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

ATMOSPHERE-INTERIOR EXCHANGE ON HOT, ROCKY EXOPLANETS

, , , and

Published 2016 September 6 © 2016. The American Astronomical Society. All rights reserved.
, , Citation Edwin S. Kite et al 2016 ApJ 828 80 DOI 10.3847/0004-637X/828/2/80

0004-637X/828/2/80

ABSTRACT

We provide estimates of atmospheric pressure and surface composition on short-period, rocky exoplanets with dayside magma pools and silicate-vapor atmospheres. Atmospheric pressure tends toward vapor-pressure equilibrium with surface magma, and magma-surface composition is set by the competing effects of fractional vaporization and surface-interior exchange. We use basic models to show how surface-interior exchange is controlled by the planet's temperature, mass, and initial composition. We assume that mantle rock undergoes bulk melting to form the magma pool, and that winds flow radially away from the substellar point. With these assumptions, we find that: (1) atmosphere-interior exchange is fast when the planet's bulk-silicate FeO concentration is low, and slow when the planet's bulk-silicate FeO concentration is high; (2) magma pools are compositionally well mixed for substellar temperatures ≲2400 K, but compositionally variegated and rapidly variable for substellar temperatures ≳2400 K; (3) currents within the magma pool tend to cool the top of the solid mantle ("tectonic refrigeration"); (4) contrary to earlier work, many magma planets have time-variable surface compositions.

Export citation and abstract BibTeX RIS

1. INTRODUCTION

Over 100 exoplanets have masses or radii in the rocky-planet range and substellar equilibrium temperatures hot enough to melt peridotite rock.5 These molten surfaces are tantalizing because they are relatively easy to detect and characterize (Rouan et al. 2011; Demory 2014; Samuel et al. 2014; Sheets & Deming 2014). What sets molten-surface composition?

The melt-coated dayside is exposed to intense irradiation, sufficient to remove H2 (Valencia et al. 2010; Owen and Jackson 2012; Owen & Wu 2013; Lopez & Fortney 2014) and to maintain a thin silicate atmosphere (Figure 1). In the case where all atmophile elements (e.g., C, H) have been removed, thin-silicate-atmosphere composition is set by silicate-surface composition. Unlike the easier-to-observe nebular-accreted atmospheres and outgassed secondary-volatile atmospheres, thin exoplanetary silicate atmospheres are only now coming into view (Forget & Leconte 2014). The most volatile rock-forming constituents of the melt (e.g., Na, K, Fe) preferentially partition into the atmosphere—fractional vaporization. Borne by winds, these volatiles make a one-way trip to the permanent nightside (Callegari & Rodríguez 2013; Makarov & Efroimsky 2013), or are lost to space (Figure 1)—trans-atmospheric distillation. If trans-atmospheric distillation is faster than mass recycling between the melt pool and the solid interior, then surface composition will differ from bulk-planet silicate composition. However, if mass recycling between the melt pool and solid interior is fast, then surface composition will be repeatedly reset toward bulk-planet silicate composition (Figure 1).

Figure 1.

Figure 1. Processes shaping the surface composition of a hot rocky exoplanet. A magma pool of depth dp and mean composition Xp (surface composition Xs) overlies a solid mantle of composition X0. Pool depth (dp) is shown greatly exaggerated. Tlo corresponds to the temperature at the melt pool's edge. θp corresponds to the angular radius of the melt pool.

Standard image High-resolution image

In the first (compositionally evolved) case with relatively slow recycling, loss of volatiles (Na, K, Fe, etc.) enriches the residue in Ca and Al, forming a refractory lag (Léger et al. 2011; Figure 2). The lag protects the vulnerable volatile-rich interior, as on a comet. After lag formation, atmospheric pressure is usually ≲O (1) Pa (Figure 2). In the second (compositionally buffered) case, Na, K, and Fe are replenished by surface-interior exchange, the exosphere fills with Na and K, and surface compositional evolution is very slow because it is buffered by the massive reservoir of the planet's interior. Surface composition will affect the atmospheric abundances of Na and K (Nikolov et al. 2014; Heng et al. 2015; Wyttenbach et al. 2015), the properties of dust plumes streaming from disintegrating rocky planets (van Lieshout et al. 2014; Budaj et al. 2015; Schlawin et al. 2016), phase curves (Demory et al. 2013, 2016a), the potential for time-variability (Demory et al. 2016b), and reflectance/emission spectra (Hu et al. 2012; Samuel et al. 2014; Ito et al. 2015).

Figure 2.

Figure 2. Trans-atmospheric distillation. Top panel: sketch of fractional vaporization. Bottom panel: decay of equilibrium vapor pressure during fractional vaporization of an initial composition corresponding to Bulk Silicate Earth (Table 2). Gray solid lines show fractional vaporization at different temperatures. Vertical dashed lines separate regions where different oxides control the density evolution of the surface. Horizontal dotted lines show the pressure below which UV-driven escape is less efficient: optical depth = 1, assuming molar mass 30 g and photoabsorption cross-section of 10−22 m2 molecule−1 (Reilman & Manson 1979), for surface gravities of 1.5 m s−2 (lower line) and 15 m s−2 (upper line).

Standard image High-resolution image

To what extent does fractional vaporization drive surface composition? To answer this question, we quantify two key controls on magma-pool surface composition (Figure 1). First, because atmospheric mass flux scales with vapor pressure has a (super-)exponential dependence on temperature, winds are more important on the hottest planets than sluggish magma currents. Magma currents are paced by diffusion and relatively insensitive to pool temperature. This means that wind transport on the hottest planets permits compositionally variegated pool surfaces (Figure 3). However, winds are unimportant on relatively cool magma planets, and so cooler magma pools are well stirred (Figure 3). The buoyancy evolution of melt pools undergoing fractional vaporization is the crucial second control: depending on initial composition, fractional evaporation can promote stratification or it can promote overturn. Together, these two controls determine whether hot rocky exoplanet surfaces develop compositionally evolved surfaces, or are repeatedly reset to bulk-silicate composition.

Figure 3.

Figure 3. Magma-pool internal structure. (a) Notation for the pool and boundary layer: dp, pool depth; E, evaporation mass flux, v, wind speed; θ0, switch from evaporation to condensation; θp, pool edge; w, upwelling velocity; κT, magma thermal diffusivity; δT, thickness of thermal boundary layer (thermocline depth); vp, net speed of the current corresponding to the surface branch of the overturning circulation; κX, magma compositional diffusivity; δX, thickness of compositional boundary layer. Downwelling occurs near the pool edge. Colored contours in (a) correspond to temperatures within the boundary layer, ranging from high (red) to low (mid-blue). The thermal boundary layer is internally temperature-stratified near the substellar point. Before downwelling, the thermal boundary layer is internally mixed by small-scale convection. (b) The ocean-circulation-dominated limit (uniform pool surface composition). (c) The atmospheric-transport-dominated limit (variegated pool surface composition). Pool-base material can redissolve or drain into the solid mantle. In (b) and (c), red = more volatile than planet bulk-silicate composition, and gray = less volatile than initial planet bulk-silicate composition.

Standard image High-resolution image

Figures 79 sum up what we found. Section 2 describes a minimal model of evaporation (winds) and circulation (currents). Section 3 describes how the pool's composition evolves. We find two pathways that might permit a chemically evolved surface composition: formation of a buoyant boundary layer within the pool, or evolution of the whole-pool composition toward a buoyant lag (Figure 3). In Section 4.1, we discuss model assumptions and limitations. We discuss albedo feedbacks in Section 4.2, links to the solid mantle in Section 4.3, links to planet formation theory in Section 4.3, and planet disintegration in Section 4.5. In Section 4.4 we discuss links to the solid mantle, links to planet formation theory, and planet disintegration. We discuss links to observations in Section 5 and conclude in Section 6. Parameters and variables are listed in Table 1.

Table 1.  Selected Parameters and Variables

Parameter Description Value/Units Source/Rationale
cp Heat capacity, atmosphere 850 J kg−1 K−1
${c}_{p,l}$ Heat capacity, magma 103 J kg−1 K−1
R Gas constant 8.314 J K−1 mol−1
H Scale height, atmosphere 50 km
lv Latent heat of vaporization, magma × 106 J kg−1 (1)
TAS Temperature, antistellar hemisphere 50 K (2)
Tlo Temperature at rheological transition ("lock-up") 1673 K (3)
α Albedo (planet surface) 0.1
γ Evaporation coefficient (in Hertz–Knudsen equation) 0.2 (4)
δ ρl Density contrast (solid versus liquid) 250 kg m−3
ηm Viscosity, mantle 1018 m2 s−1 (5)
  Viscosity, magma ≲102 m2 s−1 (6)
κ Diffusivity, thermocline 10−6–10−5 m2 s−1 Appendix C
κS Mass diffusivity, sub-Tlo mantle 10−14 m2 s−1 (7)
κT Thermal diffusivity, magma × 10−7 m2 s−1 (8)
κX Mass diffusivity, magma 10−9–10−10 m2 s−1 Appendix A
μ Molar mass, atmosphere 34.15 g (9)
μl Molar mass, magma 100 g
ρl Density, magma 2500 kg m−3
ω Fractional drainage (e.g., diapir pinch-off fraction) of the magma pool 0.3
dp Depth of pool m
E Evaporation flux kg m2 s−1
f Coriolis parameter s−1
g' Reduced gravity m s−2
P, Pss Atmospheric pressure; pressure at substellar point K
r Planet radius m
$\overline{T}$, Ts, Tss Pool-average temperature; surface temperature; temperature at substellar point K
v Speed of wind m s−1
vp Speed of magma overturning circulation in the near-surface of the pool m s−1
w Speed of upwelling in melt pool m s−1
Xs, Xp, X0 Compositions of melt-pool surface; pool; and solid silicate interior
δT, δX Thicknesses of thermal boundary layer; of compositional boundary layer m
Ω Planet rotation rate s−1
θp, θ0 Angular radius of pool; angular radius of evaporation-condensation boundary deg
ρ0, ρδ Density of silicate interior; density of chemical boundary layer kg m−3
ρh Maximum density during fractionation kg m−3
ζ Angular width of the twilight zone deg  

References. (1) Öpik (1958), (2) Léger et al. (2011), they have "50–75 K", (3) Katz et al. (2003), (4) Tsuchiyama et al. (1999), Grossman et al. (2000), Alexander (2001), Richter et al. (2002), Lauretta et al. (2006), Fedkin et al. (2006), Richter et al. (2007, 2011), (5) Zahnle et al. (2015), Solomatov (2015), (6) Dingwell et al. (2004), Russell et al. (2003), Giordano et al. (2008), Zahnle et al. (2015), Solomatov (2015), (7) Brady & Cherniak (2010), Chakraborty (2010), Van Orman & Crispin (2010), (8) Ni et al. (2015), (9) Mean of μSiO and μMg, SiO and Mg dominate at intermediate stages of fractionation.

