zz+D¯)χQSnQSχCBnCBδDL,subscriptitalic-ϕDLsubscript¯𝐾zz¯𝐷subscript𝜒QSsubscript𝑛QSsubscript𝜒CBsubscript𝑛CBsubscript𝛿DL\phi_{\rm DL}=-\left(\bar{K}_{\rm zz}+\bar{D}\right)\frac{\chi_{\rm QS}n_{\rm QS% }-\chi_{\rm CB}n_{\rm CB}}{\delta_{\rm DL}},italic_ϕ start_POSTSUBSCRIPT roman_DL end_POSTSUBSCRIPT = - ( over¯ start_ARG italic_K end_ARG start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT + over¯ start_ARG italic_D end_ARG ) divide start_ARG italic_χ start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT - italic_χ start_POSTSUBSCRIPT roman_CB end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_CB end_POSTSUBSCRIPT end_ARG start_ARG italic_δ start_POSTSUBSCRIPT roman_DL end_POSTSUBSCRIPT end_ARG , (24)

and the other for the growth rate of the depleted layer,

ϕCB=χCBnCBdδDLdt.subscriptitalic-ϕCBsubscript𝜒CBsubscript𝑛CBdsubscript𝛿DLd𝑡\phi_{\rm CB}=\chi_{\rm CB}n_{\rm CB}\frac{{\rm d}\delta_{\rm DL}}{{\rm d}t}.italic_ϕ start_POSTSUBSCRIPT roman_CB end_POSTSUBSCRIPT = italic_χ start_POSTSUBSCRIPT roman_CB end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_CB end_POSTSUBSCRIPT divide start_ARG roman_d italic_δ start_POSTSUBSCRIPT roman_DL end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_t end_ARG . (25)

The parameters χCBsubscript𝜒CB\chi_{\rm CB}italic_χ start_POSTSUBSCRIPT roman_CB end_POSTSUBSCRIPT and nCBsubscript𝑛CBn_{\rm CB}italic_n start_POSTSUBSCRIPT roman_CB end_POSTSUBSCRIPT are the mole fraction of hydrogen and the volumetric total particle number density at the bottom of the depleted layer (labeled the compositional boundary), and K¯zzsubscript¯𝐾zz\bar{K}_{\rm zz}over¯ start_ARG italic_K end_ARG start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT and D¯¯𝐷\bar{D}over¯ start_ARG italic_D end_ARG are the average eddy and molecular diffusion coefficients,

K¯zz+D¯=1RQSRCB(t)RCB(t)RQSKzz(r)+D(r)dr.subscript¯𝐾zz¯𝐷1subscript𝑅QSsubscript𝑅CB𝑡subscriptsuperscriptsubscript𝑅QSsubscript𝑅CB𝑡subscript𝐾zz𝑟𝐷𝑟d𝑟\bar{K}_{\rm zz}{+}\bar{D}=\frac{1}{R_{\rm QS}-R_{\rm CB}\left(t\right)}\int^{% R_{\rm QS}}_{R_{\rm CB}\left(t\right)}K_{\rm zz}\left(r\right){+}D\left(r% \right){\rm d}r.over¯ start_ARG italic_K end_ARG start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT + over¯ start_ARG italic_D end_ARG = divide start_ARG 1 end_ARG start_ARG italic_R start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT - italic_R start_POSTSUBSCRIPT roman_CB end_POSTSUBSCRIPT ( italic_t ) end_ARG ∫ start_POSTSUPERSCRIPT italic_R start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT roman_CB end_POSTSUBSCRIPT ( italic_t ) end_POSTSUBSCRIPT italic_K start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT ( italic_r ) + italic_D ( italic_r ) roman_d italic_r . (26)

The above equations can be solved analytically if a plane-parallel atmosphere and constant average diffusion coefficients are assumed,

K¯zz+D¯1RQSR0R0RQSKzz(r)+D(r)dr,subscript¯𝐾zz¯𝐷1subscript𝑅QSsubscript𝑅0subscriptsuperscriptsubscript𝑅QSsubscript𝑅0subscript𝐾zz𝑟𝐷𝑟d𝑟\bar{K}_{\rm zz}{+}\bar{D}\approx\frac{1}{R_{\rm QS}-R_{0}}\int^{R_{\rm QS}}_{% R_{0}}K_{\rm zz}\left(r\right){+}D\left(r\right){\rm d}r,over¯ start_ARG italic_K end_ARG start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT + over¯ start_ARG italic_D end_ARG ≈ divide start_ARG 1 end_ARG start_ARG italic_R start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT - italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ∫ start_POSTSUPERSCRIPT italic_R start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT italic_K start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT ( italic_r ) + italic_D ( italic_r ) roman_d italic_r , (27)

with the lower limit set to the planetary surface (R0subscript𝑅0R_{0}italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT). Balancing Equations 24 and 25,

δDLdδDLdt=(K¯zz+D¯)(1χQSnQSχCBnCB).subscript𝛿DLdsubscript𝛿DLd𝑡subscript¯𝐾zz¯𝐷1subscript𝜒QSsubscript𝑛QSsubscript𝜒CBsubscript𝑛CB\delta_{\rm DL}\frac{{\rm d}\delta_{\rm DL}}{{\rm d}t}=\left(\bar{K}_{\rm zz}+% \bar{D}\right)\left(1-\frac{\chi_{\rm QS}n_{\rm QS}}{\chi_{\rm CB}n_{\rm CB}}% \right).italic_δ start_POSTSUBSCRIPT roman_DL end_POSTSUBSCRIPT divide start_ARG roman_d italic_δ start_POSTSUBSCRIPT roman_DL end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_t end_ARG = ( over¯ start_ARG italic_K end_ARG start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT + over¯ start_ARG italic_D end_ARG ) ( 1 - divide start_ARG italic_χ start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT end_ARG start_ARG italic_χ start_POSTSUBSCRIPT roman_CB end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_CB end_POSTSUBSCRIPT end_ARG ) . (28)

The mole fraction and number density ratio is found by balancing Equations 23 and 24,

