Open access
Technical Papers
Aug 4, 2020

Thermal Extraction of Volatiles from Lunar and Asteroid Regolith in Axisymmetric Crank–Nicolson Modeling

This article has been corrected.
VIEW CORRECTION
Publication: Journal of Aerospace Engineering
Volume 33, Issue 6

Abstract

A physics-based computer model has been developed to support the development of volatile extraction from the regolith of the Moon and asteroids. The model is based upon empirical data sets for extraterrestrial soils and simulants, including thermal conductivity of regolith and mixed composition ice, heat capacity of soil and mixed composition ice, hydrated mineral volatile release patterns, and sublimation of ice. A new thermal conductivity relationship is derived that generalizes the cases of regolith with varying temperature, soil porosity, and pore vapor pressure. Ice composition is based upon measurements of icy ejecta from the Lunar CRater Observation and Sensing Satellite (LCROSS) impact, and the results show that the thermal conductivity and heat capacity equations for water ice provide adequate accuracy at the present level of development. The heat diffusion equations are integrated with gas diffusion equations using multiple adaptive timesteps. The entire model is placed into a Crank–Nicolson framework where the finite-difference formalism was extended to two dimensions in axisymmetry. The one-dimensional version of the model successfully predicts heat transfer that matches lunar and asteroid data sets. The axisymmetric model has been used to study heat dissipation around lunar drills and water extraction in asteroid coring devices.

Introduction

There is growing interest in extracting water from the Moon (Casanova et al. 2017), from Mars (Abbud-Madrid et al. 2016), and from asteroids (Nomura et al. 2017). NASA’s Lunar CRater Observation and Sensing Satellite (LCROSS) demonstrated the existence of water ice on the Moon when it impacted into the Cabeus crater, a permanently shadowed region (PSR) near the Moon’s south pole. The resulting ejecta blanket was found to contain water and other volatiles (Colaprete et al. 2010; Gladstone et al. 2010). Carbonaceous asteroids contain hydrated and hydroxylated phyllosilicates (Jewitt et al. 2007). These volatiles are stable at typical asteroid temperatures in near-Earth orbits in vacuum, but the water evolves when heated to moderate temperature (Zacny et al. 2016). Mars has abundant water in the form of glacial deposits, hydrated minerals including polyhydrated sulfate minerals and hydrated phyllosilicates, and water adsorbed globally at a low weight percent onto the surfaces of regolith grains (Abbud-Madrid et al. 2016).
The extraction of water on the Moon or Mars can be accomplished through strip mining with subsequent processing of the mined material or through in situ thermal techniques: injecting heat into the subsurface, providing a means for vapor or liquid to reach the surface, and collecting it in tanks. Methods on asteroids could be similar or could involve bagging the entire asteroid and heating or spallation of the rocky material with concentrated sunlight (Sercel et al. 2016). Water can be used to make rocket propellant to reduce the cost of operating in space (Sanders et al. 2008; Hubbard et al. 2013; Sowers 2016; Kutter and Sowers 2016), can serve as passive radiation shielding for astronauts (Parker 2006), and can provide life support (Kelsey et al. 2013). In addition to supporting national space agency activities, water-derived rocket propellant can be used commercially for boosting telecommunication satellites from low-Earth orbit or from geosynchronous transfer orbit into geostationary orbit, or for supporting space tourism or other nongovernmental activities (Metzger 2016).
It is difficult to test extraterrestrial water extraction technologies on Earth because of the high preparation costs for realistic test environments: large-scale beds of frozen regolith in vacuum or Mars atmosphere chambers (Kleinhenz and Linne 2013). Nevertheless, testing is vital, as shown by Zacny et al. (2016) performing thermal extraction of water from icy lunar simulant. Those tests found that water vapor moves through the regolith down the thermal gradient away from the hot mining device, so depending on the particular geometry of the system, it could either collect large amounts of water or none at all. This is highly dependent on environmental conditions including vacuum, ice characteristics, and thermal state, so testing without the correct environment would have little value. However, the environments can be simulated numerically to perform low-cost digital design evaluation in lieu of some of the testing, reducing the cost and speeding the schedule of hardware development. For this approach to work, the equations must accurately predict the thermodynamic behaviors of icy, extraterrestrial soil or hydrated minerals in hard vacuum or in low-pressure conditions and at extreme temperatures. This is still challenging because most measurements of thermal conductivity for soils have been performed in terrestrial conditions with liquid water content, Earth’s atmospheric pressure, and/or ambient temperatures. Therefore, more work is needed to develop improved models.
The application that led to the present effort is the World Is Not Enough (WINE) spacecraft concept, which is being developed by Honeybee Robotics under a NASA contract (Zacny et al. 2016). WINE will be a small spacecraft, approximately 27U in CubeSat dimensions (3 by 3 by 3 cubes), with legs for walking short distances and a steam propulsion system for hopping multiple kilometers (Metzger et al. 2016). WINE spacecraft could operate on a body such as dwarf planet Ceres obtaining water from hydrated minerals that may exist on its surface, or on a moon such as Europa where ice is abundant. WINE will drive a corer into the regolith to extract water and perform science and prospecting measurements on the regolith. The water will be extracted thermally by heating the material in the corer. Vapor will travel into a collection chamber where it is frozen onto a cold finger. After multiple coring operations have collected enough water, the tank will be heated to high pressure and vented through a nozzle to produce hopping thrust. The development and validation of the coring and water extraction system require at least two-dimensional (2D) (axisymmetric) computer modeling of heat transfer in the regolith. The modeling is needed to determine energy requirements for this process to set requirements for the spacecraft power system and to determine whether solar energy is adequate or whether radioisotopic heater units (RHUs) are needed to generate adequate thermal energy on a particular planet.
The authors were team members for another application of this modeling: NASA’s Resource Prospector mission (which was canceled while in development). It was planned to prospect for water in the Moon’s polar regions by drilling into the regolith, bringing up cuttings for physical and chemical analysis. One objective of the mission was to determine the temperature of the subsurface regolith around the drill sites. Unfortunately, drilling creates a lot of heat, and experiments showed that it takes hours or even days for the soil to cool back to the original temperature. The mission’s timeline cannot afford for the rover to sit so long in one location waiting to take a measurement. Modeling may be able to help solve this problem as well. The cooling rate around the drill bit should depend on the boundary conditions, which in cylindrical coordinates centered on the drill is the ice temperature asymptotically far from the drill. If the natural subsurface temperatures are relatively constant over distances comparable to the radius that was heated by drilling, then the asymptotic temperature will equal the original temperature at the drilling location. Therefore, measuring only the cooling rate at one or several depths down the drill bit should be adequate for modeling to determine the original temperature of the subsurface. The model will need to be populated with information obtained from the drill cuttings, including the density of the soil and ice content as a function of depth, including possibly chemistry of the ice as measured by the rover’s instruments. The measured drill torque may contribute to calculating the original density of the regolith with depth. If the model has accurate constitutive equations, then with these measurements as inputs, the model can be run repeatedly using different boundary conditions until it correctly reproduces the measured cooling rates around the drill bit. This concept needs to be developed through modeling, which requires improving the model’s constitutive equations, followed by comparison to ground testing before the mission.
Mitchell and De Pater (1994) developed a one-dimensional (1D) model of heat transfer for Mercury and the Moon. It included solar insolation at the surface, geophysical heat flux from the subsurface, and constitutive equations for heat capacity and thermal conductivity of the regolith. The model used a finite-difference framework with the Crank–Nicolson algorithm. Vasavada et al. (1999) extended the model. Vasavada et al. (2012) compared the model to radiometer data of the Moon’s surface heating and cooling throughout a lunar day. Hayne et al. (2017) used it to map apparent looseness of the lunar soil globally.
Here, the model methodology is extended in three ways. This extends the progress first reported by Metzger (2018). First, the model is converted into the axisymmetric 2D Crank–Nicolson form. Second, the constitutive equations are extended based upon additional data sets for soil and mixed composition ice over varying temperatures, porosities, and gas pore pressures. Third, the heat transfer model is merged with algorithms for gas diffusion following the methodology of Scott and Ko (1968). Another model of heat and mass transfer for extraction volatiles from regolith was recently developed by Reiss (2018) using a different methodology than the one that is followed here, so comparing the two models in future work will provide a useful test of the methodologies.