Download table as:  ASCIITypeset image

1.1. This Paper in Context

Our focus is the chemical evolution of the surface. To obtain physical insights into chemical evolution, we use a simple model and make idealizations. We use simple models of the winds and currents. We do not resolve the details of atmosphere-ocean coupling. Instead, we emphasize how the winds and currents interact with the chemical evolution of the surface.

Following Léger et al. (2011), we model a thin, hemispheric melt pool as the conduit between the solid interior and the atmosphere (Figure 1). We go beyond Léger et al. (2011) by considering horizontal convection in the magma pool, by calculating the winds driven by vapor-pressure gradients, and—most importantly—by considering the effects of fractional evaporation on residual-melt density. Because of these differences, we find that dynamic, compositionally primitive surfaces are likely. This is in contrast to Léger et al. (2011), who conclude that melt pools should have compositionally evolved CaO–Al2O3 surfaces. Our approach to winds is anticipated by Castan & Menou (2011), who consider winds in a pure-Na atmosphere on a hot rocky exoplanet and show that winds have little effect on surface temperature. We approximate the solid silicate interior of hot super-Earths as isentropic, consistent with mantle convection models (e.g., van Summeren et al. 2011) which show that solid-state convection can subdue large inter-hemispheric contrasts in interior temperature. To predict silicate-atmosphere compositions, we use the MAGMA code (Fegley & Cameron 1987; Schaefer & Fegley 2004, 2009, 2010; Schaefer et al. 2012; Figure 2). We assume negligible H2O in the silicates.

Because magma planets are at close orbital distance, they are—and will remain—intrinsically easier to detect and characterize than true Earth analogs. Already, phase curves, albedo constraints, and time variability have been reported (Rouan et al. 2011; Rappaport et al. 2012, 2014; Demory 2014; Dragomir et al. 2014; Sheets & Deming 2014; Sanchis-Ojeda et al. 2015; Vanderburg et al. 2015; Demory et al. 2016a, 2016b). Therefore, there is now a pressing need for self-consistent theory relating magma-planet geophysics to data.

2. SETTING THE SCENE: CURRENTS VERSUS WINDS

Magma-pool surface composition, Xs, is regulated by magma currents (Section 2.2), silicate-vapor winds (Section 2.3), and the development of compositionally distinct surface zones (Section 2.4). If currents transport much more mass than winds, then Xs will be the same as the pool-averaged composition Xp. However, Xs and Xp may differ if winds transport more mass than currents (Figure 3). Xp may be reset by drainage into the solid mantle (Section 2.5).

Liquid peridotite is not much more viscous than liquid water (Dingwell et al. 2004), and the almost-inviscid pool is subject to an insolation gradient (from center to edge). Circulation forced by a surface buoyancy gradient is remarkably slow (Sandström 1908, p. 6; Stommel 1961; Wunsch 2005; Hughes & Griffiths 2008). This slow stirring can allow time for Xs to diverge from Xp, despite the low fractionation rate implied by the thin air (Figure 2).

2.1. Melt Pools are Wide and Shallow

We assume that the melt pool's extent is set by a transition between fluid-like and solid-like behavior at a critical crystal fraction. Crystallizing magma acquires strength when increasing crystal fraction allows force to be transmitted through continuous crystal chains—"lock-up." Lock-up occurs at a melt fraction of ∼40% (Solomatov 2015). 40% melt fractions are reached at a lock-up temperature Tlo ≈ 1673 K (Katz et al. 2003; for a peridotitic composition; Appendix A). At Tlo, viscosity increases >1010-fold. Because of this large viscosity contrast, we treat material inside the pool as liquid, and refer to material colder than Tlo as "solid."

The pool angular radius, θp, is (in radiative equilibrium)

Equation (1)

where Tss = ${T}_{* }{(1-\alpha )}^{1/4}\sqrt{{r}_{\ast }/2a}({4}^{1/4})$ is the substellar temperature and the pool is centered on the substellar point.6 Here, a is the semimajor axis, T* is the star temperature, r* is the star radius, and we assume basalt-like albedo α = 0.1. (The optical albedo of liquid basalt is unknown, but we assume it to be similar to the albedo of solid basalt). Equation (1) applies if flow in the pool is sluggish, the atmosphere is optically thin, and atmospheric heat transport is negligible (Castan & Menou 2011; Léger et al. 2011)—all good assumptions (Sections 2.2, 2.3). Most planets with magma pools have θp > 60°.

Crystal fraction increases with depth. This is partly because pressure favors crystallization (Sleep 2007). Also, temperatures deep in the planet's solid mantle are smoothed-out by convective heat transport to the nightside, so that the mantle below the pool has lower entropy than the pool (van Summeren et al. 2011). These effects give a pool-depth estimate dp ∼ O(10) km (Appendix B). Pool depth is further reduced by within-pool circulation (Section 2.2). Typically the melt pool is a shallow, hemispheric, magma ocean.

2.2. Magma Currents Refrigerate the Pool Interior, but Are Too Slow to Perturb Surface Temperature

Magma flow is driven by gradients in T, X, or crystal fraction (Figure 3). An estimate of magma speed (vp) in the near-surface of the pool can be obtained by neglecting viscosity and inertia (this will be justified a posteriori). Then, balancing the pressure-gradient and rotational forces (geostrophic balance) yields

Equation (2)

where ∇P = ${g}^{\prime }{\rho }_{l}{\delta }_{T}/L$ = ${g}_{p}{\rm{\Delta }}{\rho }_{l}{\delta }_{T}/L$, where g' = ${g}_{{pl}}{\rm{\Delta }}{\rho }_{l}/\rho $ being the reduced gravity, gpl is the planet surface gravity, Δρl is the density contrast across the fluid interface at the bottom of the density boundary layer, f ≈ 2Ω sin(θp/2) is the Coriolis parameter evaluated at θp/2 (halfway between the substellar point and the latitudinal limits of the pool), here, Ω = 2π/p, with p the planetary period, δT is the thickness of the density boundary layer (where the density contrast can be due to T, X, or crystal fraction), and $L={\theta }_{p}R$ is the pool center-to-edge distance measured along the planet's surface (planet radius is R). This model ignores currents at the equator, which are not directly affected by Coriolis deflection.

Consider flow driven by ∇T. Substellar radiative equilibrium temperatures can exceed Tlo by >103 K. Magma expansivity is ∼10−4 K−1 (Ghiorso & Kress 2004), and so Δρl/ρl ∼ 10%. The pool thermal boundary layer grows diffusively for as long as magma takes to move from the substellar point to the pool edge. Therefore ${\delta }_{T}\approx \sqrt{{\kappa }_{T}L/{\rm{\Xi }}{v}_{p}}$, where κT is the diffusivity of heat (which we take to be the molecular thermal diffusivity, 5 × 10−7 m2 s−1; Ni et al. 2015), and Ξ ("ageostrophic flow fraction") is the scalar product of the magma-velocity unit vector and a unit vector that is directed from the center to edge of the pool. Now we can replace ∇P with ${g}^{\prime }{\rho }_{l}{L}^{-1}\sqrt{\kappa L/{\rm{\Xi }}{v}_{p}}$ and solve for the thermal-flow timescale τT:

Equation (3)

not counting any wind-driven circulation. A more sophisticated scaling (Vallis 2006)—including the variation of f with θ, which is appropriate for meridional flow—yields

Equation (4)

Equation (5)

where $\beta =(2\pi /{pr})\cos ({\theta }_{p}/2)$, i.e., evaluated at $\theta =\tfrac{1}{2}{\theta }_{p}$. κ = 10−4 m2 s−1 is possible with magma-wave breaking, but waves are likely small (Appendix C). Equation (5) is equal to Equation (3) with ${\rm{\Xi }}=L\beta /f$. Equations (3) and (5) yield the same results within 21% for Kepler's hot rocky exoplanets and Ξ ∼ 1; we use the results of Equation (5).

Magma current speed in the surface branch of the overturning circulation is vp ∼ L/τT ∼ 0.02 m s−1. The ratio of inertial to rotational forces (Rossby number Ro ≡ vp/Lf ≈ 10−3) and the ratio of viscous to rotational forces (Ekman number ${Ek}\equiv \mu /{\rm{\Omega }}{d}_{p}^{2}\ll 1$) both turn out to be small, and so our neglect of inertia and viscosity in Equation (2) is justified a posteriori.

Magma-pool heat transport Fo (column W m−2) is small. Fo is given by the product of vp and the θ-gradient in boundary-layer column thermal energy,

Equation (6)

where ${c}_{p,l}$ is melt heat capacity (103 J kg−1 K−1) and ρl is melt density (2500 kg m−3), giving

Equation (7)

104× less than insolation (Figure 9). Fo is also small for pools with gradients in crystal fraction and in X. This is because the Δρ for freezing, for compositional differences, and for temperature gradients are all O(10%), and combining all three effects (∼tripling g') only decreases the overturn time by a factor of two (Equation (5)). Including the enthalpy of crystallization in Equation (6) would not alter this conclusion. Magma oceans are much less efficient at evening-out temperatures than are liquid-water oceans. This is because, for hot planets, radiative re-equilibration (scaling as T4) defeats advection (which scales as T1; Showman et al. 2010). Because Fo is small, heat transport by melt-pool meridional overturning circulation cannot affect orbital phase curves (Esteves et al. 2015; Hu et al. 2015). Furthermore, an initially global surface magma pool cannot be sustained by melt-pool overturning circulation: instead, global surface magma pools beneath thin atmospheres will rapidly shrink to regional pools. Small F0 implies that ocean surface elevation smoothly tapers to near-zero at θp—so that levees of frozen, overspilt magma (Gelman et al. 2011) are not tall.

The thermal structure set up by horizontal convection (Figure 3; Rossby 1965; Hughes & Griffiths 2008) consists of a cool, deep layer at nearly constant potential temperature, fed by narrow downwellings near the pool edge, and topped by a thin thermal boundary layer of thickness δT:

Equation (8)

Equation (9)

Downwellings ventilate the subsurface interior of the pool with material that has been chilled near the pool edge to ≈Tlo. Because the material just below δT has T ≈ Tlo, dp will not be much greater than δT. We assume dp = 10 δT; dp = 3 δT is possible. Either option gives a pool depth that is >100× shallower than in the absence of overturning circulation (Equation (20); Léger et al. 2011).

2.3. Atmospheric Redistribution of Mass Within the Pool Outpaces Atmospheric Removal of Mass From the Pool

In this subsection, we show the following:

  • 1.  
    atmospheric pressure P adjusts to the pressure in equilibrium with local surface temperature,
  • 2.  
    atmospheric transport is $\propto \partial P/\partial \theta $;
  • 3.  
    atmospheric energy transport and atmospheric mass transport both fall rapidly near the pool edge.

Busy readers may skip the remainder of this subsection.