χQSnQSχCBnCB=K¯zz+D¯vQSδDL+K¯zz+D¯,subscript𝜒QSsubscript𝑛QSsubscript𝜒CBsubscript𝑛CBsubscript¯𝐾zz¯𝐷subscript𝑣QSsubscript𝛿DLsubscript¯𝐾zz¯𝐷\frac{\chi_{\rm QS}n_{\rm QS}}{\chi_{\rm CB}n_{\rm CB}}=\frac{\bar{K}_{\rm zz}% +\bar{D}}{v_{\rm QS}\delta_{\rm DL}+\bar{K}_{\rm zz}+\bar{D}},divide start_ARG italic_χ start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT end_ARG start_ARG italic_χ start_POSTSUBSCRIPT roman_CB end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT roman_CB end_POSTSUBSCRIPT end_ARG = divide start_ARG over¯ start_ARG italic_K end_ARG start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT + over¯ start_ARG italic_D end_ARG end_ARG start_ARG italic_v start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT roman_DL end_POSTSUBSCRIPT + over¯ start_ARG italic_K end_ARG start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT + over¯ start_ARG italic_D end_ARG end_ARG , (29)

which can be inserted into Equation 28 and integrated,

0RQSR0(δDLK¯zz+D¯+1vQS)dδDL=0tdt,subscriptsuperscriptsubscript𝑅QSsubscript𝑅00subscript𝛿DLsubscript¯𝐾zz¯𝐷1subscript𝑣QSdifferential-dsubscript𝛿DLsubscriptsuperscript𝑡0differential-d𝑡\int^{R_{\rm QS}{-}R_{0}}_{0}\left(\frac{\delta_{\rm DL}}{\bar{K}_{\rm zz}+% \bar{D}}+\frac{1}{v_{\rm QS}}\right){\rm d}\delta_{\rm DL}=\int^{t}_{0}{\rm d}t,∫ start_POSTSUPERSCRIPT italic_R start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT - italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( divide start_ARG italic_δ start_POSTSUBSCRIPT roman_DL end_POSTSUBSCRIPT end_ARG start_ARG over¯ start_ARG italic_K end_ARG start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT + over¯ start_ARG italic_D end_ARG end_ARG + divide start_ARG 1 end_ARG start_ARG italic_v start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT end_ARG ) roman_d italic_δ start_POSTSUBSCRIPT roman_DL end_POSTSUBSCRIPT = ∫ start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_d italic_t , (30)

yielding,

t=(RQSR0)22(K¯zz+D¯)+RQSR0vQS(RQSR0)22Kzz𝑡superscriptsubscript𝑅QSsubscript𝑅022subscript¯𝐾zz¯𝐷subscript𝑅QSsubscript𝑅0subscript𝑣QSsuperscriptsubscript𝑅QSsubscript𝑅022subscript𝐾zz\begin{split}t&=\frac{\left(R_{\rm QS}-R_{0}\right)^{2}}{2\left(\bar{K}_{\rm zz% }+\bar{D}\right)}+\frac{R_{\rm QS}-R_{0}}{v_{\rm QS}}\\ &\approx\frac{\left(R_{\rm QS}-R_{0}\right)^{2}}{2{K}_{\rm zz}}\end{split}start_ROW start_CELL italic_t end_CELL start_CELL = divide start_ARG ( italic_R start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT - italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 ( over¯ start_ARG italic_K end_ARG start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT + over¯ start_ARG italic_D end_ARG ) end_ARG + divide start_ARG italic_R start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT - italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_v start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT end_ARG end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL ≈ divide start_ARG ( italic_R start_POSTSUBSCRIPT roman_QS end_POSTSUBSCRIPT - italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_K start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT end_ARG end_CELL end_ROW (31)

Evaluating Equation 31 for typical values relevant to Earth’s atmosphere, with K¯zz102m2s1similar-tosubscript¯𝐾zzsuperscript102superscriptm2superscripts1\bar{K}_{\rm zz}{\sim}10^{2}~{}{\rm m^{2}~{}s^{-1}}over¯ start_ARG italic_K end_ARG start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (Liu 2021), suggests that the time required for steady state to apply is geologically negligible (40yearsless-than-or-similar-toabsent40years{\lesssim}40~{}{\rm years}≲ 40 roman_years). In fact, general circulation atmospheric models suggest that exoplanets may possess significantly higher eddy diffusion coefficients because they reside in more extreme environments (e.g., Parmentier et al. 2013; Charnay et al. 2015), indicating that steady state may be established in even shorter timescales. The amount of atmospheric depletion that occurs during the onset of steady state depends on the hydrogen outflow rate at the quasi-sonic point. Even with inviscid XUV-induced photoevaporation, the overall impact is likely negligible because total mass loss requires 103105superscript103superscript10510^{3}{-}10^{5}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT years (Figure 5). Higher mass loss rates, however, may result in complete atmospheric loss during the steady-state onset, and we discuss this possibility in section 4.1.

3.2 Representative simulations

Having assessed the rapid onset of steady state in planets with hydrogen-rich atmospheres, we now present several representative simulations for exoplanets experiencing mass loss. Because hydrogen loss at the quasi-sonic point has to be supplied from below, diffusion-limited mass loss at any point is constrained by the region just below it. The global diffusion-limited mass loss rate is therefore equal at all points in the atmosphere. Unlike Hunten (1973) and Hunten et al. (1987), we evaluate mass loss at the photosphere (subscript p; where optical depth τ=2/3𝜏23\tau{=}2/3italic_τ = 2 / 3) because it is the temperature minimum of the atmosphere and because the compositional gradient is small (KzzDmuch-greater-thansubscript𝐾zz𝐷K_{\rm zz}{\gg}Ditalic_K start_POSTSUBSCRIPT roman_zz end_POSTSUBSCRIPT ≫ italic_D; Modirrousta-Galian & Korenaga 2023), allowing us to use Equation 7. We replace the experimentally-derived binary diffusion coefficient, bijsubscript𝑏ijb_{\rm ij}italic_b start_POSTSUBSCRIPT roman_ij end_POSTSUBSCRIPT, with the more general molecular diffusion coefficient and photospheric number density, npDsubscript𝑛p𝐷n_{\rm p}Ditalic_n start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT italic_D. To estimate the diffusion coefficient we apply the Chapman-Enskog hard sphere approximation (Chapman & Cowling 1970) with the average particle kinetic radius taken as that of an atomic hydrogen-helium mixture (σ12=2×1010msubscript𝜎122superscript1010m\sigma_{\rm 12}{=}2{\times}10^{-10}~{}{\rm m}italic_σ start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT = 2 × 10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT roman_m),