2D Axisymmetric Crank–Nicolson

The 1D thermal model described above has been reproduced and extended to the 2D axisymmetric form. The 2D heat flux equation in Cartesian coordinates is
Tt=k2ρC(2Tx2+2Ty2)
(1)
where T = temperature; k = thermal conductivity of the material; ρ = density of the material; C = heat capacity of the material; and t = time. The equation is discretized for use in a finite-difference model. The left-hand side of the discretized equation calculates the change in T from before to after one time step. The right-hand side could, therefore, be evaluated either before or after that time step. The Crank–Nicolson method is simply to average these two approaches (Crank and Nicolson 1947). This results in a linear system of equations that is stable and can be solved quickly. Using Crank–Nicolson discretization in cartesian coordinates with Δy=Δx, Eq. (1) becomes
2(Δx)2ρiCinΔt(Tijn+1Tijn)=ki,jn(Ti1,jn+Ti1,jn+1)(ki,jn+ki+,jn)(Ti,jn+Ti,jn+1)+ki+,jn(Ti+1,jn+Ti+1,jn+1)+ki,jn(Ti,j1n+Ti,j1n+1)(ki,jn+ki,j+n)(Ti,jn+Ti,jn+1)+ki,j+n(Ti,j+1n+Ti,j+1n+1)
(2)
Converting to axisymmetric form requires the extra terms in the radial derivative, so in cylindrical coordinates with Δr=Δz, it becomes
2(Δz)2ρiCinΔt(Tijn+1Tijn)=ki,jn(Ti1,jn+Ti1,jn+1)(ki,jn+ki+,jn)(Ti,jn+Ti,jn+1)+ki+,jn(Ti+1,jn+Ti+1,jn+1)+ki,jn(Ti,j1n+Ti,j1n+1)(ki,jn+ki,j+n)(Ti,jn+Ti,jn+1)+ki,j+n(Ti,j+1n+Ti,j+1n+1)+ki,jn[(Ti,j+1n+Ti,j+1n+1)(Ti,j1n+Ti,j1n+1)]2j
(3)
where the radial and vertical directions are r and z, respectively, and the discretized radial distance is rj=jΔr=jΔz. Collecting terms with α=kΔt/[2(Δz)2ρC]
(1+αi,jn+αi+,jn+αi,jn+αi,j+n)Tijn+1αi+,jnTi+1,jn+1αi,jnTi1,jn+1(αi,jnαi,jn2j)Ti,j1n+1(αi,j+n+αi,jn2j)Ti,j+1n+1=(1αi,jnαi+,jnαi,jnαi,j+n)Tijn+αi+,jnTi+1,jn+αi,jnTi1,jn+(αi,jnαi,jn2j)Ti,j1n+(αi,j+n+αi,jn2j)Ti,j+1n
(4)
Adapting the method of Summers (2012) to the axisymmetric case, two operators are defined as
δz2Tij=αi,jTi1,j+(αi,j+αi+,j)Ti,jαi+,jTi+1,j
(5)
and
δr2Tij=(αi,jαi,j2j)Ti,j1(αi,j+αi,j+)Ti,j+(αi,j++αi,j2j)Ti,j+1
(6)
so the equation becomes
(1+δz2+δr2)Tijn+1=(1δz2δr2)Tijn
(7)
where the indices for α were linearized for solvability by keeping them at n instead of n+1. The fourth-order cross-derivatives are assumed to be very small and change slowly in time relative to Δt, which should be valid in realistic cases since the heat equation is diffusive and dissipative (DuChateau and Zachmann 2002)
δz2δr2Tijn+1δz2δr2Tijn0
(8)
Subtracting the left-hand side of Eq. (8) from Eq. (7) and collecting terms
(1+δz2+δr2)Tijn+1=(1δz2δr2)Tijn(δz2δr2Tijn+1δz2δr2Tijn)(1+δz2+δr2+δz2δr2)Tijn+1=(1δz2δr2+δz2δr2)Tijn(1+δz2)(1+δr2)Tijn+1=(1δz2)(1δr2)Tijn
(9)
Tij* is defined apart from the constants of integration by the relationship
(1+δz2)Tij*=(1δr2)Tijn
(10)
which is substituted into the right-hand side of Eq. (9)
(1+δz2)(1+δr2)Tijn+1=(1δz2)(1+δz2)Tij*
(11)
The derivatives commute
(1+δz2)(1+δr2)Tijn+1=(1+δz2)(1δz2)Tij*
(12)
(1+δr2)Tijn+1 and (1δz2)Tij* must therefore be equal with the correct choice of constants of integration for Tij*
(1+δr2)Tijn+1=(1δz2)Tij*
(13)
Each term in this system of equations
(1+δz2)Tij*=(1δr2)Tijn(1+δr2)Tijn+1=(1δz2)Tij*
(14)
can be represented as a tridiagonal matrix, so the tridiagonal matrix algorithm can be used to solve it efficiently.
Since this is cylindrical coordinates, the centerline j=0 is a special case that can be handled using the method of discretization by Scott and Ko (1968).
The model is parameterized for the thermal properties of the regolith. It also incorporates radiative heat transfer at its surface using albedo, emissivity, and insolation parameters following Mitchell and De Pater (1994) and Vasavada et al. (1999).

Thermal Conductivity of Regolith wthout Ice

Parameterizing the model’s soil properties relies upon published measurements in the literature, and these measurements are cited in the following paragraphs. Those measurements show that thermal conductivity is a function of temperature, porosity (equivalently, bulk density) of the granular material, and interstitial gas pressure. The Appendix summarizes the data sets that are used in this effort.
Bulk density ρ of the regolith varies on the Moon with location and depth beneath the surface, but it is not well known for asteroids. Lunar values are chosen consistent with Apollo core tubes and other Apollo measurements. Asteroid bulk densities are typically determined by fitting the results of thermal modeling to the observed thermal inertias of the asteroids. Measurements by Rosetta during a flyby of 21 Lutetia indicates the thermal inertia increases below the top few centimeters “in a manner very similar to that of Earth’s Moon” (Keihm et al. 2012). This could indicate particle sizing and/or bulk density variations over that depth, but apart from this, very little is known of possible vertical structure in asteroid regolith. The parameters in this model can be adjusted to match future spacecraft measurements to help determine asteroid regolith structure.
An important question is whether thermal conductivity also varies with particle size distribution. Chen (2008) measured thermal conductivity in terrestrial sands with different particle size distributions, varying porosity and moisture content (only the cases with zero liquid moisture are relevant to airless bodies) at ambient pressure and temperature. The D50 median particle size of these samples varied by a factor of about 6, and samples included some with narrower (well sorted, or uniform) and broader (poorly sorted, or well graded) distributions. The results found thermal conductivity to vary with porosity but not with particle size distribution. Presley and Christensen (1997) measured thermal conductivity in soda lime borosilicate glass beads of various sizes, varying pore gas pressure from 0.5 to 100 torr at ambient temperature. Only one porosity case was measured for each grain size, with finer particles generally forming more porous packings. Since Chen’s results showed thermal conductivity is independent of particle size, the differences in thermal conductivity measured by Presley and Christensen might actually be due to the samples’ porosities, not due to their grain sizes. In contrast, Chen used realistic geomaterials, while Presley and Christensen used spherical beads. The size of contact patches between spherical particles may be correlated to grain diameter, so there may be a grain size dependence that exists in Presley and Christensen’s data that does not exist in realistic regolith. A literature review found no measurements that varied temperatures for different particle sizes while keeping constant porosities or that varied temperatures for different porosities while keeping constant particle sizes; so for now, the results by Chen are the only guidance, and they indicate thermal conductivity does not vary with particle size as an independent variable for realistic geomaterials.
Thermal conductivity for actual lunar soil has been found to follow the form
k=A(1+χ(T350K)3)=A+B(T350K)3
(15)
where K = in kelvins (temperature units); and where χ, A, and B = model parameters. For example, Apollo 12 soil sample number 12001,19 was measured by Cremers and Birkebak (1971) and is shown in Fig. 1, where the dashed line is our fit using A=0.887  mW/m/K and χ=1.56 (R2=0.9929).
Fig. 1. Thermal conductivity measurements of Apollo soil sample 12001,19.
Apollo 14 soil sample 14163,133 was measured by Cremers (1972) at two different bulk densities, as shown in Fig. 2: 1,100  kg/m3 (black dots with dashed curve fit, R2=0.9934) and 1,300  kg/m3 (open circles with gray curve fit, R2=0.9926). These densities are only 17% different and consider the difficulty of maintaining local density in an experimental apparatus, but this appears inadequate to identify a trend.
Fig. 2. Thermal conductivity measurements for an Apollo 14 soil sample at two different bulk densities.
A greater variation of densities was measured by Fountain and West (1970) using crushed basalt in 10 torr vacuum at six different bulk densities, as shown in Fig. 3: ρ1=790 (pluses), ρ2=880 (circle), ρ3=980 (down-triangles), ρ4=1,130 (squares), ρ5=1,300 (diamonds), and ρ6=1,600  kg/m3 (up-triangles). Note the 980 and 880  kg/m3 samples do not follow the trend of decreasing thermal conductivity shown by the other samples probably due to experimental uncertainty.
Fig. 3. Thermal conductivity for six bulk densities of crushed basalt.
Many functional forms were analyzed to fit all these data into one overall equation. Two forms were found to provide excellent fit, and they are discussed in the following sections. The first is a power law of the porosities, and the second is an exponential of the porosities.

First Functional Form

The heat flux field may be decomposed into two contributions. The first is the flux that would exist if radiative heat transfer could be switched off. A hypothesis is that parameter A in Eq. (15) should scale as a power law of the solid fraction of the material
A=A0(1ν)a
(16)
where ν = soil porosity so (1ν) is the solid fraction; and A0 and a = model parameters obtained by fitting the data. The second heat flux contribution is the additional flux field if radiation were switched back on. That additive flux includes both the radiative field in pore spaces and the additional flux in the solid material that provides continuity to the pore flux. A hypothesis is that this passage through both the solid and radiative regions produces the product of a power law of the solid fraction and a power law of the pore fraction, so the parameter B from Eq. (14) scales as
B=B0(1ν)aνb
(17)
where B0 and b = model parameters obtained by fitting data; and a = same value as in Eq. (16). Thus, defining χ0=(B0/A0), the parameter χ from Eq. (14) scales as
χ=χ0νb
(18)
The Ai and χi fitting parameters for the six fitted curves (i=1,2,,6) in Fig. 3 were themselves fitted to Eqs. (17) and (18), with the result
A=7.02(1ν)2.08(mWmK),χ=2.16ν1.44
(19)
with R2=0.9948 and R2=0.9901, respectively. Choosing integer values a=2 and b=1 also produces excellent fits to the data with R2=0.9946 and R2=0.9881, respectively, so the integers were chosen for elegance. The best fits with these exponents are shown in Fig. 4 and are
Ai=6.122(1νi)2andχi=1.82νi
(20)
Fig. 4. Meta-fitting of the curve fitting: (a) A parameter; and (b) χ parameter.
The first form of the generalized function for thermal conductivity is therefore
k(ν,T)=A0(1ν)2[1+χ0ν(T350K)3]
(21)
where A0=6.12×103  W/m-K; and χ0=1.82 for the basalt measured by Fountain and West (1970) but should be generally different for other materials. This equation should be valid over a range wider than the range of the data to which it was fitted, but characterizing the useful extrapolation range is beyond the scope of this work. Note also that A0 cannot be interpreted as the thermal conductivity of the solid material by setting ν=0; basalt’s thermal conductivity is about 400 times larger than this value. The limit ν0 cannot be used this way because the net contact area between grains is determined not only by ν but also by the size of asperities on the grains’ surfaces.
To test Eq. (21), it is plotted in Fig. 5 against the data of Fountain and West (1970). More experimental work is needed to explain why some subsets of the data do not fit as well as others (e.g., 790  kg/m3 data, or the middle temperatures of the 1,500  kg/m3 data). The equation is theoretically elegant, but it is possibly too simple, or the experiment data might have errors due to sample handling or other deficiencies.
Fig. 5. Curve fitting using power laws of porosities.