Silicate atmospheres for Tss < 3000 K (corresponding to all Kepler's planets) are thin enough to be in vapor-equilibrium (Heng & Kopparla 2012; Wordsworth 2015). Tenuous vapor-equilibrium atmospheres consisting of a single component equilibrate with local surface temperature on horizontal distances of ∼10 atmospheric scale heights H = RT/μg ≈ (8.3 J mol−1 K−1 × 2400 K)/(∼0.03 kg mol−1 × 15 m s−2) ≈ 50 km for hot rocky exoplanet silicate atmospheres, where R is the gas constant, μ is molar mass, and g is planet surface gravity (Ingersoll 1989; Castan & Menou 2011). The appropriateness of the single-component assumption is discussed in Section 4.1. μ is set to the mean of μSiO and μMg; SiO and Mg dominate at intermediate stages of fractionation (Schaefer & Fegley 2009). Because H ≪ planet radius (r), the approximation that pressure P adjusts to surface temperature Ts locally is well-justified. With local adjustment, pressure gradients are everywhere directed away from the substellar point. Because Ts is low on the nightside, nightside pressures are close to zero. Because the dayside atmosphere expands into near-vacuum on the nightside, wind speeds are v ∼ $\sqrt{{RT}/\mu }$ ∼ 1 km s−1 (sound speed), where T ($\sim \overline{T}$) is the hot-zone atmospheric temperature. At least for the analogous case of Io, surface friction is ineffective in braking the flow (Ingersoll et al. 1985). Inertial forces modestly exceed rotational forces (v/Lf ≈ 4 for p = 3 days), so that winds are also directed away from the substellar point. Winds transport mass from an evaporation region near the substellar point to cold traps. Because cold traps are outside the pool, the loss of mass from the pool requires atmospheric flow past θp. Therefore, the temperature at θp regulates loss from the pool (given the approximation of local adjustment)—this temperature is Tlo. At Tlo, the pressure Peq in equilibrium with silicate after 20 wt% fractional evaporation from an initial composition of Bulk-Silicate Earth is 1 × 10−2 Pa (∼5 × 10−4 kg m−2 for Super-Earths; Schaefer & Fegley 2009). The flux over the pool perimeter is thus mf = vPeq/g ∼ 0.5 kg s−1 m−1. For hemispheric pools, this corresponds to a column loss rate of mf/ρlr ∼ 5 × 10−4 m yr−1 averaged throughout the pool (ρl ≈ 2500 kg m−3, r ∼ 107 m). So the time for the pool to be depleted of the constituent that is dominant in the vapor is

Equation (10)

Equation (11)

where dr is the effective depth of mixing (i.e., the column depth of melt that can be obtained by depressurization and melting on timescale τd), fg is a geometric correction equal to 1 if the pool only samples material vertically beneath it, and fc is the mass-fraction of the component in the magma.

τd falls as T rises because Peq (and thus mass flux) increases super-exponentially with T. Peq(T) (for an initial composition of Bulk-Silicate Earth, 1400 K < T < 2800 K), is well fit by

Equation (12)

where k1 = {−34.7, −60.1, −60.2} and k2 = {−0.00112, −0.00122, −0.00101} for small (<0.01), intermediate (0.5), and large (>0.9) fractions of silicate vaporization, respectively, and Peq is in bars (Figure 2). Pool mean surface temperatures $\overline{T}$ are typically 500 K higher than pool-edge temperature Tlo. Column-integrated atmospheric mass flux scales with the gradient in the pressure and with the surface pressure. Because the surface pressure increases exponentially with local temperature, atmospheric transport near the center of the pool is much faster than atmospheric transport over the edge of the pool. For example, for CoRoT-7b, a basic calculation shows (Figure 6) that the trans-atmospheric transport over the evaporation-condensation boundary within the pool is 103 kg s−1 (m perimeter)−1, which is 5000× greater than transport over the edge of the pool. Therefore the pool can internally differentiate—develop compositional boundary layers—faster than pool material can be lost by atmospheric flow over the edge of the pool. This justifies our quasi-steady-state approximation for mass exchange within the pool.

In addition to the strong T dependence, τd is sensitive to Xs. As fractional evaporation progressively removes the more volatile components, Peq falls (Figure 2).

τd is proportional to dr. dr can vary from less than pool thermocline depth δT (10 m) up to planet radius r (107 m). Although as little as 108 years are required to remove Na from an Earthlike planet (fg = 2; dr ≈ 107 m; fc ∼ 0.0035, Table 2), 4 × 1010 years are required for full planet evaporation (fg = 2; dr ≈ 107 m; fc ∼ 1).

Table 2.  Compositions Investigated (wt%)

  Bulk-silicate Mercury Bulk-silicate Earth Bulk-silicate Mars Coreless Exoplanet
SiO2 47.10 45.97 45.0 28.8
MgO 33.70 36.66 30.6 18.7
Al2O3 6.41 4.77 3.06 2.0
TiO2 0.33 0.18 0.14
FeO(T)a 3.75 8.24 18.15 48.7
CaO 5.25 3.78 2.48 1.8
Na2O 0.08 0.35 0.051
K2O 0.02 0.04 0.03
Ca:Feb 1.3 0.42 0.13 0.04
Source (1) (2) (3) (4)

Notes.

a[FeO(T)] = [FeO], except for Coreless Exoplanet, for which [FeO(T)] = [Fe2O3]. bWeight ratio of Ca to Fe, decreasing with semimajor axis in the solar system. Although [FeO]'s effect on stratification can be offset by high [CaO], Ca is refractory (Grossman & Larimer 1974) and so [FeO] and [CaO] are anti-correlated in rocky planets.

References. (1) Morgan & Anders (1980), Zolotov et al. (2013), (2) Schaefer & Fegley (2009), (3) Dreibus & Wanke (1985), (4) Elkins-Tanton & Seager (2008a).

Download table as:  ASCIITypeset image

Atmospheric heat flux (W m−2) is given by

Equation (13)

Equation (14)

where lv = 6 × 106 J kg−1 is magma's vaporization enthalpy (neglecting sensible heat transport and wind kinetic energy, which are both <10% of lv). Fa equals absorbed insolation at 3500 K. For T ≪ 3500 K, Fa is small compared to insolation (Castan & Menou 2011).

Loss of mass to cold traps is supplemented by escape to space. Atmospheric escape is 108 kg s−1 for a planet around a 1 Gyr old Sun-like star—if escape is EUV-flux-limited and 100% efficient (Valencia et al. 2010). Efficiencies for rocky planets are poorly constrained (Murray-Clay et al. 2009; Ehrenreich et al. 2015; Owen & Alvarez 2015; Tian 2015). Even at 100% efficiency, escape to space corresponds to a hemisphere-averaged vertical breeze of 10 m s−1 (4 × 10−3 m yr−1 of melt-column ablation); this does not greatly alter the conclusion that P ≈ Peq, and so liquid remains stable (Ozawa & Nagahara 1997). Hydrostatic Roche-lobe overflow is minor for Super-Earths (Rappaport et al. 2013). Summing escape to-space and loss to cold traps, Earth-sized planets cannot be wholly processed in the age of the universe, but smaller planets can disintegrate (Section 4.4).

2.4. Competition between Evaporation and Circulation Affects Magma-pool Surface Composition

Surface composition for melt pools is set by competition between the pool's overturning circulation (which refreshes the surface on timescale τT, Section 2.2), and flow in the atmosphere, which produces a chemically differentiated boundary layer on a timescale τX. Consider a pool with two chemical components, volatile A and refractory Z. "A" evaporates from the surface preferentially near the substellar point, and condenses on the surface beyond the evaporation-condensation boundary—the angular distance (θ0) where the sign of net evaporation changes. If ρA > ρZ , then the residual melt in the evaporation zone becomes buoyant, and the residual melt in the condensation zone will sink into the pool interior. If ρA < ρZ, then the residual melt in the condensation zone becomes buoyant, and the residual melt in the evaporation zone will be unstable to sinking. Therefore, regardless of whether A or Z is denser, the magma pool will (given time) acquire a variegated surface, with parts of the surface having evolved Xs and parts of the surface being close to the initial composition. (Here we assume that there is a large density difference, and that θp > θ0.) Diffusive exchange with the deeper layers of the pool will initially resupply/remove constituents faster than fractional evaporation E, and so the time to form a chemically variegated surface is set by the diffusion-ablation balance,

Equation (15)

where κX is molecular diffusivity and $\overline{{E}_{e}}$ is the mass-loss rate due to fractional evaporation. The first term in brackets in Equation (15) is the compositional boundary-layer thickness δX = ${\kappa }_{X}{\rho }_{l}/\overline{E}$, and the second term in brackets is the compositional boundary-layer processing speed. κX is 10−9–10−10 m2 s−1 for liquid silicates at hot magma-pool temperatures (Karki et al. 2010; de Koker & Stixrude 2011). The Lewis number (${\kappa }_{X}/{\kappa }_{T}\,$) is  <0.01. Increasing κX slows down chemical boundary-layer development—because chemical diffusion refreshes the surface with fresh material. By contrast, increasing κT speeds up horizontal convection—because upwelling via thermal diffusion is needed to close the circulation (Equation (5)).

When τX/τT ≪ 1, chemical differentiation between the surface boundary layer and the interior of the magma pond may occur (Figure 3(c)). If chemical fractionation occurs, then the atmosphere effectively samples the chemically fractionated skin layer δX. By contrast, when τX/τT ≫ 1, the atmosphere effectively samples a well-mixed pool (Figure 3(b)).

${\tau }_{X}/{\tau }_{T}\ll 1$ is neither necessary nor sufficient for a buoyant lag to form. If the chemical boundary-layer density ρδ exceeds the initial density ρ0, then the incompletely differentiated skin layer can founder and be replaced at the surface by material of the starting composition. Alternatively, if the chemical boundary-layer density ρδ is less than the initial density ρ0, the entire pool might evolve into a (well-mixed) lag that is compositionally buoyant with respect to the mantle, even if τX/τT ≫ 1.

2.5. Magma-pool Composition Cannot be Buffered by Molecular Diffusion, but Can be Reset by Drainage Into the Mantle

Pool depth stays steady during evaporation. This is because melt-back of the stratified solid mantle at the base of the pool keeps pace with evaporative ablation of the pool surface.7

Equation (16)

Melt-back dilutes the fractionating pool with unfractionated material, slowing compositional evolution. (We assume that any solid phases that form at the base of the pool by reactions between the liquid and the ascending solid rock, such as spinels, are swept up into the pool and remelt).

Using the atmospheric model (Appendix D) to compute E, we find that molecular diffusion within the solid mantle is too slow to delay whole-pool fractional evaporation:

Equation (17)

where κS (≲10−14 m2 s−1) is a molecular diffusivity in crystalline silicates (Brady & Cherniak 2010).8

Suppose that the pool's composition is uniform and in steady state, and the pool is much less massive than the time-integrated mass lost to trans-atmospheric transport. Then, mass balance requires that the material lost from the pool must have the same composition as the melt-back input to the pool—i.e., bulk-silicate composition (Wang et al. 1994, p. 348; Ozawa & Nagahara 2001; Richter 2004). The pool adjusts to a composition Xb (buffer) that satisfies this condition. Because of the wide range of volatilities for the component oxides (Figure 2), Xb is CaO/Al2O3-dominated, with tiny proportions of Na2O and K2O, and a low total atmospheric pressure (steady, uniform, evolved surface composition; Figure 2).