M˙1=1.1×105χp1χp(μ¯pμ11)(μ1μ2+1)12(MpM)(Tp1000K)12kgs11.1×105χp1χp(μ¯pμ11)(MpM)(Tp1000K)12kgs1,subscript˙𝑀11.1superscript105subscript𝜒p1subscript𝜒psubscript¯𝜇psubscript𝜇11superscriptsubscript𝜇1subscript𝜇2112subscript𝑀psubscript𝑀direct-sumsuperscriptsubscript𝑇p1000K12kgsuperscripts11.1superscript105subscript𝜒p1subscript𝜒psubscript¯𝜇psubscript𝜇11subscript𝑀psubscript𝑀direct-sumsuperscriptsubscript𝑇p1000K12kgsuperscripts1\begin{split}\dot{M}_{1}&=1.1{\times}10^{5}\frac{\chi_{\rm p}}{1{-}\chi_{\rm p% }}\left(\frac{\bar{\mu}_{\rm p}}{\mu_{1}}{-}1\right)\left(\frac{\mu_{1}}{\mu_{% 2}}+1\right)^{\frac{1}{2}}\left(\frac{M_{\rm p}}{M_{\oplus}}\right)\left(\frac% {T_{\rm p}}{1000~{}{\rm K}}\right)^{-\frac{1}{2}}~{}{\rm kg~{}s^{-1}}\\ &\approx 1.1{\times}10^{5}\frac{\chi_{\rm p}}{1{-}\chi_{\rm p}}\left(\frac{% \bar{\mu}_{\rm p}}{\mu_{1}}{-}1\right)\left(\frac{M_{\rm p}}{M_{\oplus}}\right% )\left(\frac{T_{\rm p}}{1000~{}{\rm K}}\right)^{-\frac{1}{2}}~{}{\rm kg~{}s^{-% 1}},\end{split}start_ROW start_CELL over˙ start_ARG italic_M end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_CELL start_CELL = 1.1 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT divide start_ARG italic_χ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT end_ARG start_ARG 1 - italic_χ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT end_ARG ( divide start_ARG over¯ start_ARG italic_μ end_ARG start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT end_ARG start_ARG italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG - 1 ) ( divide start_ARG italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG + 1 ) start_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT ( divide start_ARG italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT end_ARG ) ( divide start_ARG italic_T start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT end_ARG start_ARG 1000 roman_K end_ARG ) start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT roman_kg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL ≈ 1.1 × 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT divide start_ARG italic_χ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT end_ARG start_ARG 1 - italic_χ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT end_ARG ( divide start_ARG over¯ start_ARG italic_μ end_ARG start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT end_ARG start_ARG italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG - 1 ) ( divide start_ARG italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT end_ARG start_ARG italic_M start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT end_ARG ) ( divide start_ARG italic_T start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT end_ARG start_ARG 1000 roman_K end_ARG ) start_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 2 end_ARG end_POSTSUPERSCRIPT roman_kg roman_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , end_CELL end_ROW (32)

where (μ1/μ2+1)1/21superscriptsubscript𝜇1subscript𝜇21121\left(\mu_{1}/\mu_{2}{+}1\right)^{1/2}{\approx}1( italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT / italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + 1 ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ≈ 1. Simulations suggest that eddy diffusion is consistently greater than molecular diffusion in hydrogen-rich atmospheres (Modirrousta-Galian & Korenaga 2023), so χpsubscript𝜒p\chi_{\rm p}italic_χ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT can be approximated with MH/(MH+Mz)subscript𝑀Hsubscript𝑀Hsubscript𝑀zM_{\rm H}/\left(M_{\rm H}{+}M_{\rm z}\right)italic_M start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT / ( italic_M start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT + italic_M start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT ), where MHsubscript𝑀HM_{\rm H}italic_M start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT and Mzsubscript𝑀zM_{\rm z}italic_M start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT are the total atmospheric hydrogen and heavy components. For non-gas giant planets, the heavy component originates mainly from outgassed carbon dioxide during magma ocean solidification (Elkins-Tanton 2008; Lebrun et al. 2013; Salvador et al. 2017; Bower et al. 2019), with water being released more gradually because of its high solubility in magma (Lichtenberg et al. 2021; Miyazaki & Korenaga 2022). The initial total non-hydrous volatile budget of silicate planets is thought to be of the order 0.01%similar-toabsentpercent0.01{\sim}0.01\%∼ 0.01 % by weight (Hirschmann & Dasgupta 2009; Marty 2012) so that Mz104Mpsimilar-tosubscript𝑀zsuperscript104subscript𝑀pM_{\rm z}{\sim}10^{-4}M_{\rm p}italic_M start_POSTSUBSCRIPT roman_z end_POSTSUBSCRIPT ∼ 10 start_POSTSUPERSCRIPT - 4 end_POSTSUPERSCRIPT italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT. The initial accreted primordial atmosphere is less well known, with simulations providing a wide range of estimates (Lee & Chiang 2015; Ginzburg et al. 2016; Mordasini 2020), so we set it as a free parameter. We therefore approximate the mean molecular mass as,

μ¯=max[2.2amu,χpμH+(1χp)μCO2],¯𝜇2.2amusubscript𝜒psubscript𝜇H1subscript𝜒psubscript𝜇subscriptCO2\bar{\mu}=\max{\left[2.2~{}{\rm amu},\chi_{\rm p}\mu_{\rm H}{+}(1{-}\chi_{\rm p% })\mu_{\rm CO_{2}}\right]},over¯ start_ARG italic_μ end_ARG = roman_max [ 2.2 roman_amu , italic_χ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT italic_μ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT + ( 1 - italic_χ start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT ) italic_μ start_POSTSUBSCRIPT roman_CO start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT ] , (33)