Second Functional Form

The second functional form follows Chen (2008), which analyzed terrestrial soils at Earth-atmospheric pressure and temperature while varying porosity and moisture content, Sr
k(ν,Sr)=a^(1ν)b^ν[(1c^)Sr+c^]εν
(22)
Chen found an excellent fit using a^=7.5, b^=0.61, c^=0.0022, and ε=0.78. Lunar and asteroid regolith in vacuum are incompatible with liquid moisture content, so Sr=0, which simplifies the equation to an exponential decay
k(ν)=7.5e7.28ν
(23)
Note this lacks a separate temperature-dependent term as in Eqs. (15) and (21) because Chen’s data were all at ambient temperature (T300  K). Including this term
k(ν,T)=A(1+χ(T350K)3)=ea+b(1ν)(1+ec+dν(T350K)3)
(24)
This fits the Fountain and West data as shown in Fig. 6 with
A=e2.118+5.116(1ν)  (mW/m/K),χ=e1.301+2.256ν
(25)
with R2=0.9968 and R2=0.9907, respectfully. Simplifying
k(ν,T)=20.036e5.116ν(1+0.2723e+2.256ν(T350K)3)  (mW/m/K)
(26)
Fig. 6. Meta-fitting for curve fitting, where the solid lines are Eq. (24): (a) A parameter; and (b) χ parameter.
Eq. (26) is plotted in Fig. 7 against the data of Fountain and West (1970).
Fig. 7. Curve fitting using exponentials of porosities.
This provides a slightly better fit to the experimental data, but new measurements with a wider range of porosities should be diagnostic. No function that fits the data better than this has been identified, although many other forms and possible relationships were explored including proportional, linear, quadratic, and logarithmic functions of porosity, and products of power laws of porosity with power laws of solid fraction. The data may not fit even better than this because of experimental uncertainty. It is extremely difficult to maintain constant compaction of a granular material while evacuating the pore pressure because pressure gradients can exceed the overlying weight of soil both macroscopically and microscopically (locally). Also, thermal conductivity measurements can change the compaction of soil because thermal cycling causes grains to expand and contract, and prior work with granular materials (Chen et al. 2006) and lunar soil simulants (Gamsky and Metzger 2010; Metzger et al. 2018) shows this is an effective compaction mechanism. Metzger et al. (2018) found it extremely difficult to maintain low compaction states of lunar simulant because even tiny mechanical shocks cause internal avalanches and compaction. Therefore, it may be difficult to obtain experimental results that fit better than in Fig. 7. The Apollo lunar soil data in Fig. 2 were checked, and they are well-fitted by Eq. (26). For now, this second form is selected for the remainder of the study.

Comparison with Other Data Sets

Fig. 8 compares data from Fountain and West (1970) (FW), Presley and Christensen (1997) (PC), and Chen (2008). The pore pressure differences are discussed in the following section on pore pressure dependence. This section discusses discrepancies in the forms of the curves. The top black points are from Chen’s (2008) data for four sand samples of varied particle size in four packing porosities, each, with no moisture content, at ambient pressure (760  torr), and at ambient temperature (300  K). The solid curve is the fit by Chen evaluated for no moisture content, and the curve is dashed where extrapolating beyond the measurements. The middle graphs (from 100 to 0.5 torr) are from PC with glass spheres in eight samples with each having a different mean particle diameter (each sample is a vertically-aligned set of points correlated to one porosity value) measured at 17 pore pressures (the lines connect different samples at the same pore pressure as a guide to the eye) and ambient temperature (300  K). The bottom black solid line is Eq. (26) fit to FW at 108  torr, evaluated here at T=300  K for consistency with Chen and PC, dashed where extrapolating beyond the range of measured porosities. The error bars were calculated for the six porosities where FW measurements were taken.
Fig. 8. Thermal conductivity versus porosity at different pore pressures, comparing the three data sets.
PC does not fit the form of Eq. (26), but Chen and FW both fit that form. This might be explained by the difference in particle shapes. PC samples were smooth spheres, which under compression have large contact patches that are a function of particle diameter, whereas the Chen and FW samples were geologic materials with irregular shapes and asperities that create much smaller contact patches uncorrelated to particle diameter. However, larger contact patches should produce greater k, but in PC they produced smaller k than the trends of Chen and FW.
Alternatively, the fact that PC does not fit the form of Eq. (26) might be explained as experimental disturbances affecting the coarse particles in PC more than the fine particles. This could happen either through random mechanical vibrations in the laboratory environment (Metzger et al. 2018) or by the drag forces of gas permeation since PC measurements were taken at a variety of pressures. The Kozeny–Carman relationship (Kozeny 1927; Carman 1956; Carrier 2003) applied to the PC data shows the permeability would be two orders of magnitude greater for the coarsest (but least porous particles) than for the finest (but more porous) ones, so gas drag forces would be two orders of magnitude weaker for the larger particles. Cohesive energy to stabilize a granular packing scales as the number of contacts per volume, which scales as the inverse of particle diameter cubed and decreases with porosity due to the decreasing grain contact coordination number by a factor of three over the range of porosities in PC (Murphy 1982). Cohesive energy per grain contact scales proportionally to particle diameter for spheres (Walton 2007). Overall, cohesive forces scale as two and a half orders of magnitude stronger for the finest particles than for the coarsest ones. This rough analysis indicates the finest PC particles should be more resistant to gas permeation disturbance by a factor of five compared to the coarsest particles and more resistant to incidental mechanical shock and vibration disturbance by a factor of 240 compared to the coarsest particles. This supports the hypothesis that the coarser particles (lower porosities in Fig. 8) were more disturbed during the experiments, affecting the shapes of the curves. The reduced thermal conductivity for the coarser particles (relative to the trend lines of Chen and FW) suggests these cases were more porous than believed, indicating the gas exiting the material during vacuum pump-down fluffed these cases and reduced their grain-to-grain contacts.
The consistent curve shape for FW and Chen provides confidence that Eq. (26) can be extrapolated modestly beyond the range of porosities measured by FW. The combined range of porosities measured by Chen and FW is 0.355<ν<0.73. Lunar soil bulk densities are primarily in the range 900<ρ<2,200  kg/m3, corresponding to 0.26<ν<0.71, so only modest extrapolation is required at the low end of the range where Chen data provide high confidence in the functional form. However, a thin surface veneer of epiregolith may exist globally on the Moon, a “fairy castle” state with ν0.9 made possible by low gravity and photoionization in the strong ultraviolet light (Mendell and Noble 2010). Also, experimental work suggests surficial regolith may be more porous in PSRs due to the absence of thermal cycling (Gamsky and Metzger 2010; Metzger et al. 2018). The impact dynamics of the LCROSS spacecraft in the Cabeus crater, a PSR, suggests ν0.7 to a depth of two or more meters. If the geologic processes of a PSR compacted it to only ν0.7 with such overburden, it is possible the soil may be even less compacted in the upper layers where there is less overburden. Extrapolation ν>0.73 via k(ν,T) may therefore be needed, and further laboratory measurements should be performed to validate the model over the wider range of porosities.