If $\rho ({X}_{b})\lt \rho ({X}_{0})$, then Xp = Xb is a steady state. However, if $\rho ({X}_{b})\gt \rho ({X}_{0})$, then the intrinsically dense material in the pool may be unstable to finite-amplitude perturbations in the pool-base. Xp can then be intermittently reset to X0 by pool drainage into the mantle.

The pool can drain by infiltrating, by diking, or by forming one or more approximately spherical blobs (diapirs). Diapirism is rate-limited by the time needed to concentrate a hemispheric shell into a spherical diapir (Honda et al. 1993; Reese & Solomatov 2006). In numerical simulations (Honda et al. 1993), diapirs form on a timescale

Equation (18)

Equation (19)

where ηm is the mantle viscosity (∼1018 Pa s; Zahnle et al. 2015), G is the gravitational constant, r is the planet radius, and ω ≈ 0.3 is the fraction of the hemispheric shell's volume that contributes to the diapir. To obtain Equation (19), we assume ρ(Xb)/ρ0 ≈ 1.1. In order to drain, the diapir density must exceed the solid density, which might occur through compositional evolution, or through partial freezing by ventilation by cool currents at the base of the pool. τf ≈ τκ (Equation (16)). Diapir volume will be ${V}_{d}=2\pi {\delta }_{p}\omega {r}^{2}$ giving diapir radius O(102) km. As the diapir sinks, a maximum of ${V}_{d}{\rho }_{m}\,{\rm{\Delta }}\rho \,G\,{M}_{{pl}}/r\sim {10}^{26}$ J of gravitational potential energy is converted to heat, enough to warm the planet's interior by 0.01 K per event—delayed differentiation. (Alternatively, melting of the adjacent mantle from viscous dissipation as the diapir sinks may entrain lighter fluid and halt the descent of the diapir). Although a single diapir is the most linearly-unstable mode (Ida et al. 1987, 1989), smaller diapirs (or dikes, or magma solitons) may be more realistic, reducing ω. As $\omega \to 0$, composition becomes steady. For $\omega \ne 0$, composition is unsteady.

3. HOW MANTLE COMPOSITION REGULATES SURFACE-INTERIOR EXCHANGE: DENSITY EVOLUTION AND ITS IMPLICATIONS

Evaporation-versus-circulation competition (Section 2) determines whether pool surface composition evolves in lock-step with whole-pool composition. In cooler magma pools, circulation defeats evaporation and homogenizes pool composition. However, hotter magma pools will have compositionally variegated surfaces. Compositionally evolved portions of the surface may spread to form buoyant shroud layers depending on initial composition. Dense basal layers will form, and may drain into the mantle.

Now, we use a melt-density model (Appendix A) to track the density of residual magma during fractional vaporization. Residual-magma density determines whether the fractionated reservoir is unstable to sinking (Section 3.1). To calculate evaporation rates, we use a 1D atmospheric model (Appendix D).

3.1. Initial Bulk-silicate Composition X0 Determines Magma-pool Stability

Rocky-planet destruction by evaporation has five steps: (1) loss of atmophiles (H, C, N, S, P, halogens, noble gases), oceans, and continents; (2) loss of Na+K, (3) loss of Fe+Mg+Si, (4) loss of the residual Al+Ca(+Ti); (5) loss of the metal core. We consider steps 2, 3, and 4, for a range of bulk-silicate (mantle + crust) compositions (Table 2). We do not consider carbide-rich planets (e.g., Wilson & Militzer 2014). The main density-determining components are SiO2, CaO, Al2O3, MgO, FeO(T), Na2O, and K2O (Figure 2). (FeO(T) can stand for FeO, Fe2O3, Fe3O4, or a mixture, depending on oxidation state). Cations are lost from the melt in the order Na > K > Fe > (Si, Mg) > Ca > Al.

Output from the MAGMA code for fractional evaporation of melt at 2000 K (Bulk-Silicate Earth composition) is shown in Figure 4. Initially the magma outgasses a Na/K rich mix which (upon condensation) has a low density. The residue density does not change much at this stage because Na2O and K2O are minor components of the melt. Next to outgas is an Fe-rich mix. Upon condensation, this material is denser than the magma underlying it; the Fe-rich material will sink to the base of the pool. Continuing beyond ∼22 wt% fractional evaporation, the density of the residual magma increases. This will lead to small-scale compositional convection within the evaporation-zone compositional boundary layer, but the boundary layer as a whole still has lower density than the initial density (ρ0). The still-buoyant boundary layer continues to evolve to a more-fractionated composition (the gas is now dominated by Mg, Si, and O). Beyond 81 wt% fractional vaporization (point A in Figure 4), the density of the now-well-mixed boundary layer exceeds ρ0. That is because the boundary layer is now CaO-rich (∼20% molar) and CaO is dense (Appendix A). We define ρh as the maximum density at >70 wt% fractionation. Beyond A, the boundary layer sinks into the underlying magma. Fractional evaporation beyond A is not possible.9 The surface reaches a dynamic steady state, with parts of the surface covered by fresh material, and parts of the surface covered by evolved material that is nearly ready to sink. Atmospheric pressure will not be much less than the pressure above an unfractionated melt. The relevant reservoir for long-term net pool compositional evolution is the entire silicate mantle (assuming dense material drains efficiently into the solid mantle).

Figure 4.

Figure 4. Fractional vaporization at 2000 K of an initial composition corresponding to Bulk-Silicate Earth. Upper panel: residual-magma compositional evolution. Na and K are lost rapidly. Lower panel: density evolution. Thin black solid curve corresponds to the density of residual magma, and thick gray dashed curve corresponds to the density-upon-condensation of the gas. ρ0 corresponds to unfractionated magma density. ρh corresponds to the maximum density at >70 wt% fractional vaporization. At point A, the surface boundary layer sinks into the interior.

Standard image High-resolution image

Fractional vaporization of an FeO-rich exoplanet (Elkins-Tanton & Seager 2008a) proceeds differently (Figure 5). At 0 wt% fractionation, the melt is ∼10% denser than the unfractionated melt of Bulk-Silicate Earth. The FeO-rich pool will initially develop a stably stratified layer due to loss of Fe. At 60 wt% fractionation, density starts to rise and small-scale convection develops within the boundary layer (Figure 5). At 95 wt% fractionation, the extent of convection reaches its maximum—but convection is still confined within the boundary layer. The boundary layer stays buoyant; re-equilibration with the deep interior is inhibited. The relevant reservoir depth for fractionation is δX or δT for τT ≫ τX, or the entire pool for τT ≪ τX. Because τd (δX), τd(δT) and τd(δp) are all ≪1 Gyr (Equation (11)), FeO-rich planets are vulnerable to surface-composition evolution. This leads to an extremely low-pressure atmosphere (Figure 2). Results for additional compositions, and additional details, are shown in Figures 1417.

Figure 5.

Figure 5. Fractional vaporization at 2000 K of an initial composition corresponding to the Coreless Exoplanet of Elkins-Tanton & Seager (2008a). Upper panel: residual-magma compositional evolution. Lower panel: density evolution. The thin black solid curve corresponds to the density of residual magma, and the thick gray dashed curve corresponds to the density-upon-condensation of the gas. ρ0 corresponds to unfractionated magma density. ρh corresponds to the maximum density at >70 wt% fractional vaporization. At point A, small-scale convection within the surface boundary layer is maximal, but the boundary layer as a whole remains buoyant.

Standard image High-resolution image

Vulnerability to stratification is proportional to (ρ0ρh). (ρ0ρh) is shown in Figure 7 for a range of pool $\overline{T}$ and mantle compositions (Table 2). High [FeO] makes stratification more likely because FeO is both dense and volatile. When [FeO] is high, ρ0 is high. For a high-[FeO] world, the near-complete loss of FeO before 70 wt% fractionation greatly decreases residual-magma density relative to ρ0. Therefore the Ca-driven "uptick" in density at high fractionation does not reach ρ0 (Figure 5). Therefore, (ρ0ρh) is high (Figure 7). If [FeO] is low, then stratification is impossible. If [FeO] is high, then stratification is possible.

To calculate the time τX required for compositional evolution (Figure 8), we need to know the mean evaporation rate ($\overline{E}$) within the pool. To find $\overline{E}$, we use a 1D atmospheric model (Appendix D), which largely follows Castan & Menou (2011) and Ingersoll (1989). Typical 1D model output is shown in Figure 6.

Figure 6.

Figure 6. Example of atmospheric-model output (for CoRoT-7). Top panel: thick black lines correspond to pressure (solid is uncorrected for evaporative flux, dashed is corrected for atmospheric flux). Bottom panel: light gray line corresponds to evaporation (negative for condensation). Thin solid line corresponds to air temperature (K). Thick dashed line corresponds to wind speed (m s−1). The magma pool extends from a region of moderately high pressure (75 Pa) to a region of very low pressure. Results are shown at 50 wt% fractionation, for an initial composition corresponding to Bulk-Silicate Earth. (See Appendix D for details).

Standard image High-resolution image

3.2. Summary of Model Output

By combining Equations (5)–(11) and (23)–(26), we can predict the surface composition and atmospheric pressure for each magma planet. Predictions for specific magma planets are shown in Table 3. Figures 79 show the expected steady-state outcomes.

  • 1.  
    If hot rocky exoplanets are FeO-rich, then pool compositions will be steady, uniform, and dominated by CaO and Al2O3. We refer to this state as a "buoyant shroud." Atmospheres will be ≲O(1) Pa (Figure 2). Venting of Na and K from the pool will be minor, and escape to space may be reduced.
  • 2.  
    If hot rocky exoplanets are FeO-poor, then pool composition will be time-variable, and probably variegated. Averaged over timescales >τf, atmospheres will be thick. Escape to space will not be limited by the supply of atmosphere (Figure 2).
  • 3.  
    τT increases as $\overline{T}$ increases because hotter pools are wider—with increased L(θp) and f(θp).

Table 3.  Predictions of Atmospheric Pressure and Surface Composition for Selected Magma Planets

Planet p (day) Tss (K) Melt-pool Radius (°) Mean Pool Temperature $\overline{T}$ (K) Pool Depth (δp × 10) (m) (θp > θ0)? Pool-overturn Timescale (s) Atmosphere Timescale (s) Pss at 95 wt% Fractionation (Pa) Pss at 20 wt% Fractionation (Pa)
Not known to be disintegrating:        
CoRoT-7b 0.85 2425 77° 2146 100 Y 108 107 100 102
55 Cnc e 0.74 2671 88° 2263 200 Y 109 106 101 103
Kepler-36b 13.8 1476 n.a. n.a. n.a. n.a. n.a. <10−8 10−5
Kepler-93b 4.73 1569 n.a. n.a. n.a. n.a. n.a <10−7 10−4
Kepler-78ba 0.36 3056 102° 2423 200 Y 109 102 102 104
Kepler-10b 0.84 3007 96° 2464 200 Y 109 103 102 104
Disintegrating:        
KIC 12557548b 0.65 2096 66° 1919 100 Y 108 107 10−2 100
K2-22b 0.62 2135 68° 1945 100 Y 109 106 10−2 100
KOI-2700b 0.91 1780 39° 1730 60 Y 108 1011 10−5 10−2

Notes. Dominant timescales are underlined. Timescales calculated for 20 wt% fractionation. An initial composition corresponding to that of Bulk-Silicate Earth is assumed. Pressures <O (1) Pa are too low for disintegration (Perez-Becker & Chiang 2013).

aThe assumptions made in obtaining the atmosphere and ocean timescales are only marginally satisfactory for planets as hot as Kepler-78; see Section 3.3.