with μHsubscript𝜇H\mu_{\rm H}italic_μ start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT and μCO2subscript𝜇subscriptCO2\mu_{\rm CO_{2}}italic_μ start_POSTSUBSCRIPT roman_CO start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUBSCRIPT being the molecular masses of atomic hydrogen and carbon dioxide. Our test planet is GJ 357 b, which is a super-Earth with a mass, radius, and equilibrium temperature of 1.84±0.31Mplus-or-minus1.840.31subscript𝑀direct-sum1.84{\pm}0.31~{}M_{\oplus}1.84 ± 0.31 italic_M start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT (Luque et al. 2019), 1.20±0.06Rplus-or-minus1.200.06subscript𝑅direct-sum1.20{\pm}0.06~{}R_{\oplus}1.20 ± 0.06 italic_R start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT (Oddo et al. 2023), and 525K525K525~{}{\rm K}525 roman_K (Luque et al. 2019) respectively. The X-ray luminosity and age of the host star were determined through XMM-Newton observations to be Lx=1018.73Wsubscript𝐿xsuperscript1018.73WL_{\rm x}{=}10^{18.73}~{}{\rm W}italic_L start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 18.73 end_POSTSUPERSCRIPT roman_W and 5Gyr5Gyr5~{}{\rm Gyr}5 roman_Gyr (Modirrousta-Galian et al. 2020). The X-ray and ultraviolet luminosity evolution of the star are estimated through empirical relations (Penz & Micela 2008; Sanz-Forcada et al. 2011, see also Chadney et al. 2015; Johnstone et al. 2021), which we extend back to right after the T Tauri stage (age10Myrsimilar-toabsent10Myr{\sim}10~{}{\rm Myr}∼ 10 roman_Myr). For the atmospheric size we employ the model of Lopez & Fortney (2014), which provides the atmospheric radius as a function of the planetary mass, the atmospheric mass fraction, the equilibrium temperature, and the planetary age. To provide a basis for comparison with the diffusion-limited mass loss model, we also use the energy-limited approximation of Watson et al. (1981) and the hydro-based model of Kubyshkina et al. (2018). Both models assume inviscid free advection, but the latter considers thermal heating and chemistry, usually resulting in higher mass loss rates.

Refer to caption
Figure 5: Atmospheric mass as a function of time for super-Earth GJ 357 b (Mp=1.84±0.31Msubscript𝑀pplus-or-minus1.840.31subscript𝑀direct-sumM_{\rm p}{=}1.84{\pm}0.31~{}M_{\oplus}italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = 1.84 ± 0.31 italic_M start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT, Rp=1.20±0.06Rsubscript𝑅pplus-or-minus1.200.06subscript𝑅direct-sumR_{\rm p}{=}1.20{\pm}0.06~{}R_{\oplus}italic_R start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT = 1.20 ± 0.06 italic_R start_POSTSUBSCRIPT ⊕ end_POSTSUBSCRIPT, Teq=525Ksubscript𝑇eq525KT_{\rm eq}{=}525~{}{\rm K}italic_T start_POSTSUBSCRIPT roman_eq end_POSTSUBSCRIPT = 525 roman_K, Lx=1018.73Wsubscript𝐿xsuperscript1018.73WL_{\rm x}{=}10^{18.73}~{}{\rm W}italic_L start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 18.73 end_POSTSUPERSCRIPT roman_W) modeled using the hydro-based (red line), energy-limited (orange line), and diffusion-limited (blue line) mass loss models. The solid and dashed lines are for initial atmospheric hydrogen budgets of MH=0.01Mpsubscript𝑀H0.01subscript𝑀pM_{\rm H}{=}0.01M_{\rm p}italic_M start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT = 0.01 italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT and 0.1Mp0.1subscript𝑀p0.1M_{\rm p}0.1 italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT respectively.

Figure 5 shows the mass loss evolution of GJ 357 b for initial atmospheric hydrogen budgets of MH=0.1Mpsubscript𝑀H0.1subscript𝑀pM_{\rm H}{=}0.1M_{\rm p}italic_M start_POSTSUBSCRIPT roman_H end_POSTSUBSCRIPT = 0.1 italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT and 0.01Mp0.01subscript𝑀p0.01M_{\rm p}0.01 italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT. The diffusion-limited model predicts lower mass loss rates than the hydro-based and energy-limited models, enabling the retention of atmospheric hydrogen over significantly longer timescales. Assuming the stellar age of 5Gyr5Gyr5~{}{\rm Gyr}5 roman_Gyr is coeval with that of the planet (Modirrousta-Galian et al. 2020), GJ 357 b probably accreted an atmosphere less than 0.16Mpsimilar-toabsent0.16subscript𝑀p{\sim}0.16~{}M_{\rm p}∼ 0.16 italic_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT; otherwise, it would have survived photoevaporation and still existed. This model, therefore, provides tighter constraints on the mass loss evolution of GJ 357 b than the hydro-based model, which predicts an initial hydrogen reservoir less than 21Mpsimilar-toabsent21subscriptMp{\sim}21~{}{\rm M_{p}}∼ 21 roman_M start_POSTSUBSCRIPT roman_p end_POSTSUBSCRIPT (Modirrousta-Galian et al. 2020).

4 Discussion

4.1 On the self-consistency of mass loss and fractionation

Having evaluated diffusion-limited mass loss (section 2.1), the crossover mass (section 2.2), and the transition to the diffusion-limited regime (section 3.1), we now discuss the self-consistency of mass loss and fractionation.

In Modirrousta-Galian & Korenaga (2023), atmospheric evaporation is categorized into three regimes. In the first regime, mass loss arises from internal energy generated during accretion; in the second regime, mass loss occurs through a combination of internal energy and incoming thermal radiation; and in the third regime, mass loss is driven by X-ray and ultraviolet irradiation. Selecting the appropriate mass loss regime depends on the thermodynamic properties of the system, which we explored in our previous paper. To keep our explanation concise and avoid reiterating what has already been discussed in Modirrousta-Galian & Korenaga (2023), we focus only on regimes one and three, which represent endmember cases when mass loss is driven by either internal energy or incoming X-ray and ultraviolet irradiation.

Regime one (also known as core-powered mass loss) is very efficient and can remove a primordial atmosphere in a few days (Modirrousta-Galian & Korenaga 2023), which is too fast for steady state to activate (40yearsless-than-or-similar-toabsent40years{\lesssim}40~{}{\rm years}≲ 40 roman_years) and diffusion-limited mass loss to apply. This suggests that fractionation is unlikely to occur as a result of geological events that bring a planet into regime one, such as giant impacts (e.g., Canup 2008; Lock et al. 2018) or late accretion (e.g., Marchi et al. 2018, 2023). In contrast, regime three (also known as XUV photoevaporation) proceeds more slowly, requiring 103105superscript103superscript10510^{3}{-}10^{5}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT years to remove an atmosphere from super-Earth and sub-Neptune exoplanets (Figure 5). These longer timescales allow for steady state to activate and diffusion-limited mass loss to apply. Consequently, whereas a planet in regime three may initially experience mass loss according to traditional inviscid free advection models (e.g., the energy-limited and the hydro-based models), it will rapidly shift to diffusion-limited mass loss after the onset of steady state.