Thermal Conductivity of Lunar Ice

For asteroids, the volatile molecules are bound in the crystalline structure of the hydrated minerals and not generally in the form of physical ice. For the Moon, the volatiles include adsorbed molecules on the surfaces of the grains as well as solid ice mixed in the regolith. The contribution of ice to thermophysical properties of regolith is determined by its chemistry and its physical state: e.g., amorphous or crystalline, snow mixed in the pore spaces, and solid ice cobbles like hail. The thermophysical properties of amorphous ice can vary by orders of magnitude depending on density and microstructure (Mastrapa et al. 2013). Amorphous ice crystallizes exothermically when there is adequate activation energy. This has been considered a mechanism for comet outbursts as heat diffuses into the interior reaching amorphous material (Sekanina 2009). In the lunar case, impact gardening could provide the activation energy, crystallizing the deposits as it matures. For now, this thermal model is based on the geological picture presented by Hurley et al. (2012) that the ice began as a homogeneous sheet and was fragmented by impact gardening, mixing grains of pure crystalline ice among grains of otherwise dry soil. The LCROSS impact did detect crystalline ice in the ejecta (Anthony Colaprete, personal communication, 2016). If the fragments are smaller than a volume element in the model, then a volumetric mixing model is adequate. These mixing models have been investigated for icy regolith by Siegler et al. (2012).
The composition of lunar ice was calculated by Tony Colaprete (personal communication, 2016) of NASA based on LCROSS impact ejecta measurements by combining measurements from the two instruments (Colaprete et al. 2010; Gladstone et al. 2010). The calculated volatile concentrations are shown in Table 1. Hydrogen gas was detected, but it should not be stable even at the temperatures of the lunar polar craters. The hydrogen gas and hydroxyl may have been products of chemistry driven by heat of the LCROSS spacecraft impact. It is beyond our present scope to back-calculate what chemicals must have been present in the ice prior to the impact. For now, this remains the best estimate of the composition of lunar ice.
Table 1. Volatiles in LCROSS Ejecta
CompoundSymbolConcentration (% by weight)
WaterH2O5.5
Hydrogen sulfideH2S1.73
Sulfur dioxideSiO20.61
AmmoniaNH30.32
Carbon dioxideCO20.29
EthyleneC2H40.27
MethanolCH3OH0.15
MethaneCH40.03
HydroxylOH0.0017
Carbon monoxideCO0.000003
CalciumCa0.0000008
Hydrogen gasH20.0000007
MercuryHg0.0000006
MagnesiumMg0.0000002
The saturation curves of these volatiles (NIST 2017) shown in Fig. 9 illustrate that temperatures adequate to release water from the regolith will also release many other volatiles. For now, only water sublimation has been modeled. Modeling here may treat the sublimation of each species separately based on partial pressures and treat the diffusion of gas through the pore spaces based on overall pressure and molecular collision rates in the mixed gas. This assumption needs to be checked with measurements of actual lunar ice. Kouchi et al. (2016) found that ice mixtures of CO to H2O in ratios 501 and 101 subjected to conditions for the sublimation of the CO but not the H2O left the water ice in a porous amorphous state with a density similar to high-density amorphous ice. They also found that at 140 K this matrix-sublimated high-density amorphous ice transitioned to cubic ice. Doubtless, the porosity resulting from matrix sublimation decreases thermal conductivity of the remaining matrix and transitioning back to crystalline ice increases it, so these effects may occur when subliming mixed composition lunar ice. The nonwater species of lunar ice constitute less than 50% by weight of the combined ices, so the induced porosity should be much less than reported by Kouchi et al. (2016). For now, the effect is ignored. Future work may add an ad hoc parameter to treat it simplistically, but it would be little more than a guess. To inform a better model, experimental measurement is needed for the thermal conductivity of matrix sublimed, mixed composition ice.
Fig. 9. Saturation curves for chemicals in the lunar ice (L-V = liquid-vapor; S-V = solid-vapor; and TP = triple point, blue dots).
For thermal conductivity, the contribution of pure crystalline water ice is calculated using the data points of Ehrlich et al. (2015), reproduced in Fig. 10 with the added curve fit
k=1.582+11.458exp(T95.271)
(27)
in W/m/K, and where temperature T is in kelvins. Available thermal conductivity data for the other volatiles are limited. Sumarokov et al. (2009) measured CO ice in the range 1–20 K. At 20 K, it is about 0.5  W/m/K, more than an order of magnitude less than the extrapolation of water to that temperature per Fig. 10. Koloskova et al. (1974), cited in Sumarokov et al. (2003), measured CO2 and found it about 1  W/m/K at 100 K, about 1/6 the value of water. Manzhelii et al. (1972) report solid ammonia about 1.8  W/m/K at 100 K, about 1/3 the value of water. Lorenz and Shandera (2001) found ammonia-rich (approximately 10%–30%) water ice has a thermal conductivity of about 1/2 to 1/3 that of pure water ice.
Fig. 10. Thermal conductivity of ice.
In the 1D model of sandwiched materials, the net thermal conductivity is
keff=(diki)1
(28)
where di and ki = thickness and thermal conductivity of each layer. To first-order approximation, which is the limit of accuracy considering the other unknowns, a mixture of dry regolith with ice grains would scale as
kbulk=(Vdryregolithkdryregolith+Vicegrainskicegrains)1(77%O(103)+23%O(6.0))1
(29)
in W/m/K, where 23% volume fraction of ice is derived from 8.9% by weight of ice, a rough estimate based on Table 1 approximating the mixed chemistry as if it were all water. This indicates kbulk0.0013, only about 30% higher than dry regolith. If kicegrains=2  W/m/K to reflect the mixed chemistry instead of 6  W/m/K for pure water ice, then kbulk0.0013, not measurably changed. The mixed chemistry of ice cobbles can safely be ignored for modeling thermal conductivity.

Thermal Conductivity with Gas in the Pores

As volatiles are released, the increasing pore pressure increases the thermal conductivity by orders of magnitude, as shown in Fig. 8, where FW is in hard vacuum, PC is in a range of pore pressures, and Chen is at Earth ambient pressure. The order of magnitude of k compares reasonably for all data sets when pore pressure is accounted for, although the shape of PC disagrees with the other two, as discussed previously. There are inadequate data regarding how to reconcile this, but the following observations lead to a hypothesis. First, as shown in Fig. 8, the PC data with higher porosity are better distributed between the end points formed by the Chen and FW curves than they are at the lower porosities. Second, as shown in Fig. 11, the data at high porosity are continuous with the upper pressure end-point, while the low porosity data are discontinuous. Fig. 11 plots the ratio
Ratio=kPC(ν,P)k(ν,300K)kPC(ν,0.5Pa)k(ν,300K)
(30)
where kPC(ν,P) = PC data; and k(ν,300K) = Eq. (26) based on FW data evaluated at T=300  K (the temperature chosen to match the temperature of the PC data). kPC is evaluated at P=0.5  Pa in the denominator, which is apparently below the floor where pore gas does not contribute significantly to thermal conduction in the soil, as discussed in the following section. Appended on the right side of each set of points is one data point
Ratio=kChen(ν)k(ν,300K)kChen(ν,0.5Pa)k(ν,300K)
(31)
where kChen(ν) = Eq. (22) with the fitting parameters found by Chen and zero moisture content. The continuity for the high porosity cases suggests a hypothesis that the high porosity cases are correct, while the low porosity cases (coarse particles with less cohesion) suffered experimental disturbance. Third, it is understandable why the more porous cases of PC would be more accurate than the less porous cases because they are more stabilized by higher cohesion, as discussed.
Fig. 11. Ratio of thermal conductivity differences for two data sets.
Consistent with the assumption that the most porous data of PC are the least disturbed in the laboratory measurements, trendlines were projected on log-log axes in Fig. 12 such that they intersect at the same point where the FW and Chen trendlines are projected to intersect. These projections pass through the most porous case of PC data. Dots on the top and bottom trendlines are calculations using the Chen and FW fitted functions at the porosities of the PC data.
Fig. 12. Data from Fig. 8 compared to hypothesized model (thin gray lines).
The assumption is that these trend lines are what would have been measured in PC had there been no mechanical disturbance of the samples. This assumption is necessary to reconcile the existing data sets and create a constitutive equation. The family of trend lines is viewed in Fig. 13 through two different projections: into the (ν,k) plane and into the (P,k) plane, where P is pore pressure in the soil. In the (ν,k) plane, the top line is for 760 torr pore pressure (Chen), the bottom is for 108  torr (FW), and the intermediate are for 100 torr (upper) to 0.5 torr (lower) (PC). These trendlines were chosen to intercept at the same point off the right side of the figure where FW and Chen also intercept.
Fig. 13. Trendlines: (a) in the (ν,k) plane; and (b) in the (P,k) plane.
In the (P,k) plane, the far right vertical column of points is from Chen, the left vertical column of points is from FW, and the second column of points from the left is not from any dataset but is the point where the PC data project to an intersection with their corresponding floor. Each floor represents conduction and radiation through the grains without a significant contribution from pore gas. Note that the existence of this floor implies that the gas contribution is additive to the other contributions (not multiplicative).
The general fitting function for this family of curves is
k(ν,P)=P^c1Exp{c2c3Ln2P^+(c4νc5)(Ln2P^c6LnP^c7)}
(32)
where P^=Max(P,P0); and P0 = pressure below which is the floor of minimum conductivity. At constant pressure, this reduces to the form of a single exponential
k(ν)=d1ed2ν
(33)
whereas Eq. (26) at constant temperature reduces to the form of a double exponential
k(ν)=d1ed2ν+d3ed4ν
(34)
The second term is the coefficient for the radiation term in T3. Radiation ought to be independent of gas pressure to good approximation in the rarefied conditions considered here, so the additive form of Eq. (34) agrees with expectations. Eq. (32) was developed from data measured at T=300  K. Therefore, the T3 in Eq. (26) can be added to these curve-fits after subtracting the assumed (300  K)3 contribution. With some manipulation, this yields the full model
k(ν,P,T)=k1P^(k2k3ν)e(k4ν)(1k5T3)+k6ek7n+(k8νk9)Ln2P^
(35)
in W/m/K, where curve fitting with T in kelvins and P in pascals provided the following constants:
P^=Max(P,P0)
(36)
P0=13.68508622330367  Pa
k1=3.419683995668
k2=1.3409114952195769
k3=0.680957757428219
k4=2.8543969429430347
k5=0.000000037037037037
k6=0.799089591748905
k7=2.637142687697802
k8=0.024344154876476995
k9=0.04793741867125248
The decimal places are not all significant, but they are the exact values coded into the model. Quantifying significant digits of model parameters is left to future work when better empirical datasets are available and when the knowledge gaps in the physics have been reduced.