Download table as:  ASCIITypeset image

3.3. Stratification Should Occur for Mantles Formed From Oxidized Planetesimals

[FeO(T)] measures the extent to which a planet's building-blocks have been oxidized. (For this paper, [FeO(T)] = [FeO] for the Bulk-Silicate Earth composition, the Bulk-Silicate Mercury composition, and the Bulk-Silicate Mars composition, and [FeO(T)] = [Fe2O3] for the Coreless Exoplanet composition.) The most likely oxidant is water.

FeO(T) is not expected from condensation-from-vapor of an solar-composition protoplanetary disk: fast condensation of solar-composition gas at the snowline (145–170 K) forms large quantities of iron (Fe + FeS), Mg-silicates, and H2O, but FeO(T) in silicates is not expected because of kinetic effects (Krot et al. 2000; Lewis 2004; Podolak & Zucker 2004; Moynier & Fegley 2015). However, FeOx production occurs readily via metal-water reactions (Lange & Ahrens 1984; Dreibus & Wanke 1987; Rosenberg et al. 2001),

on planetesimals and larger bodies, as well as via water-rock reactions, such as (Früh-Green et al. 2004; McCollom & Bach 2009; Klein et al. 2013)

and by condensation of silicates from a water-enriched vapor, for example, a plume produced by impact into a water-rich target planetesimal (Fedkin & Grossman 2016). FeO(T) produced by Reactions 1 and 2 remains in the silicate mantle. Because of accretion energy, growing planets get hotter. Above ∼900 K, Reaction 2 is thermodynamically unfavorable. However, iron oxidation can continue if accreted material includes both Fe-metal and H2O (e.g., via Reaction 1; Kuwahara & Sugita 2015). [FeO(T)] production is also favored by partitioning of metallic Fe into the core and by H-escape (Wade & Wood 2005; Frost et al. 2008; Zahnle et al. 2013).

Because Reactions 1 and 2 involve water, they occur more readily where water content is enhanced: e.g., beyond the snowline or in objects that accrete material from beyond the snowline. Consistent with this, rocky objects that formed further out in the solar nebula have more mantle FeO(T) (Figure 7; Rubie et al. 2011, 2015). With 18 wt% FeO in silicates, Mars is (just) unstable to stratification under fractional vaporization (although only for relatively cool magma; Figure 7). FeO-rich silicate compositions (>25 wt%) are also obtained for the parent body of the CI-class meteorites, the parent body of the CM-class meteorites, and some angrite-class meteorites (Anders & Grevesse 1989; Lewis 2004; Keil 2012). Planets with even more FeO are theoretically reasonable, and Ceres may be a solar system example (McCord & Sotin 2005; Elkins-Tanton & Seager 2008a; Ciesla et al. 2015; Rubie et al. 2015). [FeO] can affect hot rocky-planet atmospheric thickness and time variability (Figures 2, 7). Because atmospheric thickness and magma-planet variability are potentially observable, this suggests a route to constrain hot rocky-planet composition. Such routes are valuable because mass and radius measurements only weakly constrain the composition of exoplanet silicates (Rogers & Seager 2010; Swift et al. 2012; Gong & Zhou 2012; Fogtmann-Schulz et al. 2014; Dorn et al. 2015).

Figure 7.

Figure 7. Magma-planet phase diagram. Blue and gray lines divide the phase diagram into quadrants. Blue lines correspond to stratification index (ρ0ρh) contoured at +50 kg m−3 (top dashed blue line), 0 kg m−3 (thick solid blue line), and −50 kg m−3 (bottom dashed blue line), using the Ghiorso & Kress (2004) equation of state. Planets below the line are unlikely to have CaO/Al2O3-dominated surfaces, planets above the line are likely to have CaO/Al2O3-dominated surfaces. Gray lines (near vertical) correspond to the ocean-dominance index τX/τT, contoured at 10 (left thin solid line), 1 (thick line), and 0.1 (right thin solid line), for 50 wt% vaporization. The lower left quadrant corresponds to ocean-dominated planets with uniform but time-variable surfaces driven by thermal overturn. The lower right quadrant corresponds to atmosphere-dominated planets with time-variable and compositionally variegated surfaces driven by evaporative overturn. The upper two quadrants correspond to planets with stable, stratified, CaO–Al2O3-dominated surfaces (compositionally evolved). Calculations fix p (to 0.84 days), r (to 1.47 R), and g (to 1.9 g) appropriate for Kepler-10b. Named compositions (red) are from Table 2. Additional results are shown in Figure 18.

Standard image High-resolution image

Furthermore, because water is less available inside the snowline, [FeO] probes the hot rocky planet's birth location relative to the snowline (and thus migration distance). Assuming a radial temperature gradient similar to the solar nebula, and further assuming that only a negligible fraction of Fe-silicates from the birth molecular cloud (Jones 1990; Min et al. 2007) survive to be incorporated into hot rocky exoplanets, evidence for mantle [FeO] in hot rocky exoplanets is evidence against in situ accretion (Chiang & Laughlin 2013). Evidence for mantle [FeO] in hot rocky exoplanets might be used to test inside-out planet formation (Chatterjee & Tan 2014). This is because [FeO] formation in the Solar System occurred on >km-sized objects, but in the Chatterjee & Tan (2014) model, >km-sized objects are assembled close to the star from dehydrated boulders. If hot rocky exoplanets formed through migration of objects of planetesimal size or larger from beyond the snowline (Cossou et al. 2014; Raymond et al. 2014, p. 595), then we expect high [FeO] on hot rocky exoplanets.

4. DISCUSSION

4.1. Overview of Approximations and Model Limitations

Our model omits or simplifies geologic processes that are not well understood even for the rocks of Earth. For example, two-phase (mush) effects, such as magma solitons, filter-pressing, and fingering instabilities (Scott & Stevenson 1984; Katz et al. 2006; Solomatov 2015), are not included in our model. Melt-residue separation at modest temperatures (low melt fractions) will yield Ca-rich, Al-rich melts. Ca/Al-rich melts have low Tlo and tend to favor overturning during fractional vaporization. This overturning-promoting effect is less strong for the high melt fractions (hot rocky exoplanets) that are considered here. However, even at high melt fraction, olivine seperation from melt might still be important, and the effect of this process on surface composition could be a target for future work (Asimow et al. 2001; Suckale et al. 2012).

Bulk-rock compositions are assumed to be similar to Solar System silicates, consistent with rocky-planet densities, stellar spectra, and white-dwarf data (Lodders et al. 2009; Jura & Young 2014; Adibeykan et al. 2015; Dressing et al. 2015; Gaidos 2015; Thiabaud et al. 2015). Some accretion simulations predict strongly varying Mg/Si (Carter-Bond et al. 2012; Carter et al. 2015). Future work might compute (ρ0ρh) for a broad range of silicate composition (Figure 7).

We assume a well-stirred mantle. Well-stirred mantles are predicted for large, hot planets by simple theories. Simple theories are undermined by 142Nd and 182W anomalies, which show that the Early Earth's mantle was not well-stirred (Debaille et al. 2013; Carlson et al. 2014; Rizo et al. 2016). This is not understood.

We do not discuss loss of atmophiles (i.e., primary and outgassed atmospheres), oceans and crust (Lupu et al. 2014). These steps should complete in <1 Ga even if the efficiency of EUV-driven atmospheric escape is low, although CO2 might resist ablation through 4.3 μm-band cooling (Tian et al. 2009; Tian 2015). Thick remnant atmospheres are not ruled out by density data (e.g., Ballard et al. 2014). Figure 19 shows results of a sensitivity test for crustal evaporation.

We also do not pursue the question of what happens to the magma-pool circulation on planets where the magma-pool circulation is dominated by the mass loading of atmospheric condensates (Tokano & Lorenz 2016). Such mass loading could throw the magma-pool circulation into reverse, but only on strongly atmosphere-dominated planets where the magma-pool circulation is much less efficient at transporting heat than is the atmosphere.

The atmospheric model has several limitations. (1) We assume that the temperature of evaporating gases is equal to the substellar temperature. (2) The model assumes a single atmospheric species. Although a single species typically dominates the silicate atmosphere (Schaefer & Fegley 2009), real atmospheres have multiple species. This is not a big problem in the evaporation zone (where the total saturation vapor pressure exceeds the total pressure of the overlying atmosphere). However, in the condensation zone, less-refractory atmospheric constituents form a diffusion barrier to condensation, and so condensation-zone atmospheric pressures will be higher in reality than in our model. This could increase P at θp. (3) Feedbacks between Xs and P(θ) during surface compositional evolution are neglected. (4) Magnetic effects (Batygin et al. 2013; Rauscher & Menou 2013; Koskinen et al. 2014; Lopez-Morales et al. 2011) are not included. Thermal ionization reaches ∼10−3 (fractional) at 3100 K for Bulk-Silicate Earth. (5) We find the initial winds, not the more complex wind pattern that may subsequently develop after compositional variegation develops. (6) Our model lacks shocks (Heng 2012).

Io's atmosphere is the best solar system analog to magma-planet atmospheres. Over most of Io's dayside, the approach of Ingersoll (1989) matches Io-atmosphere data (Jessup et al. 2004; Walker et al. 2012). Walker et al. (2010) find (in a sophisticated model that includes both sublimation and volcanic loading) that Io's atmosphere's evaporation zone extends 45°–105° from the substellar point, which is more than predicted by Ingersoll (1989).

Crystallization is inevitable near the edge of a compositionally evolved pool because Tlo is only just above the CaO–Al2O3 solidus (Berman 1983; Mills et al. 2014). If crystals are small enough to remain entrained in the melt, then evolved fluid near the margin of the pool will increase in density and sink (increased buoyancy forcing in Equation (5)). If crystallization increases viscosity, then the drainage of dense, evolved material into the solid mantle will slow, increasing the likelihood that the surface will have time to evolve into a CaO–Al2O3-rich composition. Lag formation is hard to avoid if the entire melt pond freezes.

We do not consider light scattering by rock clouds (Juhász et al. 2010). Cloud grain-size depends on P (and thus Ts), and so feedback involving silicate-dust clouds may modulate magma-pool activity (Rappaport et al. 2012).

Thermo-chemical boundary layers can have a complex substructure, which we simplify. Thermal stirring will tend to mix the boundary layer in the small-scale convection zone θ > $\theta ({T}_{s}=\overline{T})$, inhibiting compositional boundary-layer development. Therefore, the development of a buoyant layer segregation is most likely if the compositional boundary layer is developed before $\theta \gt \theta ({T}_{s}=\overline{T})$. This effect reduces the critical τX/τT by a factor of ∼2 (within the gray bars in Figures 8 and 9).