4.2 On modeling photoevaporation accurately

Of the three regimes of atmospheric evaporation, the third is the most widely discussed because of its prolonged duration and observability through Lyman α𝛼\alphaitalic_α spectroscopy (Linsky et al. 2010; Lecavelier des Etangs et al. 2012; Vidal-Madjar et al. 2013; Rockcliffe et al. 2023). The prevailing model employed for assessing mass loss in the third regime is the energy-limited model (Watson et al. 1981). Despite its apparent simplicity and intuitive nature, this model is beset by several well-documented limitations (e.g., Kubyshkina et al. 2018; Krenn et al. 2021; Modirrousta-Galian & Korenaga 2023). Most notably, it neglects the influence of incoming thermal radiation and erroneously assumes that incoming XUV energy is deposited at the XUV-photosphere (τXUV=2/3subscript𝜏XUV23\tau_{\rm XUV}{=}2/3italic_τ start_POSTSUBSCRIPT roman_XUV end_POSTSUBSCRIPT = 2 / 3). These assumptions are hard to justify because exoplanets are often highly irradiated, and in the case of primordial atmospheres, the XUV photosphere is usually above the sonic point (Sekiya et al. 1980, 1981). Therefore, the sonic point (or more correctly, the quasi-sonic point; Modirrousta-Galian & Korenaga submitted) is a better location for assessing XUV energy deposition.

To overcome these limitations, Kubyshkina et al. (2018) proposed the Hydro-based model, which incorporates a more realistic chemical analysis and addresses the limitations of the energy-limited approximation. However, this model also makes several assumptions that are likely to be invalid. First, they assume that gas accelerates indefinitely, which is not always accurate (Modirrousta-Galian & Korenaga submitted). Second, it assumes that mass loss occurs through inviscid free advection for prolonged durations, and it is therefore efficient in stripping hydrogen from planetary atmospheres. Consequently, the question still remains as to why some exoplanets maintain their hydrogen-rich atmospheres while others do not. If, indeed, fractionation is significant, this could provide a plausible explanation for the observed diversity in exoplanet atmospheres because extreme hydrogen loss is unsustainable if fractionation takes place and mass loss becomes diffusion-limited. In other words, the energy-limited and hydro-based models are both limited in their ability to describe mass loss, and they should be applied judiciously. In contrast, the diffusion-limited mass loss model incorporates the effects of chemical fractionation, and it therefore provides a more realistic description of mass loss.

4.3 On the observability of chemical fractionation in exoplanetary atmospheres

Our model suggests that planetary atmospheres become chemically fractionated because of XUV-induced photoevaporation, transitioning from a primordial hydrogen-rich atmosphere to a secondary heavier one. The specific details of this evolution depend on the planet’s bulk properties and its environment, requiring a case-by-case evaluation. This evolution may be detectable using atmospheric spectroscopy (Burrows 2014; Madhusudhan 2019). However, observations are limited to cloudless planets because clouds obstruct visibility into deeper atmospheric layers (Bétrémieux & Swain 2017, 2018) and create pronounced chemical stratification, likely masking the effects of diffusion-limited chemical fractionation. The recent launch of the James Webb Space Telescope and upcoming missions like Ariel and Twinkle hold promise for advancing our understanding of the link between chemical fractionation and planetary evolution.

5 Concluding Remarks

In this study, we suggest that X-ray and ultraviolet-induced photoevaporation in primordial atmospheres results in compositional fractionation. When studying Earth, unlike earlier research, we assess fractionation at the exobase. We find that the crossover mass ranges from 415amu415amu4{-}15~{}{\rm amu}4 - 15 roman_amu for Earth, with it depending primarily on the altitude at which it is evaluated, the hydrogen mole fraction, the temperature ratio to the one-fourth power, and the local mean molecular mass. Building on this result, we demonstrate that mass fractionation results in a transient disequilibrium, forming a hydrogen-depleted layer that grows from the top of the atmosphere downward. After reaching steady state, hydrodynamic outflow stops because diffusion can no longer supply the loss of hydrogen from advection. Last, we find that subsequent mass loss is diffusion-limited, yielding significantly lower mass loss rates than free advection, thus allowing planets to sustain hydrogen-rich atmospheres for significantly longer timescales.

Acknowledgements

This work was sponsored by the US National Science Foundation EAR-2224727 and the US National Aeronautics and Space Administration under Cooperative Agreement No. 80NSSC19M0069 issued through the Science Mission Directorate. This work was also supported in part by the facilities and staff of the Yale University Faculty of Arts and Sciences High Performance Computing Center. We thank the anonymous reviewer, whose comments helped us to improve the clarity of our manuscript.