Specific Heat of Regolith wthout Ice

For specific heat, this model uses a mass-weighted mixing model of ice and regolith. The contribution of the dry regolith is informed by the measurements previously made for Apollo soil samples and analog materials. Fig. 14 shows a representative comparison. Apollo samples 10084 and 10057 are from Winter and Saari (1969). The empirical fitting functions by Winter and Saari (1969)
C(T)=0.034T1/2+0.008T0.0002T3/2
(37)
and by Hemingway, Robie, and Wilson (1973)
C(T)=23.173+2.127T+0.015009T27.3699×105T3+9.6552×108T4
(38)
both in J/kg/K, are very close to one another and only slightly higher than the experimental data above 250 K. New fifth-order and fourth-order polynomial fits were tried. The fifth order diverges from the probable trend just outside the range of experimental measurements, so it is rejected. The fourth order is marginally better than the one by Hemingway, Robie, and Wilson (HRW) and seems to preserve the trends in extrapolation. The fit by HRW is actually based on a larger set of measurements, including Apollo soil samples 14163, 15301, 60601, and 10084 to represent the average of lunar soil, so HRW is selected. It fits the data with less than a 10% error.
Fig. 14. Specific heat of lunar soil versus temperature.
The heat capacity of dry regolith should vary proportionally to the soil’s solid fraction (1ν), which varies by about ±16% from the median value in the lunar case. For the asteroid case, the bulk density may vary widely due to large changes in particle size and ultra-low gravity. Measurement of asteroid regolith density in situ, including any stratigraphic variation in the subsurface, is required to guide more accurate models.

Specific Heat of Lunar Ice

The specific heat of pure water ice was measured by Giauque and Stout (1936) from about 15 to 270 K and is
CH2Oice=100.5+11.43T+7.101×103T23.987×104T3+2.075×106T43.200×109T5
(39)
in J/kg/K, where T is in kelvins.
The specific heats of the major volatiles in lunar ice are shown in Fig. 15: (in order of prevalence) water by Giauque and Stout (1936), hydrogen sulfide by Giauque and Blue (1936), sulfur dioxide by Giauque and Stephenson (1938), ammonia by Overstreet and Giauque (1937), carbon dioxide by Giauque and Egan (1937), ethylene by Egan and Kemp (1937), methanol by Carlson and Westrum (1971), methane by Colwell et al. (1963), and carbon monoxide measured by Clayton and Giauque (1932). Weighting these according to Table 1, the composite heat capacity is shown in Fig. 16. This neglects the hydrogen and hydroxyl that were also measured in the lunar ice ejecta, which are assumed to have come from the decomposition of unidentified components.
Fig. 15. Specific heats of components of lunar ice.
Fig. 16. Specific heat of water ice and composite lunar ice.
The composite heat capacity for ice is calculated by a mass-weighted sum of the individual components. This assumes a linear mixing model which may not be correct depending on the actual crystalline or amorphous form of the ice, but until measurements are taken on the Moon, this is the best assumption that can be made. Above the sublimation temperature of each component, the weighting is renormalized for the reduced mass. The heat capacity for the composite ice and for pure water are compared in Fig. 16. The integrated heat capacity of pure water is found to be always within 14% of the integrated heat capacity for composite ice. In the present accuracy of the approximation, pure water’s heat capacity can be used as an adequate representation of the composite ice. The combined specific heat of regolith with 8.9% by weight ice (now approximating it is all water) is shown in Fig. 17. The specific heat of water ice is roughly three times higher than the specific heat of dry lunar soil, but since it constitutes only 8.9% by weight of the regolith, it raises total heat capacity of the mixture by only about 29% at 40 K and about 16% at 200 K.
Fig. 17. Thermal conductivity of water ice, lunar soil, and 8.9% by weight water ice in lunar soil.

Phase Change of Ice

The sublimation of water ice on the Moon at temperatures below the triple point is treated by Andreas (2007) by relating the saturation vapor pressure of water ice, esat,i(T), which is a function of temperature T, to the sublimation rate S0
S0=esat,i(T)(MW2πRT)12
(40)
in kg/m2/s, where MW = molecular weight of water; and R = universal gas constant. Kossiacki and Leliwa-Kopystynski (2015) modified this by subtracting the partial pressure of the vapor P from the saturation pressure
S0=[esat,i(T)P](MW2πRT)12
(41)
When partial pressure reaches saturation pressure, then sublimation should cease. Here, the model uses the vapor pressure relationship provided by Murphy and Koop (2005)
esat,i(T)=14050.7T3.53068exp(5723.265T0.00728332T)
(42)
in pascals.
The free surface area of the ice where sublimation takes place depends on the physical state of the ice, whether it exists as large cobbles of ice surrounded by regolith fines, fine particles of ice comparable to the size of regolith particles intermixed with the mineral grains, a rind of ice coating the mineral grains, amorphous material residing in the pore spaces between grains, or another form. How it is modeled depends on which physical state is assumed. One simple way to model the exposed surface area of the ice Δs (in a cell toroidal of radius r about the model’s axis because this is an axisymmetric model) is Δs=2πrΔzσ, where σ = parameter based on the expected physical state of the ice informed by the geological model. The net sublimed mass during the nth timestep in cell location (i,j) is therefore
mS,ijn=2πrijΔzσij[esat,i(Tijn)Pij](MW2πRTijn)12Δt
(43)
The mass of water in the regolith’s pore spaces in the toroidal cell could be modeled as
mw,ij0=π(rij2ri,j+12)ΔzφijνijρI
(44)
where φij = fraction of the pore space filled by ice; and ρI = density of the ice. A relationship is needed between σij and φij to represent the physical state of the ice. For now, the model uses the simplification σij=φij. Heat capacity is a simple scaling between ice and regolith heat capacities by the amount of each mass within a cell. The temperature in a cell may continue to rise even as ice sublimes until it reaches the triple point because sublimation is a slow process at these temperatures and the system remains in nonequilibrium. In practical cases that have been modeled, the temperature never rose as high as the triple point. As sublimation occurs, the heat of fusion is subtracted from the internal energy of the cell, and the temperature is lowered accordingly before time-stepping the model to calculate the conduction of heat again. The gas and the solid components in each cell are assumed to have the same temperature.

Release of Volatiles from Asteroid Regolith

Hydrated minerals in asteroid regolith release their volatiles as a function of temperature. For testing an asteroid mining prototype, it was economically beneficial to use lower temperature materials for early tests, so a lower temperature simulant was developed using primarily epsomite because it releases most of its water of hydration below 300°C. Curves were obtained through thermo-gravimetric analysis (TGA) to determine the mass of released volatiles at each temperature increment. Examples of these curves are shown in Fig. 18 for asteroid simulant UCF-CI-1 measured by Metzger et al. (2019), the Orgueil meteorite by King et al. (2015), and epsomite by Ruiz-Agudo et al. (2007).
Fig. 18. Thermo-gravimetric analysis (TGA): (a) meteoritic and simulated asteroid materials; and (b) epsomite.
To model this, as a region in the model reaches a new high temperature, the volatiles up to that temperature per the TGA curve are released as vapor, and the model remembers that no more volatiles will be released from that location until an even higher temperature is achieved. The appropriate amount of gas per the TGA curve is added to the gas already in the pore space at that location. Energy spent liberating the volatiles in each step is removed from the regolith appropriately in each time step. Thermal and gas diffusion are then iterated. The equation to fit epsomite [Fig. 18(b)] is
w=[tanh(T508.404)tanh(273.15508.404)]×26.61%  byweight
(45)
where w = weight percent (of a cell’s material) that has sublimed; and T = in kelvins.
The model does not modify the thermal conductivity of the solid fraction of the soil as mass is converted to vapor, although it should reduce that term because (especially with epsomite) a large fraction of the solid mass is lost, reducing the solid conduction contact network. That effect is offset by the increase in conductivity due to rising pore pressure as shown in Fig. 13, but there are no empirical data at present to guide this improvement. Those experimental measurements and modeling are left to future work.

Gas Diffusion

The model incorporates diffusion using the finite-difference equations of Scott and Ko (1968). As vapor is evolved as described, the pressure differences drive it into neighboring cells. To couple the fast gas diffusion equations and the slow thermal diffusion equations while maintaining the stability of the model, it was necessary to implement different time steps for each set of equations. Adaptive timesteps were implemented for the fast diffusion process, dividing each heat flow timestep into the minimum number of smaller timesteps necessary for a stable solution of the gas diffusion equations. As pressure gradients increase, the gas diffusion timesteps become smaller. The resulting model is fast, allowing simulation of an hour-long physical test in just 5 or 10 min on a standard laptop computer.