Figure 8.

Figure 8. Pool-overturn timescales vs. surface-fractionation timescales. Relatively cool magma pools are "ocean-dominated"—stirred by currents more rapidly than boundary-layer chemical segregation can occur (Figure 3(b)). Color of dots corresponds to pool $\bar{T}$ (K). Hotter magma pools are "atmosphere-dominated," with compositionally variegated surfaces (Figure 3(c)). The black line corresponds to chemical fractionation time (τX) equal to pool-overturn time (τT). The gray lines show 10× uncertainty. τX decreases rapidly as planets get hotter. The slope of the shallow branch (small magma pools) is set by τT's control of θp; for large magma pools, θp ∼ 90°, and the slope of the steep branch is set by the effect of p on rotational forces (Equation (5)). Selected planets are labeled; labels without prefixes are Kepler planet numbers. Bold highlights planets with measured densities. We exclude the putative planets orbiting KIC 05807616 (Krzesinski 2015). Results assume fractional vaporization of 20% of fractionating volume. Disintegrating rocky planets are assigned r = 0.3 r, M ≈ 0.01 M.

Standard image High-resolution image
Figure 9.

Figure 9. Atmospheric latent-heat transport compared to magma-current heat transport and to insolation. Color of dots corresponds to pool $\overline{T}$ (K). Contours show the ratio of the total advected flux (currents + winds), divided by insolation for an "average" hot rocky exoplanet. Results assume fractional vaporization of 20% of fractionating volume. The black line corresponds to Fa = Fo. This is not the same as the black line in Figure 8 and does not separate the same populations of planets. The gray lines show 10× uncertainty.

Standard image High-resolution image

We do not consider how condensed volatiles might return to the pool by gravity-current flow of viscous condensates from the permanent nightside back into the light (Leconte et al. 2013). For example, the Na from Earth corresponds to a hemispheric sheet of thickness 80 km, thick enough to flow back onto the light-side (Leconte et al. 2013). Sublimation of this Na flow might allow an Na-dominated (and thus much thicker) atmosphere to persist even after Na has been removed from the silicate mantle. Analogous flow may occur for S, which is the dominant volatile on Io's surface and makes up 250 ppm of Earth's mantle (McDonough & Sun 1995). Escape to-space is necessary to finally remove these materials from the planet.

For Ts > 3500 K, atmospheric heat transport melts the nightside, and the magma pool covers the entire surface—a magma sea. This limit is reached for accreting planets (Lupu et al. 2014; Hamano et al. 2015; Kuwahara & Sugita 2015), planets around post-main-sequence stars, and the hard-to-observe rock surfaces of planets with thick volatile envelopes (e.g., Howe & Burrows 2015; Chen & Rogers 2016; Owen & Morton 2016).

4.2. Albedo Feedbacks

Only a few relevant albedo measurements exist (Gryvnak & Burch 1965; Adams 1967; Nowack 2001; Petrov 2009), and so albedo-composition feedbacks are not considered in detail in this paper. High surface albedo would reduce atmospheric P and mass fluxes, and so magma currents are relatively more important on high-albedo worlds. Limited data (Zebger et al. 2005) indicate increasing UV/VIS reflectance with increasing CaO/SiO2 ratio in the system CaO–Fe2O3–SiO2. Taken at face value, these results suggest CaO-rich surfaces could explain data that have been interpreted as indicating high magma-planet albedos (e.g., Rouan et al. 2011; Demory 2014). Per our calculations, CaO-rich surfaces indicate FeO-rich initial compositions. In turn, this is a strike against in situ planet formation. This argument is necessarily speculative because of the paucity of relevant lab data. Motivated by the connection between magma-planet albedo and planet formation theory, gathering more laboratory data relevant to magma-planet albedos could make these arguments more rigorous.

4.3. Links to the Solid Mantle: Tectonic Refrigeration

On Earth, horizontal convection fills the deep sea with cool fluid (Lodders & Fegley 1998; Hughes & Griffiths 2008). Therefore, Earth's sea-floor temperature is ∼1°C, even though Earth's mean sea-surface temperature is 18°C. Similarly, magma-pool downwellings irrigate the magma-solid interface ("sea-floor") with cool magma. Cool-magma downwellings set a low and uniform top temperature boundary condition on the dayside mantle circulation. For a dayside-spanning shallow magma pool, the mean temperature at the upper boundary of a convecting-mantle simulation is the mean of the pool-edge temperature Tlo and the antistellar-hemisphere surface temperature TAS, i.e., ∼850 K (Figure 10). This is because Tlo is the characteristic temperature for the magma-solid interface, and magma floods the dayside, while the nightside remains very cold. Because mantle convection (in the absence of tidal heating) will transport much less heat than magma currents, this is a constant-T boundary condition. These low temperatures make 1:1 synchronous rotation more likely because pseudo-synchronous rotation requires a hot (> Tlo) interior (Makarov 2015). The "tectonic refrigeration" effect is large (Figure 10). Convection models (van Summeren et al. 2011; Foley et al. 2012; Lenardic & Crowley 2012; O'Rourke & Korenaga 2012; Tackley et al. 2013; Noack et al. 2014; Miyagoshi et al. 2014) could investigate the effects of a hemispherically uniform upper temperature boundary condition on mantle convection.

Figure 10.

Figure 10. Tectonic refrigeration. Upper symbols (squares) show the inter-hemispheric (dayside-nightside) temperature contrast at the top boundary of the solid mantle for hot rocky exoplanets. Black filled squares correspond to a circulating magma pool; blue open squares correspond to the unrealistic case of radiative equilibrium. The lower symbols (circles) show the global mean upper temperature boundary condition for mantle circulation on hot rocky exoplanets. Black filled squares correspond to a circulating magma pool; blue open squares correspond to the unrealistic case of radiative equilibrium. Global mean upper temperature boundary conditions for mantle convection on rocky planets with 1 day periods are not much greater than Venus' surface temperature (735 K).

Standard image High-resolution image

4.4. Stratification Should Be Common if Small-radius Exoplanets Have Outgassed Atmospheres

Planet density decreases with increasing planet radius in the range 1 < r < 4 (Wu & Lithwick 2013; Weiss & Marcy 2014). This density trend can be simply explained by models in which small planets consist of varying proportions of Earth-composition cores and H2-rich envelopes (Hadden & Lithwick 2014; Lissauer et al. 2014; Lopez & Fortney 2014; Dressing et al. 2015; Wolfgang et al. 2015). H2-rich envelopes might form by nebular accretion (Bodenheimer & Lissauer 2014; Jin et al. 2014; Lee et al. 2014; Inamdar & Schlichting 2015; Ogihara et al. 2015), or by outgassing of H2-rich material (Sleep et al. 2004; Elkins-Tanton & Seager 2008b; Sharp et al. 2013). The hottest rocky exoplanets either fail to develop a H2-rich envelope, or lose H2 in <1 Gyr (Chiang & Laughlin 2013; Chatterjee & Tan 2014; Lopez & Fortney 2014; Schlichting 2014; Ogihara et al. 2015). The upper limit on H2 outgassing via iron oxidation, corresponding to [FeO(T)] = 48.7 wt% (Table 2), is 1.7 wt% (Rogers et al. 2011). 1.7 wt% H2 is not very much greater than the amount required to explain the radii of low-mass Kepler planets (Lopez & Fortney 2014; Howe et al. 2014; Howe & Burrows 2015; Wolfgang & Lopez 2015). Therefore, if outgassing explains the observed radii of Kepler Super-Earths, then their rocky cores must contain abundant FeO(T). Upon H2 removal via photo-evaporation and impact erosion (Ehrenreich & Desert 2011; Schlichting et al. 2015), the rocky cores of the most-oxidized planets may develop stratified surfaces (Figure 4), forming a CaO–Al2O3-rich lag. This is speculative because the percentage of oxidation needed to match Kepler data depends on the molecular weight of the outgassed atmosphere, and that percentage is poorly constrained. The lag might be detectable through its effect on albedo (e.g., Demory 2014) and atmospheric composition.

4.5. Size Dependence and Planet Disintegration

The molten surfaces of disintegrating rocky planets are the single most likely source for time-variable dust plumes orbiting KIC 12557548, KOI-2700, K2-22, and WD 1145+017 (Rappaport et al. 2012, 2014; Sanchis-Ojeda et al. 2015; Vanderburg et al. 2015). Planet disintegration involves vapor pressures ≳O(1) Pa (Perez-Becker & Chiang 2013). Because evolved-surface-composition worlds usually have lower vapor pressures (Figure 2), they usually resist disintegration. Therefore, observed disintegrating planets require active surface-interior exchange, or entrainment of solids by escaping gas.

Smaller-radius worlds (e.g., KIC 12557548b, KOI-2700b, K2-22b) are more likely (relative to other worlds with the same temperature) to be atmosphere-dominated in our model. This is because the same gradient in atmospheric pressure with angular separation from the substellar point (∂P/∂θ) corresponds (for smaller-radius worlds) to a larger mass flux $\tfrac{\partial M}{\partial x}\propto {\rm{\nabla }}(P/{rg})\propto {r}^{-3}$ (assuming r M1/4; Seager et al. 2007). Larger atmospheric mass fluxes favor variegated, time-variable surfaces.

5. IMPLICATIONS FOR OBSERVATIONS

Current observatories can demonstrably constrain magma-planet time variability, and place limits on magma-planet albedo (Léger et al. 2011; Rouan et al. 2011; Rappaport et al. 2012, 2014; Demory 2014; Dragomir et al. 2014; Sheets & Deming 2014; Sanchis-Ojeda et al. 2015; Vanderburg et al. 2015). VLT/UVES data set upper limits on Ca i, Ca ii, and Na in the atmosphere of CoRoT-7b (Guenther et al. 2011). 55 Cnc e, which orbits a V = 6 star at 12 pc, is the most attractive candidate for magma-planet observations currently (Dragomir et al. 2014; Demory et al. 2016a; Ridden-Harper et al. 2016). TESS should increase the number of known nearby magma planets (Sullivan et al. 2015), given a Kepler-based estimate of 6 × 10−4 for magma-planet frequency (Ito et al. 2015). Here, we outline the implications of magma-pool composition for (in turn) observations of the exosphere + upper atmosphere, lower atmosphere/clouds, surface, and time domain. Our discussion draws on the work of Samuel et al. (2014) and Ito et al. (2015).

Magma-planet exospheres, upper atmospheres, and dust tails are mainly seen in transit. For the Na resonance doublet, assuming Hubble Space Telescope (HST)/STIS resolution, one can expect to see variations of ∼5 scale heights in a cloud-free atmosphere (giving 10–15 ppm signal for 55 Cnc e). With HST/STIS, detection of this signal would require ∼100 transits (scaling from Charbonneau et al. 2002). However, the signal is easily detectable in a single transit with HST/STIS (3000 ppm) if the Na (or K) absorption extends to fill the ∼3 Rp Hill sphere of 55 Cnc e. The James Webb Space Telescope (JWST)/NIRSPEC should improve over HST/STIS for magma-planet observations. (This improvement is hard to quantify because JWST/NIRSPEC error for bright stars may be dominated by systematics.) Na detection suggests that the planet's surface is little-fractionated. [FeO] might be constrained by dust-tail spectroscopy (Croll et al. 2014, 2015; Bochinski et al. 2015).