References

  • (1)
  • Bétrémieux & Swain (2017) Bétrémieux, Y. & Swain, M. R. (2017), ‘An analytical formalism accounting for clouds and other ‘surfaces’ for exoplanet transmission spectroscopy’, MNRAS 467(3), 2834–2844.
  • Bétrémieux & Swain (2018) Bétrémieux, Y. & Swain, M. R. (2018), ‘The Hidden Depths of Planetary Atmospheres’, ApJ 865(1), 12.
  • Bower et al. (2019) Bower, D. J., Kitzmann, D., Wolf, A. S., Sanan, P., Dorn, C. & Oza, A. V. (2019), ‘Linking the evolution of terrestrial interiors and an early outgassed atmosphere to astrophysical observations’, A&A 631, A103.
  • Burrows (2014) Burrows, A. S. (2014), ‘Spectra as windows into exoplanet atmospheres’, Proceedings of the National Academy of Science 111(35), 12601–12609.
  • Caldiroli et al. (2021) Caldiroli, A., Haardt, F., Gallo, E., Spinelli, R., Malsky, I. & Rauscher, E. (2021), ‘Irradiation-driven escape of primordial planetary atmospheres. I. The ATES photoionization hydrodynamics code’, A&A 655, A30.
  • Canup (2008) Canup, R. M. (2008), ‘Lunar-forming collisions with pre-impact rotation’, Icarus 196(2), 518–538.
  • Catling & Kasting (2017) Catling, D. C. & Kasting, J. F. (2017), Escape of Atmospheres to Space, Cambridge University Press, p. 129–168.
  • Chadney et al. (2015) Chadney, J. M., Galand, M., Unruh, Y. C., Koskinen, T. T. & Sanz-Forcada, J. (2015), ‘XUV-driven mass loss from extrasolar giant planets orbiting active stars’, Icarus 250, 357–367.
  • Chapman & Cowling (1970) Chapman, S. & Cowling, T. G. (1970), The mathematical theory of non-uniform gases. an account of the kinetic theory of viscosity, thermal conduction and diffusion in gases.
  • Charnay et al. (2015) Charnay, B., Meadows, V. & Leconte, J. (2015), ‘3D Modeling of GJ1214b’s Atmosphere: Vertical Mixing Driven by an Anti-Hadley Circulation’, ApJ 813(1), 15.
  • Elkins-Tanton (2008) Elkins-Tanton, L. T. (2008), ‘Linked magma ocean solidification and atmospheric growth for Earth and Mars’, Earth and Planetary Science Letters 271(1-4), 181–191.
  • Emmert (2015) Emmert, J. T. (2015), ‘Thermospheric mass density: A review’, Advances in Space Research 56(5), 773–824.
  • Emmert et al. (2021) Emmert, J. T., Drob, D. P., Picone, J. M., Siskind, D. E., Jones, M., Mlynczak, M. G., Bernath, P. F., Chu, X., Doornbos, E., Funke, B., Goncharenko, L. P., Hervig, M. E., Schwartz, M. J., Sheese, P. E., Vargas, F., Williams, B. P. & Yuan, T. (2021), ‘NRLMSIS 2.0: A Whole Atmosphere Empirical Model of Temperature and Neutral Species Densities’, Earth and Space Science 8(3), e01321.
  • Fulton et al. (2017) Fulton, B. J., Petigura, E. A., Howard, A. W., Isaacson, H., Marcy, G. W., Cargile, P. A., Hebb, L., Weiss, L. M., Johnson, J. A., Morton, T. D., Sinukoff, E., Crossfield, I. J. M. & Hirsch, L. A. (2017), ‘The California-Kepler Survey. III. A Gap in the Radius Distribution of Small Planets’, AJ 154, 109.
  • Ginzburg et al. (2016) Ginzburg, S., Schlichting, H. E. & Sari, R. (2016), ‘Super-Earth Atmospheres: Self-consistent Gas Accretion and Retention’, ApJ 825, 29.
  • Gross (1972) Gross, S. H. (1972), ‘On the exospheric temperature of hydrogen-dominated planetary atmospheres.’, Journal of Atmospheric Sciences 29, 214–218.
  • Hirschmann & Dasgupta (2009) Hirschmann, M. M. & Dasgupta, R. (2009), ‘The H/C ratios of Earth’s near-surface and deep reservoirs, and consequences for deep Earth volatile cycles’, Chemical Geology 262(1-2), 4–16.
  • Hunten (1973) Hunten, D. M. (1973), ‘The Escape of Light Gases from Planetary Atmospheres.’, Journal of Atmospheric Sciences 30(8), 1481–1494.
  • Hunten et al. (1987) Hunten, D. M., Pepin, R. O. & Walker, J. C. G. (1987), ‘Mass fractionation in hydrodynamic escape’, Icarus 69, 532–549.
  • Johnstone et al. (2021) Johnstone, C. P., Bartel, M. & Güdel, M. (2021), ‘The active lives of stars: A complete description of the rotation and XUV evolution of F, G, K, and M dwarfs’, A&A 649, A96.
  • Joshi et al. (2019) Joshi, P. P., Phal, Y. D. & Waldrop, L. S. (2019), ‘Quantification of the Vertical Transport and Escape of Atomic Hydrogen in the Terrestrial Upper Atmosphere’, Journal of Geophysical Research (Space Physics) 124(12), 10,468–10,481.
  • Krenn et al. (2021) Krenn, A. F., Fossati, L., Kubyshkina, D. & Lammer, H. (2021), ‘A critical assessment of the applicability of the energy-limited approximation for estimating exoplanetary mass-loss rates’, A&A 650, A94.
  • Kubyshkina et al. (2018) Kubyshkina, D., Fossati, L., Erkaev, N. V., Cubillos, P. E., Johnstone, C. P., Kislyakova, K. G., Lammer, H., Lendl, M. & Odert, P. (2018), ‘Overcoming the Limitations of the Energy-limited Approximation for Planet Atmospheric Escape’, ApJL 866, L18.
  • Lebrun et al. (2013) Lebrun, T., Massol, H., ChassefièRe, E., Davaille, A., Marcq, E., Sarda, P., Leblanc, F. & Brandeis, G. (2013), ‘Thermal evolution of an early magma ocean in interaction with the atmosphere’, Journal of Geophysical Research (Planets) 118(6), 1155–1176.
  • Lecavelier des Etangs et al. (2012) Lecavelier des Etangs, A., Bourrier, V., Wheatley, P. J., Dupuy, H., Ehrenreich, D., Vidal-Madjar, A., Hébrard, G., Ballester, G. E., Désert, J. M., Ferlet, R. & Sing, D. K. (2012), ‘Temporal variations in the evaporating atmosphere of the exoplanet HD 189733b’, A&A 543, L4.
  • Lee & Chiang (2015) Lee, E. J. & Chiang, E. (2015), ‘To Cool is to Accrete: Analytic Scalings for Nebular Accretion of Planetary Atmospheres’, ApJ 811(1), 41.
  • Leuenberger & Lang (2002) Leuenberger, M. & Lang, C. (2002), Thermal diffusion: An important aspect in studies of static air columns such as firn air, sand dunes and soil air, Technical report, International Atomic Energy Agency (IAEA).