1D Thermal Model Validation

Only limited testing of the model has been performed. The following four cases demonstrate aspects of the thermal algorithms and the overall code structure with increasing complexity. The first case is 1D simulations of the lunar regolith heating and cooling in sunlight at various latitudes in comparison with Lunar Reconnaissance Orbiter (LRO) Diviner data per Fig. 9(a) of Vasavada et al. (2012). The lunar surface albedo as a function of angle and other parameters choices by Vasavada et al. (2012) were intertwined with choices of thermal conductivity to make their model match lunar data sets. Vasavada et al. (2012) used k=0.6  mW/m/K for the most porous soil at the lunar surface (ν0=0.58, bulk density ρ0=1,300  kg/m3), asymptotically approaching k=7  mW/m/K for the least porous soil at depth (ν=0.42, bulk density ρ=1,800  kg/m3), with the porosity exponentially decaying as a function of depth, z
ν=ν0(ν0ν)ez/H
(46)
where H = single remaining model parameter. Values of H can be iterated until the model makes predictions that match observations of lunar surface temperature rising and falling as the Moon rotates in the sunlight. The value of H is thus a proxy to characterize how rapidly the soil compactifies with depth at each location on the Moon in some averaged sense.
Following the choices of Vasavada et al. (2012), the thermal conductivity form in Eq. (15) becomes
A=e6.898+15.232(1ν)  (mW/m/K),χ=e0.9933
(47)
Ice content is set to zero. The specific heat is represented by Eq. (38). Simulations were performed for the Moon rotating in the sunlight over 37 lunations (months) to achieve a steady state. The final lunation is shown in Fig. 19 for three cases: H=0.5  cm (short dots, top curve, most compacted soil so highest thermal inertia), 3.5 cm (best fit, solid curve), and 30 cm (long dashes, bottom curve, loosest soil so least thermal inertia). The model was successful in predicting lunar temperatures, indicating the model is structured correctly. Future work will use the improved parameterization of Eq. (25), which will make it necessary to determine how albedo and the other model parameters must be changed from the values of Vasavada et al. (2012) to match lunar measurements. This should produce improved characterization of H.
Fig. 19. 1D modeling of the lunar case.
The second case is for equatorial conditions on asteroid 101955 rotating in sunlight. Fine-tuning of the model has not been performed for the asteroid because adequate data sets from asteroids do not exist, but better data are expected soon from spacecraft missions currently in progress. Parameterization is therefore speculative. Many cases were modeled, and they produced similar results with differences that can be tested when the mission data become available. This particular case shown in Fig. 20 used a three-layer regolith model assuming the surface and deepest layers of the asteroid have identical properties, while a layer with different properties exists between 0.5 and 6.0 cm depth. Theory says such an intermediate layer might form on asteroids by thermal cracking as the asteroid rotates in the sunlight, while the uppermost layer loses the fines in the low gravity. The surface and deepest layers were assumed to be very porous with bulk density ρ=1,300  kg/m3, while the intermediate layer was assumed to have ρ=2,340  kg/m3. The specific heat was assumed the same as lunar soil in all layers per Eq. (38). The thermal conductivities of all three layers were assumed to follow Eq. (15). For the surface and deepest layers, parameter A was estimated by taking the value of Vasavada et al. (2012) for the most porous lunar soil, k=0.6  mW/m/K, and then multiplying by the particle size factor suggested by Presley and Christensen (1997), (Dasteroid/Dlunar)0.5, where Dasteroid = average particle size of the asteroid regolith; and Dlunar = average particle size of lunar soil. This could be interpreted as the expected larger contact patches because asteroid regolith is dominated by large gravel particles. Dasteroid=1.5  cm and Dlunar=60  μm result in A=9.49  mW/m/K. χ=2.7 is kept, matching Vasavada et al. (2012). The thermal conductivity of the intermediate layer was assumed to have A=3.736  mW/m/K, corresponding to an intermediate value between the most and least compacted lunar soil, and χ=0.434 for reduced radiative heat transfer due to reduced porosity. The simulation replicated the solar insolation conditions for Bennu and its approximately 4.3 h rotation rate for 500 rotations to achieve steady state. The resulting range of temperatures shown in Fig. 20 correctly matches the range observed on Bennu, as it rotates in the sun per Lauretta et al. (2015).
Fig. 20. 1D modeling of asteroid 101955 Bennu.

2D Axisymmetric Model Validation

The third case adds complexity by using the 2D axisymmetric version of the Crank–Nicolson formulation while retaining the lunar soil property equations of the 1D model [following Vasavada et al. (2012)]. It is a simulation for a drilling test in which a warm drill bit is embedded in soil that carries away its heat. The soil is in a tall, narrow, cylindrical container 14 cm in radius and 120 cm tall. The experiment is inside a warm vacuum chamber. This is the geometry of simulated tests done with a Honeybee Robotics drill at a NASA Glenn Research Center vacuum chamber where a liquid nitrogen bath kept the soil container at constant temperature (77 K) and removed heat from the soil conductively.
In these simulations, four different boundary temperatures (133, 153, 173, and 193 K) instead of the liquid nitrogen bath temperature were successively used to test how the Resource Prospector Mission drill bit could measure cooling rate while embedded in soil. The initial soil temperature before drilling was set to the boundary temperature. The soil model had no ice and was set to ρ0=1,300  k/m3 (ν0=0.58), ρ=1,950  k/m3 (ν=0.37), H=5  cm, porosity following Eq. (46), heat capacity per Eq. (38), thermal conductivity per Eq. (15) parameterized by
A=e5.424+11.717(1ν)  (mW/m/K),χ=e0.9933
(48)
Albedo and emission at the surface follow Vasavada et al. (2012). The vacuum chamber walls are set to 193 K for radiative heat transfer. The drill bit was initially held at 213 K for 5,000 time steps (2 s each) while the soil came to equilibrium, shown in Fig. 21. Then, it was allowed to cool to determine the cooling rate of the bit. The bit and drilling mechanism attached to its upper end were modeled for realistic thermal inertia. The model showed that the cooling process is so slow in lunar soil that it takes hours for soil around a warm drill bit to return to ambient temperature, matching the experimental observations. It is impractical for a lunar rover to pause its mission until the soil cools to take a subsurface temperature measurement. Fig. 22 shows that the cooling rate of the bit depends on the boundary temperature condition, which represents the original subsurface soil temperature before drilling. Thus, the lunar drill can quickly measure the cooling rate and continue its mission, relying on modeling to calculate the original subsurface temperature. The actual lunar case would be more complex than what was simulated here because boundary temperatures (temperature asymptotically far from the drill) should vary with depth. Additional uncertainty exists from the compaction of the soil and ice content with depth. These additional unknowns will be informed in part by drill torque as it is inserted into the subsurface and by analysis of the cuttings for ice content as the drill brings the cuttings up to instruments at the surface. All these measurements would need to be analyzed to inform model parameterization. A future study could analyze the accuracy of this overall method and its sensitivity on the several parameters.
Fig. 21. 2D axisymmetric simulation of a warm drill bit in frozen lunar simulant. Lighter colors represent hotter soil.
Fig. 22. (a) Bit temperature while cooling for four cases with different boundary temperatures 133–193 K; and (b) initial cooling rate of the bit versus boundary temperature.

Thermal Extraction of Water from an Asteroid Simulant

The fourth case integrated the full set of new constitutive equations given by Eqs. (36), (38), and (45) into the 2D axisymmetric Crank–Nicolson model. This was used to simulate the extraction of water from asteroid regolith by the WINE spacecraft coring device (Zacny et al. 2016; Metzger et al. 2016). The corer is a hollow tube with flutes on the exterior that enable it to drill into the subsurface, filling the hollow center of the tube with regolith. The corer walls are heated on the inside, while its layered insulating structure minimizes heat transfer to its exterior. Regolith increases in thermal conductivity as it warms, so the process becomes increasingly efficient. It releases volatiles according to the TGA curve, further increasing thermal conductivity.
The simulations were performed both for terrestrial test conditions with a 1 bar background pressure in the regolith and for space applications with the pores initially in vacuum. Fig. 23 is a case inside of a vacuum chamber with zero ambient pressure and wall temperatures at 293 K. The soil begins at 272 K temperature, and the soil container’s boundaries are kept at 272 K throughout the simulation. In each timestep, 200 W thermal energy is delivered to the inside walls of the corer uniformly along its heated surface and then diffused from the tube through the soil.
Videos were created from the simulation data showing the resulting temperature and pressure fields in the regolith. Fig. 23 shows a series of snapshots of the temperature field in the cross section through the corer. The simulation demonstrated that corer design successfully keeps most of the thermal energy inside the interior, although some energy leaks to the exterior. The videos show that the pressure builds up almost immediately and then decays and the pressure field becomes more uniform. This decay is because the simulant’s water of hydration becomes depleted. Fig. 24 shows a series of snapshots of the vapor pressure field. The semicircular pressure gradient at the top inside the corer is where the vapor diffuses to the collection tube located on the centerline. In the corresponding experiments, the tube leads to the cold trap where volatiles are frozen, keeping the tube at near vacuum conditions, but in the simulation, the vapor that reaches the tube’s entrance is simply accounted for and then removed from the simulation to maintain the tube entrance at vacuum conditions. In the initial simulations, a significant fraction of the vapor can be seen exiting the bottom of the corer rather than diffusing into the collection tube, reducing the system’s mining efficiency. In this particular case, the soil’s initial temperature started near the triple point, so as it was warmed, the vapor exiting the bottom of the corer did not freeze elsewhere in the soil but diffused to the soil’s upper surface where it escaped into the surrounding vacuum.
Fig. 23. Temperature field in the Honeybee Corer at (a) t=18  s; (b) t=90  s; and (c) t=1,080  s.
Fig. 24. Pressure field in the Honeybee Corer driven fully into the soil at (a) t=18  s; (b) t=90  s; and (c) t=1,080  s.
To study how to capture a larger fraction of the vapor, additional simulations were performed where a gap was left between the top of the soil and the inside top of the coring tube. This can be achieved experimentally by not driving to the corer all the way into the soil. This is the case shown in Fig. 25. The gap is so small it is not visible, but it is simulated by appropriate choice of model parameters so the gas can diffuse out from the soil into vacuum all along its top surface inside the corer. This produced a flat pressure gradient across the entire top of the soil inside the corer instead of the semicircular pressure gradient in Fig. 24. Comparing with Fig. 24, the vapor pressure outside the coring tube was reduced because vapor was transported upward through the corer more efficiently. This increased water capture by 520%. This illustrates how modeling can be used to drive the design of mining devices for improved performance in an extraterrestrial environment.
Fig. 25. Pressure field in the Honeybee Corer leaving a gap at the top of the soil at (a) t=18  s; (b) t=90  s; and (c) t=1,080  s.

Conclusions

Thermal volatile extraction modeling has been successfully developed for the 1D and axisymmetric 2D cases for asteroid and lunar regolith. The modeling includes parameterization for regolith thermal conductivity and heat capacity based on measurements of lunar soil samples, simulants, and terrestrial soil and ices. This is apparently the first time a soil constitutive model has successfully reconciled datasets for temperature, porosity, and gas pore pressure variables into a single equation. The model has been only partially tested. It produced excellent agreement with LRO Diviner data of the Moon and estimates of asteroid Bennu heating and cooling as they rotate in the sun. The 2D axisymmetric features have been demonstrated by simulating the Resource Prospector drill cooling after insertion into lunar soil. The model can also simulate the effects of ices upon the thermal conductivity and heat capacity of the lunar regolith. (For the asteroid case, ice is not expected as the volatiles are in the form of hydrated minerals.) The model is based on the assumption that lunar ice is crystalline rather than amorphous, which is supported by some data although the presence of amorphous phases cannot be ruled out. More work is needed to adapt the model to include the effects of amorphous ice. The model also successfully integrated equations for volatile release and gas diffusion along with the thermal diffusion equations, employing multiple adaptive time steps to handle the different characteristic times of each part of the physics. The fully integrated model has been demonstrated for the case of a corer heating and extracting volatiles from asteroid regolith.