Magma-planet clouds and lower-atmospheres can be observed using secondary-eclipse spectroscopy. Ito et al. (2015) find that a photon-limited JWST-class telescope can detect SiO, Na and K in the atmosphere of 55 Cnc e with 10 hours of observations. Albedo measurements are possible at lower signal-to-noise ratio; they may correspond to the albedo of clouds, or the albedo of the surface (Rouan et al. 2011; Demory 2014; Dragomir et al. 2014).

Illustrating the imminent detectability of magma-surface properties, a phase curve for 55 Cnc e was reported while this paper was in review (Demory et al. 2016a). The phase curve shows a ∼40° eastward hotspot offset and an antistellar-hemisphere temperature of ∼1400 K. This high antistellar-hemisphere temperature requires heat transport (Section 2). Transport by magma currents is insufficient to explain the antistellar-hemisphere temperature (Figure 9), and so an atmosphere is indicated. 55 Cnc e's density is consistent with an Earthlike composition (Demory et al. 2016a), but the planet may nevertheless have retained an envelope of non-silicate volatiles. Near-future intruments (e.g., NIRSPEC on JSWT) can meaure CoRoT-7b's Bond albedo to ±0.03, test the synchronous-rotation assumption, and set limits on the atmospheric density (Samuel et al. 2014). This will help to distinguish the compositionally evolved from the compositionally primitive endmembers discussed in this paper (Figure 2).

Disintegrating magma planets have dust plumes that are time-variable (Figure 11). Short-timescale variations are plausibly linked to limit cycles involving the dust-plume optical depth and the surface magma temperature (Perez-Becker & Chiang 2013). However, longer-timescale variations (e.g., quiescent periods, Rappaport et al. 2012; van Werkhoven et al. 2014) must be rooted in a reservoir with a correspondingly longer relaxation time, such as the magma pool. For example, KOI-2700b's dust cloud faded from 2009 to 2013 (Rappaport et al. 2013). This may be connected to the magma-pool thermal-overturn timescale τT, which is usually a few years (Figures 8 and 11). On this timescale, a patch of pool surface that is anomalously dark (thus hot, with increased gas output) will drift to the pool edge and subside. High-[FeO] worlds have stable surfaces (Figure 7), and so any light-side transient implies low [FeO]. From Figure 8, period p > 1.5 day magma planets (p > 1 day for albedo 0.5) should lack surface-driven variability on timescales <106 s.

Figure 11.

Figure 11. Observed magma-planet time variability, compared to possible source mechanisms. Dust-plume-crossing timescale from Rappaport et al. (2012). Planet destruction timescale from Perez-Becker & Chiang (2013), assuming 2145 K isothermal Parker wind and vaporization of pure olivine. The left edge of the red bar is for destruction of a 100 km radius planet. Timescales for thermal-overturn (τT) and chemical boundary-layer relaxation (τX) from this paper.

Standard image High-resolution image

We conclude our discussion of observability with observational possibilities that are intriguing, but less likely. The induced power from a ring of metal (e.g., Na, Fe, etc.) condensed just beyond the terminator might affect magnetism. J2 constraints (Batygin et al. 2009) are unlikely to break degeneracies because of the trade-off between unknown mantle FeO and unknown planet Fe/Si (Taylor & McLennan 2009). Ito et al. (2015) state that planetary radial velocities are marginally detectable for Na and K for a Giant Magellan Telescope-class telescope (see also Ridden-Harper et al. 2016). Direct proof of fractional vaporization might involve measurement of isotopomers of SiO (Campbell et al. 1995).

6. CONCLUSIONS

Magma-pool overturning circulation and differentiation represents a new tectonic mode for rocky planets at temperatures too high for plate tectonics, stagnant-lid convection, or heat-pipes (Sleep 2000; Korenaga 2013; Moore & Webb 2013; Stamenković & Breuer 2014).

Surface-interior exchange on hot rocky exoplanets is driven by near-surface contrasts in melt density (and can shut down if the surface layer becomes stably buoyant). In turn, these density effects are regulated by two factors (Figure 7).

  • 1.  
    Relative vigor of evaporation and circulation. For "magma-pool-dominated" worlds (substellar temperature ≲2400 K), magma-pool overturning circulation outruns net evaporation. Pool surface composition tracks bulk-pool composition. For "atmosphere-dominated" worlds (substellar temperature ≳2400 K), pool overturning circulation is slow compared to atmospheric transport. Fractional evaporation drives pool surface composition away from the composition of the bulk of the pool.
  • 2.  
    Exposure of the planet's building-blocks to oxidants such as H2O. If the planetesimals that formed the planet grew ≳1 au from the star, water-rock reactions will lead to high Fe-oxide concentrations in the planet's silicate mantle. Close to the star, preferential evaporation of volatile and dense Fe favors stable stratification of the residual magma. This may allow a buoyant, stable lag to form—a compositionally evolved surface. However, if the planetesimals that formed the planet are more reduced, fractionally evaporated residual melt will sink. The concomitant resurfacing will repeatedly reset the surface composition to the planet-averaged silicate composition.

We thank Bruce Buffett, Michael Manga, Ruth Murray-Clay, Paul Asimow, Larry Grossman, Diana Dragomir, Valeri Makarov, Dorian Abbot, Michael Efroimsky, Ray Pierrehumbert, Eric Ford, Bethany Ehlmann, Brice-Olivier Demory, and especially Jacob Bean and Malte Jansen for discussions. We thank the anonymous reviewer and the editor, Steinn Sigurdsson. B.F. was supported by NSF grant AST-1412175. E.S.K. thanks the Astrophysics and Geosciences Departments at Princeton University for providing financial support and a convivial home while the ideas in this paper were marinating.

APPENDIX A: MATERIAL PROPERTIES

Density. We obtain molar volumes for melts in the early stages of fractionation from the silicate-melt equation of state of Ghiorso & Kress (2004) (Figure 12). This EoS is calibrated against a wider range of experiments, and so is preferred to, the EoS of Lange & Carmichael (1987). Mixture densities assume ideal mixing (Bottinga & Weill 1970); nonideal-mixing density errors are small for early-stage fractionation. Ghiorso & Kress (2004) include a fictitious "book-keeping" oxide, FeO1.3, which we ignore; the inclusion of this fictitious oxide would make FeO-rich silicates even denser, and therefore would not alter our conclusions. Both equations of state assume constant ∂V/∂T, consistent with simulations (Guillot & Sator 2007).

Figure 12.

Figure 12. Densities of pure oxides according to the fits of Lange & Carmichael (1987; thin lines) and Ghiorso & Kress (2004; thick lines). Error bars not shown. Vertical dashed line corresponds to Tlo.

Standard image High-resolution image

Late stages of fractionation produce CaO–MgO–Al2O3–SiO2 (CMAS) melts. The Ghiorso & Kress (2004) and Lange & Carmichael (1987) models are not calibrated for these melts (Figure 13). Instead, we use the CMAS equation of-state of Courtial & Dingwell (1999), including their SiO2–CaO nonideal-mixing term. To check, we compare to the data of Courtial & Dingwell (1995, 1999; both using the double-bob Archimedean method), and Aksay et al. (1979). Errors are small, as expected, except for densities measured with the X-ray radiography method (Aksay et al. 1979). A switch from four-oxygen to six-oxygen coordination of Al at high Al contents increases the density of Al-rich melts (Jakse et al. 2012).

Figure 13.

Figure 13. Comparison of models to data for the CaO–MgO–Al2O3–SiO2 system (appropriate for high degrees of fractional vaporization of silicate magmas). Crosses correspond to experiments using the double-bob Archimedean method. Triangles correspond to X-ray radiography experiments. Dots and diamonds correspond to modeled densities: gray dots for the model of Lange & Carmichael (1987); black dots for the model of Ghiorso & Kress (2004); and red diamonds for the model of Courtial & Dingwell (1999). Error bars on the experiments (not shown) are small compared to the mean data-model discrepancy.

Standard image High-resolution image

Molecular diffusivity. Diffusivity in the melt, κX, is set to $2.8\times {10}^{-7}\exp (-7.9\times {10}^{-4}{\mu }_{l}^{-1}/(8.314{\mu }_{l}^{-1}\overline{T}))$. μl =  100 g is the assumed molar mass in the liquid. This follows simulations of Mg self-diffusion in hydrous melts by Karki et al. (2010), but reduced by a factor of 2 to take account of anhydrous effects, per de Koker & Stixrude (2011). There is only a small dependence on component mass (Tsuchiyama et al. 1994). Using self-diffusivities is an approximation to the real, multicomponent diffusion. Self-diffusivities at 3000 K are slightly lower for O than for Mg, and ∼3 times lower for Si than for Mg.

Figure 14.

Figure 14. Fractional vaporization of a "coreless" exoplanet (Elkins-Tanton & Seager 2008a). The gray lines correspond to atmospheric pressure at temperatures of 1400, 1600, 2000, 2400, and 2800 K (in order of decreasing line thickness). The vertical dashed lines separate regions where different oxides (text) control the density evolution of the residual fluid. The horizontal dotted lines show the pressure below which UV-driven escape is less efficient (optical depth = 1 for surface gravities of 1.5 m s−2 (lower dotted line) and 15 m s−2 (upper dotted line), assuming molar mass 30 Da and photoabsorption cross-section of 10−22 m2 molecule−1; Reilman & Manson 1979).

Standard image High-resolution image
Figure 15.

Figure 15. Residual-magma density evolution for fractional vaporization of an initial composition corresponding to Bulk-Silicate Earth.

Standard image High-resolution image
Figure 16.

Figure 16. Residual-magma density evolution for fractional vaporization of an initial composition corresponding to a "coreless" exoplanet(Elkins-Tanton & Seager 2008a). The wiggles around 93 wt% fractionation are artifacts caused by the transition in our model from the Ghiorso & Kress (2004) equation of-state to the Courtial & Dingwell (1999) equation of-state.

Standard image High-resolution image
Figure 17.

Figure 17. Fractional vaporization (at 2000 K) of an initial composition corresponding to Bulk-Silicate Mars. Upper panel: residual-magma compositional evolution. Lower panel: density evolution. Thin black solid curve corresponds to the density of residual magma, and thick gray dashed curve corresponds to the density-upon-condensation of the gas. ρ0 corresponds to unfractionated magma density. ρh corresponds to the maximum density at >70 wt% fractional vaporization. At point A, the surface boundary layer is unstable to sinking. For Bulk-Silicate Mars initial composition, the density-upon-condensation curve is truncated at 0.97, due to numerical artifacts beyond 0.97.

Standard image High-resolution image
Figure 18.