  • Lichtenberg et al. (2021) Lichtenberg, T., Bower, D. J., Hammond, M., Boukrouche, R., Sanan, P., Tsai, S.-M. & Pierrehumbert, R. T. (2021), ‘Vertically Resolved Magma Ocean-Protoatmosphere Evolution: H22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT, H22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO, CO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT, CH44{}_{4}start_FLOATSUBSCRIPT 4 end_FLOATSUBSCRIPT, CO, O22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT, and N22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT as Primary Absorbers’, Journal of Geophysical Research (Planets) 126(2), e06711.
  • Lindzen (1971) Lindzen, R. S. (1971), Tides and Gravity Waves in the Upper Atmosphere, in G. Fiocco, ed., ‘Mesospheric Models and Related Experiments’, Vol. 25 of Astrophysics and Space Science Library, p. 122.
  • Lindzen (1981) Lindzen, R. S. (1981), ‘Turbulence and stress owing to gravity wave and tidal breakdown’, J. Geophys. Res.: Oceans 86(C10), 9707–9714.
  • Linsky et al. (2010) Linsky, J. L., Yang, H., France, K., Froning, C. S., Green, J. C., Stocke, J. T. & Osterman, S. N. (2010), ‘Observations of Mass Loss from the Transiting Exoplanet HD 209458b’, ApJ 717(2), 1291–1299.
  • Liu (2021) Liu, H.-L. (2021), ‘Effective Vertical Diffusion by Atmospheric Gravity Waves’, Geophys. Res. Lett. 48(1), e91474.
  • Lock et al. (2018) Lock, S. J., Stewart, S. T., Petaev, M. I., Leinhardt, Z., Mace, M. T., Jacobsen, S. B. & Cuk, M. (2018), ‘The Origin of the Moon Within a Terrestrial Synestia’, Journal of Geophysical Research (Planets) 123(4), 910–951.
  • Lopez & Fortney (2014) Lopez, E. D. & Fortney, J. J. (2014), ‘Understanding the Mass-Radius Relation for Sub-neptunes: Radius as a Proxy for Composition’, ApJ 792, 1.
  • Luque et al. (2019) Luque, R., Pallé, E., Kossakowski, D., Dreizler, S., Kemmer, J., Espinoza, N., Burt, J., Anglada-Escudé, G., Béjar, V. J. S., Caballero, J. A., Collins, K. A., Collins, K. I., Cortés-Contreras, M., Díez-Alonso, E., Feng, F., Hatzes, A., Hellier, C., Henning, T., Jeffers, S. V., Kaltenegger, L., Kürster, M., Madden, J., Molaverdikhani, K., Montes, D., Narita, N., Nowak, G., Ofir, A., Oshagh, M., Parviainen, H., Quirrenbach, A., Reffert, S., Reiners, A., Rodríguez-López, C., Schlecker, M., Stock, S., Trifonov, T., Winn, J. N., Zapatero Osorio, M. R., Zechmeister, M., Amado, P. J., Anderson, D. R., Batalha, N. E., Bauer, F. F., Bluhm, P., Burke, C. J., Butler, R. P., Caldwell, D. A., Chen, G., Crane, J. D., Dragomir, D., Dressing, C. D., Dynes, S., Jenkins, J. M., Kaminski, A., Klahr, H., Kotani, T., Lafarga, M., Latham, D. W., Lewin, P., McDermott, S., Montañés-Rodríguez, P., Morales, J. C., Murgas, F., Nagel, E., Pedraz, S., Ribas, I., Ricker, G. R., Rowden, P., Seager, S., Shectman, S. A., Tamura, M., Teske, J., Twicken, J. D., Vanderspeck, R., Wang, S. X. & Wohler, B. (2019), ‘Planetary system around the nearby M dwarf GJ 357 including a transiting, hot, Earth-sized planet optimal for atmospheric characterization’, A&A 628, A39.
  • Madhusudhan (2019) Madhusudhan, N. (2019), ‘Exoplanetary Atmospheres: Key Insights, Challenges, and Prospects’, Annu. Rev. Astron. Astrophys. 57, 617–663.
  • Marchi et al. (2018) Marchi, S., Canup, R. M. & Walker, R. J. (2018), ‘Heterogeneous delivery of silicate and metal to the Earth by large planetesimals’, Nature Geoscience 11(1), 77–81.
  • Marchi et al. (2023) Marchi, S., Rufu, R. & Korenaga, J. (2023), ‘Long-lived volcanic resurfacing of Venus driven by early collisions’, Nature Astronomy 7, 1180–1187.
  • Marrero & Mason (1972) Marrero, T. R. & Mason, E. A. (1972), ‘Gaseous Diffusion Coefficients’, Journal of Physical and Chemical Reference Data 1(1), 3–118.
  • Marty (2012) Marty, B. (2012), ‘The origins and concentrations of water, carbon, nitrogen and noble gases on Earth’, Earth and Planetary Science Letters 313, 56–66.
  • Mason & Marrero (1970) Mason, E. & Marrero, T. (1970), The diffusion of atoms and molecules, Vol. 6 of Advances in Atomic and Molecular Physics, Academic Press, pp. 155–232.
    https://www.sciencedirect.com/science/article/pii/S0065219908602055
  • Miyazaki & Korenaga (2022) Miyazaki, Y. & Korenaga, J. (2022), ‘Inefficient Water Degassing Inhibits Ocean Formation on Rocky Planets: An Insight from Self-Consistent Mantle Degassing Models’, Astrobiology 22(6), 713–734.
  • Modirrousta-Galian & Korenaga (2023) Modirrousta-Galian, D. & Korenaga, J. (2023), ‘The Three Regimes of Atmospheric Evaporation for Super-Earths and Sub-Neptunes’, ApJ 943(1), 11.
  • Modirrousta-Galian & Korenaga (submitted) Modirrousta-Galian, D. & Korenaga, J. (submitted), ‘The exobase–sonic point relationship in hydrodynamic planetary atmospheres’.
  • Modirrousta-Galian et al. (2020) Modirrousta-Galian, D., Locci, D. & Micela, G. (2020), ‘The Bimodal Distribution in Exoplanet Radii: Considering Varying Core Compositions and H22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT Envelope’s Sizes’, ApJ 891(2), 158.
  • Mordasini (2020) Mordasini, C. (2020), ‘Planetary evolution with atmospheric photoevaporation. I. Analytical derivation and numerical study of the evaporation valley and transition from super-Earths to sub-Neptunes’, A&A 638, A52.