Notation

The following symbols are used in this paper:
A, B
model fitting coefficients defined in the text;
a, b
model fitting exponents defined in the text;
a^,b^,c^
model fitting parameters defined in the text;
C
soil heat capacity;
c1,,c7,
model fitting exponents defined in the text;
d
soil layer thickness;
d2,,d4
model fitting exponents defined in the text;
Exp
exponential function;
esat,i
saturation vapor pressure of water ice;
i
index of model cell location in vertical direction;
j
index of model cell location in radial direction;
k
thermal diffusivity constant;
k1,,k9
model fitting exponents defined in the text;
kChen
thermal diffusivity values measured by Chen (2008);
kPC
thermal diffusivity values measured by Presley and Christensen (1997);
Ln
natural logarithm function;
MW
molecular weight;
mS
mass of ice sublimed;
n
index of model timesteps;
O()
order of magnitude;
P
pore gas pressure in the soil;
R
universal gas constant;
r
location in radial direction;
Sr
moisture saturation of soil;
S0
sublimation rate of ice;
T
temperature;
t
time;
V
volume;
w
weight percent sublimed;
z
depth into the soil column;
α
model thermal parameter defined in the text;
Δ
model step difference in variable r, z, or t;
δz2,δr2
calculus operators defined in the text;
ε
model fitting exponent defined in the text;
ν
soil porosity;
π
pi (the number);
ρ
bulk density of soil;
ρI
density of water ice;
σ
parameter to define the surface area of ice in a model cell;
φ
fraction of soil’s mass that is ice; and
χ
radiative transfer coefficient.

Appendix. Particle Samples Analyzed

SourceAbbreviationMean size (μm)Range (m) or gradingPorosityBulk density (kg/m3)Material
Chen (2008)Chen320aUniform0.3961,600bQuartz sand
Chen (2008)Chen320aUniform0.4231,530bQuartz sand
Chen (2008)Chen320aUniform0.4571,440bQuartz sand
Chen (2008)Chen320aUniform0.491,350bQuartz sand
Chen (2008)Chen570aUniform0.4341,500bQuartz sand
Chen (2008)Chen570aUniform0.4721,400bQuartz sand
Chen (2008)Chen570aUniform0.5091,300bQuartz sand
Chen (2008)Chen570aUniform0.5471,200bQuartz sand
Chen (2008)Chen120aUniform0.4341,500bQuartz sand
Chen (2008)Chen120aUniform0.4721,400bQuartz sand
Chen (2008)Chen120aUniform0.5091,300bQuartz sand
Chen (2008)Chen120aUniform0.5471,200bQuartz sand
Chen (2008)Chen310aWell graded0.3541,710bquartz sand
Chen (2008)Chen310aWell graded0.3961,600bQuartz sand
Chen (2008)Chen310aWell graded0.4341,500bQuartz sand
Chen (2008)Chen310aWell graded0.4721,400bQuartz sand
Presley and Christensen (1997)PC805c710–9000.231d2,000Glass spheres
Presley and Christensen (1997)PC510c500–5200.308d1,800Glass spheres
Presley and Christensen (1997)PC262.5c250–2750.231d2,000Glass spheres
Presley and Christensen (1997)PC170c160–1800.346d1,700Glass spheres
Presley and Christensen (1997)PC154.5c149–1600.346d1,700Glass spheres
Presley and Christensen (1997)PC127.5c125–1300.423d1,500Glass spheres
Presley and Christensen (1997)PC95c90–1000.346d1,700Glass spheres
Presley and Christensen (1997)PC72.5c70–750.423d1,500Glass spheres
Presley and Christensen (1997)PC27.5c25–300.462d1,400Glass spheres
Presley and Christensen (1997)PC17.8c15.6–200.538d1,200Glass spheres
Presley and Christensen (1997)PC13.3c11–15.60.654d900Glass spheres
Fountain and West (1970)FW49.5c37–620.737e790Crushed basalt
Fountain and West (1970)FW49.5c37–620.707e880Crushed basalt
Fountain and West (1970)FW49.5c37–620.673e980Crushed basalt
Fountain and West (1970)FW49.5c37–620.623e1,130Crushed basalt
Fountain and West (1970)FW49.5c37–620.567e1,300Crushed basalt
Fountain and West (1970)FW49.5c37–620.500e1,500Crushed basalt
Cremers and Birkebak (1971)12001,1966f<1,000g0.580h1,300Apollo 12 lunar soil
Cremers (1972)14163,13368i<1,000j0.645h1,100Apollo 14 lunar soil
Cremers (1972)14163,13368i<1,000j0.580h1,300Apollo 14 lunar soil
a
These mean sizes are D50 values scaled from Fig. 1 in Chen (2008).
b
Calculated using the reported porosity and 2,650  kg/m3 for the mineral density of quartz.
c
Calculated as the mean of the end points of the reported range.
d
Calculated using the reported bulk density and 2.60 as the approximate specific gravity of the glass.
e
Calculated using the reported bulk density and 3.00 as the specific gravity of basalt.
f
Average of measurements reported in Meyer (2011a).
g
From Meyer (2011a).
h
Calculated using the reported bulk density and 3.10 as the mean specific gravity of lunar soil.
i
Average of measurements reported in Meyer (2011b).
j
From Meyer (2011b).

Data Availability Statement

Some or all data, models, or code generated or used during the study are available from the corresponding author by request: Mathematica notebook containing data from Figs. 18, 1014, 1620, and 2325; Excel spreadsheet containing data from Figs. 9 and 15.

Acknowledgments

This work was directly supported by NASA SBIR contract no. NNX15CK13P, “The World is Not Enough (WINE): Harvesting Local Resources for Eternal Exploration of Space.” This work was directly supported by NASA’s Solar System Exploration Research Virtual Institute cooperative agreement award NNA14AB05A.