Figure 18. Detailed magma-planet phase diagram (a simplified version of this figure is shown in Figure 7). Black and gray lines correspond to stratification index (ρ0ρh) contoured at 100 kg m−3 intervals. Negative values solid, positive values dashed. The zero line is highlighted by thick black and gray lines. Black contours correspond to the Ghiorso & Kress (2004) equation of-state, and gray contours correspond to the Lange & Carmichael (1987) equation of-state (Appendix A). Colored lines correspond to (red lines, left) τX/τT = 10, (green lines, middle) τX/τT = 1, and (blue lines, right) τX/τT = 0.1. Among the colored lines, the thin lines correspond to <1 wt% vaporization, the mediumthickness lines correspond to 50 wt% vaporization, and the thick lines correspond to 80 wt% vaporization. A choice of one colored ∼vertical line and one black-or-gray diagonal line divides the plot into quadrants. Then, the lower-left quadrant corresponds to ocean-dominated planets with uniform, but time-variable surfaces, driven by thermal overturn. The lower-right quadrant correspond to atmosphere-dominated planets with variable, variegated surfaces driven by evaporative overturn. The upper two quadrants correspond to planets with stable, stratified, CaO–Al2O3-dominated surfaces (compositionally evolved). Calculations assume orbital period p = 0.84 days, planet radius 1.47 R, gravity 1.9 g, appropriate for Kepler-10b.

Standard image High-resolution image
Figure 19.

Figure 19. Crust-vaporization sensitivity test. Residual-magma density evolution during fractional evaporation of the crust vs. mantle, for Earth (top panel) and Mars (lower panel). Thick lines correspond to fractional evaporation of crust (early stages in planet destruction) and thin lines correspond to fractional evaporation of mantle (later stages in planet destruction). The Earth oceanic crust composition is from Klein (2005) and the Mars crust composition is from Taylor & McLennan (2009); mantle compositions are shown in Table 2. Peak density during fractionation (ρh) moves to the left for fractional evaporation of crust due to the partitioning of Ca and Al into the melt during crust formation; initial densities (ρ0) are also lower.

Standard image High-resolution image

Diffusivity in the solid (κS) is ≲10−14 m2 s−1 for olivine at ∼1700 K (Brady & Cherniak 2010). For example, Chakraborty (2010) reports κS ∼ O(10−14) m2 s−1 at ∼1770 K for Mg diffusion in olivine (and κS < 10−17 m2 s−1 for Si and O). Similarly low values are reported for Mg diffusion in periclase (Van Orman & Crispin 2010), so that the true Lewis number may be even lower than used here.

Liquidus temperature and lock-up temperature. The Katz et al. (2003) crystallization-temperature parameterization interpolates between "0 GPa" experiments carried out on a rock that is representative of Earth's mantle: KLB-1 peridotite (Takahashi 1986). Katz et al. (2003) recommends a parameterization that yields 1673 K for Tlo.

APPENDIX B: DEPTH OF THE MELT POOL IN THE ABSENCE OF AN OVERTURNING CIRCULATION

Combining the effects of pressure-induced crystallization and the cooling of near-surface material by underlying cooler solid mantle, the bottom boundary of the melt pool is the depth where T(z) = Tlo(z). This depth is given by

Equation (20)

where $\overline{T}$ is the surface temperature of a well-stirred pond, $\overline{T}=({T}_{{ss}}\pi \,{\sin }^{2}\,{\theta }_{p})/2\pi {(1-\cos {\theta }_{p})}^{1/4}$, and dbl is the thickness of the mantle boundary layer. Because pressure favors crystallization, Tlo(z) increases with depth: $(\partial {T}_{{lo}}/\partial z)$ = 2 × 10−3 K m−1 for g = g (Sleep 2007; Solomatov 2015). $(\partial {T}_{{lo}}/\partial z)$ is proportional to gravity because the adiabatic temperature gradient and the slope of the curve of constant melt fraction are both proportional to pressure (Kite et al. 2009). For $\overline{T}=2400$ K, Tmantle = 1500 K, and dbl = 70 km, Equation (20) gives dp ∼ 50 km. In reality, lateral flow in the pool further suppresses dp. Setting dbl = 70 km implies mantle circulation speeds of the same order of magnitude as those calculated for Earth (Watters et al. 2009; Gelman et al. 2011; van Summeren et al. 2011).

APPENDIX C: STIRRING THE MELT POOL

The melt pool overturns more quickly if the effective (molecular or eddy) diffusivity is high. Wind-driven waves can stir the pool, especially if they break. Following Lorenz & Hayes (2012), transfer of energy J from the atmosphere to the wave-field scales as

Equation (21)

where Cd ≈ 2 × 10−3 is a surface exchange coefficient (Emanuel 1994) and ξ ∼ 0.01 is a correction factor for near-surface wind speed. Since ρa ≈ P/(gH) ≈ 10−6 kg m−3, therefore ∂J/∂x ∼ 10−7 J m−2 m−1, and so E = 1 J m−2 for an L = 107 m magma pool and a 1 Pa atmosphere. Lorenz & Hayes (2012) give $J=0.125\,{\rho }_{l}{{gh}}_{w}^{2}$, and (ignoring dissipation) this gives a maximum wave height hw = $0.015\sqrt{P}$ m. Therefore waves are unlikely to be big enough to mix the boundary layer, even if they break. An analogy is diapycnal viscosity at Earth's thermocline (10−5 m2 s−1; Munk & Wunsch 1998). Stirring will increase if insolation penetrates below the thermocline. We assume insolation is absorbed near the surface, above the thermocline. Increasing stirring has only modest effect on τT (100-fold increase in κT only increases ocean circulation speeds 4-fold). Faster stirring delays the development of a compositional boundary layer, which would allow thicker atmospheres to persist (Figure 2). However, because increasing ρa also increases E, this is unlikely to shift the boundary between ocean-dominated pools and atmosphere-dominated pools that is shown in Figure 8.

Tides can speed up ocean circulation (Munk & Wunsch 1998; Jackson et al. 2008; Běhounková et al. 2010, 2011; Henning & Hurford 2014; Makarov & Efroimsky 2014). Tidal heating in the solid mantle might allow nightside volcanism (Moore et al. 2007, p. 89; Tyler et al. 2015). The WASP-47 system may be an example.

APPENDIX D: ATMOSPHERE MODEL: FROM $\overline{T}$ TO τX

Volatile transport by the vapor-equilibrium atmosphere depends on pressure gradients. Equilibrium pressure depends on Ts(θ) and depends on surface composition. Ts(θ) is given by

Equation (22)

where TAS = 50 K is the antistellar temperature corresponding to a geothermal heat flux of 0.35 W m−2, and ${\rm{\Delta }}T={T}_{{ss}}-{T}_{\mathrm{AS}}$. In the twilight zone, we replace Equation (22) with the temperature corresponding to a linear interpolation of the stellar flux (footnote 6). The resulting Ts(θ) is similar to that in Léger et al. (2011).

Using a uniform-surface-composition assumption to get Peq(θ) (via Equation (12)), we next find the wind driven by ∂P/∂θ, assuming P ≈ Peq. The wind is assumed to be everywhere directed away from the substellar point. Equations (23)–(26) follow the approach of Ingersoll (1989; similar to Castan & Menou 2011). This approach has several limitations, which are discussed in Section 4.1. Neglecting friction and assuming that injected gas has a constant temperature T0, energy conservation yields wind speed v:

Equation (23)

Temperature is set using

Equation (24)

Mass conservation constrains sublimation, E (kg m−2 s−1):

Equation (25)

E is proportional to pressure and to (temperature)1/2. We integrate (23)–(26) by shooting from θ = 0, with initial conditions P = P(Ts(θ = 0)), v = 0, and T = T0. Example output is shown in Figure 6.

The switch from evaporation to surface condensation occurs at θ0 = (35 ± 5)°. θ0 is insensitive to Ts, P(θ = 0), or Xs. Therefore, chemical fractionation (lower P) does not change the basic evaporation-condensation structure shown in Figure 6.

Maintaining a sublimation flux E requires a fractional deviation from the equilibrium pressure of

Equation (26)

(the Hertz–Knudsen equation), assuming a single-component system, with the evaporation coefficient γ ∼ 0.03–0.3 near the liquidus, and rising with temperature (Tsuchiyama et al. 1999; Grossman et al. 2000; Alexander 2001; Richter et al. 2002, 2007, 2011; Fedkin et al. 2006; Lauretta et al. 2006); we set γ = 0.2. We find ΔP/P ≲ 10% $(0.1/\gamma )$, validating the approximation $P\approx {P}_{s}({T}_{s}(\theta ))$. As a second self-consistency check, we feed the latent-heat flux back into the Ts equation (Equation (22)). Changes in Ts are minor (<3%), except for the hottest planet Kepler-78b.

Footnotes

  • Peridotite rock comprises most of Earth's upper mantle. According to the NASA Exoplanets Archive (8/2015), 66 planets have substellar temperature Tss > 1673 K and r < 1.6 r (Rogers 2015), and 103 planets have Tss > 1673 K (assuming albedo = 0.1) and r < 2.5 r (Dumusque et al. 2014). Hot planets with masses and radii in the rocky-planet range include CoRoT-7b, Kepler-10b, Kepler-78b, Kepler-97b, Kepler-99b, Kepler-102b, Kepler-131c, Kepler-406b, Kepler-406c, and WASP-47e, with Kepler-36b and Kepler-93b being slightly cooler than 1673 K (Léger et al. 2009; Batalha et al. 2011; Hatzes et al. 2011; Carter et al. 2012; Howard et al. 2013; Moutou et al. 2013; Pepe et al. 2013; Weiss & Marcy 2014; Dai et al. 2015).

  • Equation (1) assumes a point source of light, but the host star fills ∼1 sr. Therefore, Equation (1) is used only for θp < $\tfrac{1}{2}$($\pi -\zeta $), where ζ is the width (°) of the twilight zone. To find θp for θp > $\tfrac{1}{2}$($\pi -\zeta $), we linearly interpolate the stellar flux between limits of $\tfrac{1}{2}$($\pi -\zeta $) and $\tfrac{1}{2}$($\pi +\zeta $). This crude approximation sets Kepler 10b's pool to cover 61% of the planet's surface, which is reasonable.

  • Latent heats of fusion are <20% lv.

  • Melt-pool depth (i.e., a few times greater than the δT given in Equation (9)) is the relevant length scale for diffusion. An order-unity depletion (relative to original bulk-silicate composition) in the abundance of a chemical component in the pool can be replenished by drawdown of the same chemical component from a layer in the solid mantle that is δp-thick.

  • To see that ρh is also a barrier for condensates beyond the evaporation-condensation boundary (θ0), suppose that after some time the supply of fresh gas from the evaporation zone to the condensation zone is shut off. After shut-off, some gas will condense on the solid planet and will fractionate toward an CaO–Al2O3 lag. For rocky planets with θp > θ0 (∼80 wt% of known hot rocky exoplanets), the dense FeOx-rich residual liquids will sink to dp. The lighter materials will float and so will evolve toward CaO–MgO–Al2O3–SiO2 (CMAS) liquids, but because ρh > ρ0 their boundary layers will be homogenized by small-scale convection and then sink.

Please wait… references are loading.
10.3847/0004-637X/828/2/80