  • Oddo et al. (2023) Oddo, D., Dragomir, D., Brandeker, A., Osborn, H. P., Collins, K., Stassun, K. G., Astudillo-Defru, N., Bieryla, A., Howell, S. B., Ciardi, D. R., Quinn, S., Almenara, J. M., Briceño, C., Collins, K. I., Colón, K. D., Conti, D. M., Crouzet, N., Furlan, E., Gan, T., Gnilka, C. L., Goeke, R. F., Gonzales, E., Harris, M., Jenkins, J. M., Jensen, E. L. N., Latham, D., Law, N., Lund, M. B., Mann, A. W., Massey, B., Murgas, F., Ricker, G., Relles, H. M., Rowden, P., Schwarz, R. P., Schlieder, J., Shporer, A., Seager, S., Srdoc, G., Torres, G., Twicken, J. D., Vanderspek, R., Winn, J. N. & Ziegler, C. (2023), ‘Characterization of a Set of Small Planets with TESS and CHEOPS and an Analysis of Photometric Performance’, AJ 165(3), 134.
  • Öpik (1963) Öpik, E. J. (1963), ‘Selective Escape of Gases?’, Geophysical Journal 7(4), 490–506.
  • Oran et al. (1998) Oran, E., Oh, C. & Cybyk, B. (1998), ‘Direct simulation monte carlo: Recent advances and applications’, Annual Review of Fluid Mechanics 30(1), 403–441.
    https://doi.org/10.1146/annurev.fluid.30.1.403
  • Parker (1958) Parker, E. N. (1958), ‘Dynamics of the Interplanetary Gas and Magnetic Fields.’, ApJ 128, 664.
  • Parmentier et al. (2013) Parmentier, V., Showman, A. P. & Lian, Y. (2013), ‘3D mixing in hot Jupiters atmospheres. I. Application to the day/night cold trap in HD 209458b’, A&A 558, A91.
  • Penz & Micela (2008) Penz, T. & Micela, G. (2008), ‘X-ray induced mass loss effects on exoplanets orbiting dM stars’, A&A 479, 579–584.
  • Penz et al. (2008) Penz, T., Micela, G. & Lammer, H. (2008), ‘Influence of the evolving stellar X-ray luminosity distribution on exoplanetary mass loss’, A&A 477, 309–314.
  • Pope (2000) Pope, S. B. (2000), Turbulent Flows.
  • Porcelli & Pepin (2014) Porcelli, D. & Pepin, R. (2014), 6.17 - the origin of noble gases and major volatiles in the terrestrial planets, in H. D. Holland & K. K. Turekian, eds, ‘Treatise on Geochemistry (Second Edition)’, second edition edn, Elsevier, Oxford, pp. 383–406.
    https://www.sciencedirect.com/science/article/pii/B9780080959757004125
  • Rockcliffe et al. (2023) Rockcliffe, K. E., Newton, E. R., Youngblood, A., Duvvuri, G. M., Plavchan, P., Gao, P., Mann, A. W. & Lowrance, P. J. (2023), ‘The Variable Detection of Atmospheric Escape around the Young, Hot Neptune AU Mic b’, AJ 166(2), 77.
  • Salinas et al. (2016) Salinas, C. C. J. H., Chang, L. C., Liang, M.-C., Yue, J., Russell, J. & Mlynczak, M. (2016), ‘Impacts of SABER CO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT-based eddy diffusion coefficients in the lower thermosphere on the ionosphere/thermosphere’, Journal of Geophysical Research (Space Physics) 121(12), 12,080–12,092.
  • Salvador et al. (2017) Salvador, A., Massol, H., Davaille, A., Marcq, E., Sarda, P. & Chassefière, E. (2017), ‘The relative influence of H22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPTO and CO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT on the primitive surface conditions and evolution of rocky planets’, Journal of Geophysical Research (Planets) 122(7), 1458–1486.
  • Sanz-Forcada et al. (2011) Sanz-Forcada, J., Micela, G., Ribas, I., Pollock, A. M. T., Eiroa, C., Velasco, A., Solano, E. & García-Álvarez, D. (2011), ‘Estimation of the XUV radiation onto close planets and their evaporation’, A&A 532, A6.
  • Sekiya et al. (1981) Sekiya, M., Hayashi, C. & Kanazawa, K. (1981), ‘Dissipation of the Primordial Terrestrial Atmosphere Due to Irradiation of the Solar Far-UV during T Tauri Stage’, Progress of Theoretical Physics 66(4), 1301–1316.
  • Sekiya et al. (1980) Sekiya, M., Nakazawa, K. & Hayashi, C. (1980), ‘Dissipation of the Primordial Terrestrial Atmosphere Due to Irradiation of the Solar EUV’, Progress of Theoretical Physics 64(6), 1968–1985.
  • Shematovich et al. (2015) Shematovich, V. I., Bisikalo, D. V. & Ionov, D. E. (2015), Suprathermal Particles in XUV-Heated and Extended Exoplanetary Upper Atmospheres, in H. Lammer & M. Khodachenko, eds, ‘Characterizing Stellar and Exoplanetary Environments’, Vol. 411 of Astrophysics and Space Science Library, p. 105.
  • Swenson et al. (2021) Swenson, G. R., Vargas, F., Jones, M., Zhu, Y., Kaufmann, M., Yee, J. H. & Mlynczak, M. (2021), ‘Intra-Annual Variation of Eddy Diffusion (kzz𝑧𝑧{}_{zz}start_FLOATSUBSCRIPT italic_z italic_z end_FLOATSUBSCRIPT) in the MLT, From SABER and SCIAMACHY Atomic Oxygen Climatologies’, Journal of Geophysical Research (Atmospheres) 126(23), e2021JD035343.
  • Tian et al. (2005) Tian, F., Toon, O. B., Pavlov, A. A. & De Sterck, H. (2005), ‘Transonic Hydrodynamic Escape of Hydrogen from Extrasolar Planetary Atmospheres’, ApJ 621(2), 1049–1060.
  • Vidal-Madjar et al. (2013) Vidal-Madjar, A., Huitson, C. M., Bourrier, V., Désert, J. M., Ballester, G., Lecavelier des Etangs, A., Sing, D. K., Ehrenreich, D., Ferlet, R., Hébrard, G. & McConnell, J. C. (2013), ‘Magnesium in the atmosphere of the planet HD 209458 b: observations of the thermosphere-exosphere transition region’, A&A 560, A54.
  • Watson et al. (1981) Watson, A. J., Donahue, T. M. & Walker, J. C. G. (1981), ‘The dynamics of a rapidly escaping atmosphere - Applications to the evolution of earth and Venus’, Icarus 48, 150–166.
  • Zahnle et al. (2019) Zahnle, K. J., Gacesa, M. & Catling, D. C. (2019), ‘Strange messenger: A new history of hydrogen on Earth, as told by Xenon’, Geochim. Cosmochim. Acta 244, 56–85.
  • Zahnle & Kasting (1986) Zahnle, K. J. & Kasting, J. F. (1986), ‘Mass fractionation during transonic escape and implications for loss of water from Mars and Venus’, Icarus 68(3), 462–480.
  • Zahnle et al. (1990) Zahnle, K., Kasting, J. F. & Pollack, J. B. (1990), ‘Mass fractionation of noble gases in diffusion-limited hydrodynamic hydrogen escape’, Icarus 84(2), 502–527.