References

Abbud-Madrid, A., et al. 2016. “Report of the mars water in-situ resource utilization (ISRU) planning (M-WIP) study.” Accessed May 29, 2020. https://www.nasa.gov/sites/default/files/atoms/files/mars_water_isru_planning.pdf.
Andreas, E. L. 2007. “New estimates for the sublimation rate for ice on the Moon.” Icarus 186 (1): 24–30. https://doi.org/10.1016/j.icarus.2006.08.024.
Carlson, H. G., and E. F. Westrum, Jr. 1971. “Methanol: Heat capacity, enthalpies of transition and melting, and thermodynamic properties from 5–300 K.” J. Chem. Phys. 54 (4): 1464–1471. https://doi.org/10.1063/1.1675039.
Carman, P. C. 1956. The flow of gases through porous media. New York: Academic Press.
Carrier, W. D., III. 2003. “Goodbye, Hazen; Hello, Kozeny-Carman.” J. Geotech. Geoenviron. Eng. 129 (11): 1054–1056. https://doi.org/10.1061/(ASCE)1090-0241(2003)129:11(1054).
Casanova, S., et al. 2017. “Enabling deep space exploration with an in-space propellant depot supplied from lunar ice.” In Proc., 2017 AIAA SPACE and Astronautics Forum and Exposition, 5376. Reston, VA: American Institute of Aeronautics and Astronautics. https://doi.org/10.1061/9780784479018.ch05.
Chen, K., J. Cole, C. Conger, J. Draskovic, M. Lohr, K. Klein, T. Scheidemantel, and P. Schiffer. 2006. “Granular materials: Packing grains by thermal cycling.” Nature 442 (7100): 257. https://doi.org/10.1038/442257a.
Chen, S. X. 2008. “Thermal conductivity of sands.” Heat Mass Transfer 44 (10): 1241–1246. https://doi.org/10.1007/s00231-007-0357-1.
Clayton, J. O., and W. F. Giauque. 1932. “The heat capacity and entropy of carbon monoxide. Heat of vaporization. Vapor pressures of solid and liquid. Free energy to 5000 K. From spectroscopic data.” J. Am. Chem. Soc. 54 (7): 2610–2626. https://doi.org/10.1021/ja01346a004.
Colaprete, A., et al. 2010. “Detection of water in the LCROSS ejecta plume.” Science 330 (6003): 463–468. https://doi.org/10.1126/science.1186986.
Colwell, J. H., E. K. Gill, and J. A. Morrison. 1963. “Thermodynamic properties of CH4 and CD4. Interpretation of the properties of the solids.” J. Chem. Phys. 39 (3): 635–653. https://doi.org/10.1063/1.1734303.
Crank, J., and P. Nicolson. 1947. “A practical method for numerical evaluation of solutions of partial differential equations of the heat-conduction type.” Math. Proc. Cambridge Philos. Soc. 43 (1): 50–67. https://doi.org/10.1017/S0305004100023197.
Cremers, C. J. 1972. “Thermal conductivity of Apollo 14 fines.” In Vol. 3 of Proc., 3rd Lunar Science Conf., 2611–2617. Cambridge, MA: MIT Press.
Cremers, C. J., and R. C. Birkebak. 1971. “Thermal conductivity of fines from Apollo 12.” In Vol. 3 of Proc., 2nd Lunar Science Conf., 2311–2315. New York: Pergamon.
DuChateau, P., and D. W. Zachmann. 2002. Applied partial differential equations. Mineola, NY: Dover Publications.
Egan, C. J., and J. D. Kemp. 1937. “Ethylene. The heat capacity from 15° K. to the boiling point. The heats of fusion and vaporization. The vapor pressure of the liquid. The entropy from thermal measurements compared with the entropy from spectroscopic data.” J. Am. Chem. Soc. 59 (7): 1264–1268. https://doi.org/10.1021/ja01286a031.
Ehrlich, L. E., J. S. G. Feig, S. N. Schiffres, J. A. Malen, and Y. Rabin. 2015. “Large thermal conductivity differences between the crystalline and vitrified states of DMSO with applications to cryopreservation.” PLoS One 10 (5): e0125862. https://doi.org/10.1371/journal.pone.0125862.
Fountain, J. A., and E. A. West. 1970. “Thermal conductivity of particulate basalts as a function of density in simulated lunar and Martian environments.” J. Geophys. Res. 75 (20): 4063–4069. https://doi.org/10.1029/JB075i020p04063.
Gamsky, J., and P. T. Metzger. 2010. “The physical state of lunar soil in the permanently shadowed craters of the moon.” In Proc., Earth and Space 2010 Conf. Reston, VA: ASCE. https://doi.org/10.1061/41096(366)27.
Giauque, W. F., and R. W. Blue. 1936. “Hydrogen sulfide. The heat capacity and vapor pressure of solid and liquid. The heat of vaporization. A comparison of thermodynamic and spectroscopic values of the entropy.” J. Am. Chem. Soc. 58 (5): 831–837. https://doi.org/10.1021/ja01296a045.
Giauque, W. F., and C. J. Egan. 1937. “Carbon dioxide. The heat capacity and vapor pressure of the solid. The heat of sublimation. Thermodynamic and spectroscopic values of the entropy.” J. Chem. Phys. 5 (1): 45–54. https://doi.org/10.1063/1.1749929.
Giauque, W. F., and C. C. Stephenson. 1938. “Sulfur dioxide. The heat capacity of solid and liquid. Vapor pressure. Heat of vaporization. The entropy values from thermal and molecular data.” J. Am. Chem. Soc. 60 (6): 1389–1394. https://doi.org/10.1021/ja01273a034.
Giauque, W. F., and J. W. Stout. 1936. “The entropy of water and the third law of thermodynamics. The heat capacity of ice from 15 to 273° K.” J. Am. Chem. Soc. 58 (7): 1144–1150. https://doi.org/10.1021/ja01298a023.
Gladstone, G. R., et al. 2010. “LRO-LAMP observations of the LCROSS impact plume.” Science 330 (6003): 472–476. https://doi.org/10.1126/science.1186474.
Hayne, P. O., et al. 2017. “Global regolith thermophysical properties of the Moon from the diviner lunar radiometer experiment.” J. Geophys. Res. Planets 122 (12): 2371–2400. https://doi.org/10.1002/2017JE005387.
Hemingway, B. S., R. A. Robie, and W. H. Wilson. 1973. “Specific heats of lunar soils, basalt, and breccias from the Apollo 14, 15, and 16 landing sites, between 90 and 350 K.” Geochim. Cosmochim. Acta 3 (4): 2481–2487.
Hubbard, M. S., P. K. Davidian, D. Gump, J. Keravala, C. Lewicki, T. Gavin, D. Morrison, and W. Hanson. 2013. “Deep space resources: Can we utilize them?” New Space 1 (2): 52–59. https://doi.org/10.1089/space.2013.1505.
Hurley, D. M., D. J. Lawrence, D. B. J. Bussey, R. R. Vondrak, R. C. Elphic, and G. R. Gladstone. 2012. “Two-dimensional distribution of volatiles in the lunar regolith from space weathering simulations.” Geophys. Res. Lett. 39 (9): L09203. https://doi.org/10.1029/2012GL051105.
Jewitt, D., L. Chizmadia, R. Grimm, and D. Prialnik. 2007. “Water in the small bodies of the solar system.” In Protostars and planets, 863–878. Tucson, AZ: Univ. of Arizona Press.
Keihm, S., et al. 2012. “Interpretation of combined infrared, submillimeter, and millimeter thermal flux data obtained during the Rosetta fly-by of Asteroid (21) Lutetia.” Icarus 221 (1): 395–404. https://doi.org/10.1016/j.icarus.2012.08.002.
Kelsey, L., S. A. Padilla, P. Pasadilla, and R. Tate. 2013. “Contaminant robust water extraction from lunar and martian soil for in situ resource utilization-system testing.” In Proc., 43rd Int. Conf. on Environmental Systems, 3438. Reston, VA: American Institute of Aeronautics and Astronautics. https://doi.org/10.2514/6.2013-3438.
King, A. J., J. R. Solomon, P. F. Schofield, and S. S. Russell. 2015. “Characterising the CI and CI-like carbonaceous chondrites using thermogravimetric analysis and infrared spectroscopy.” Earth Planets Space 67 (1): 198. https://doi.org/10.1186/s40623-015-0370-4.
Kleinhenz, J., and D. L. Linne. 2013. “Preparation of a frozen regolith simulant bed for ISRU component testing in a vacuum chamber.” In Proc., 51st AIAA Aerospace Sciences Meeting, 0732. Reston, VA: American Institute of Aeronautics and Astronautics. https://doi.org/10.2514/6.2013-732.
Koloskova, L. A., I. N. Krupskii, V. G. Manzheli, B. Y. Gorodilov, and Y. G. Kravchenko. 1974. “Some peculiarities of heat transport in solid N2O and CO2.” Fiz.Tverdogo Tela 16 (10): 3089–3091.
Kossacki, K. J., and J. Leliwa-Kopystynski. 2015. “Temperature dependence of the sublimation rate of water ice: Influence of impurities.” Icarus 233 (May): 101–105. https://doi.org/10.1016/j.icarus.2014.01.025.
Kouchi, A., T. Hama, Y. Kimura, H. Hidaka, R. Escribano, and N. Watanabe. 2016. “Matrix sublimation method for the formation of high-density amorphous ice.” Chem. Phys. Lett. 658 (Aug): 287–292. https://doi.org/10.1016/j.cplett.2016.06.066.
Kozeny, J. 1927. “Über die Kapillare Leitung des Wassers im Boden.” Proc. Royal Acad. Sci. 136: 271–306.
Kutter, B. F., and G. F. Sowers. 2016. “Cislunar-1000: Transportation supporting a self-sustaining space economy.” In Proc., AIAA SPACE 2016, 5491. Reston, VA: American Institute of Aeronautics and Astronautics. https://doi.org/10.2514/6.2016-5491.
Lauretta, D. S., et al. 2015. “The OSIRIS-REx target asteroid (101955) Bennu: Constraints on its physical, geological, and dynamical nature from astronomical observations.” Meteorit. Planet. Sci. 50 (4): 834–849. https://doi.org/10.1111/maps.12353.
Lorenz, R. D., and S. E. Shandera. 2001. “Physical properties of ammonia-rich ice: Application to Titan.” Geophys. Res. Lett. 28 (2): 215–218. https://doi.org/10.1029/2000GL012199.
Manzhelii, V. G., A. M. Tolkachev, I. N. Krupskii, E. I. Voitovich, V. A. Popov, and L. A. Koloskova. 1972. “Thermal properties of solid ND3.” J. Low Temp. Phys. 7 (1–2): 169–182. https://doi.org/10.1007/BF00629127.
Mastrapa, R. M. E., W. M. Grundy, and M. S. Gudipati. 2013. “Amorphous and crystalline H2O-ice.” In The science of solar system ices, 371–408. New York: Springer.
Mendell, W., and S. Noble. 2010. “The Epiregolith.” In Proc., 41st Lunar and Planetary Science Conf., 1348. Houston: Lunar and Planetary Institute.
Metzger, P. T. 2016. “Space development and space science together, an historic opportunity.” Space Policy 37 (Aug): 77–91. https://doi.org/10.1016/j.spacepol.2016.08.004.
Metzger, P. T. 2018. “Modeling the thermal extraction of water ice from regolith.” In Proc., Earth and Space 2018 Conf. Reston, VA: ASCE. https://doi.org/10.1061/9780784481899.046.
Metzger, P. T., S. Anderson, and A. Colaprete. 2018. “Experiments indicate regolith is looser in the lunar polar regions than at the lunar landing sites.” In Proc., Earth and Space 2018 Conf. Reston, VA: ASCE. https://doi.org/10.1061/9780784481899.009.
Metzger, P. T., D. T. Britt, S. Covey, C. Schultz, K. M. Cannon, K. D. Grossman, J. G. Mantovani, and R. P. Mueller. 2019. “Measuring the fidelity of asteroid regolith and cobble simulants.” Icarus 321 (Mar): 632–646. https://doi.org/10.1016/j.icarus.2018.12.019.
Metzger, P. T., K. Zacny, K. Luczek, and M. Hedlund. 2016. “Analysis of thermal/water propulsion for cubesats that refuel in space.” In Proc., Earth and Space 2016 Conf. Reston, VA: ASCE. https://doi.org/10.1061/9780784479971.044.
Meyer, C. 2011a. “12001–2216 grams, 12003–∼300 grams, reference soil.” Accessed December 23, 2019. https://curator.jsc.nasa.gov/lunar/lsc/12001.pdf.
Meyer, C. 2011b. “14163. Bulk soil sample. 7,776 grams.” Accessed December 23, 2019. https://curator.jsc.nasa.gov/lunar/lsc/14163.pdf.
Mitchell, D. L., and I. De Pater. 1994. “Microwave imaging of Mercury’s thermal emission at wavelengths from 0.3 to 20.5 cm.” Icarus 110 (1): 2–32. https://doi.org/10.1006/icar.1994.1105.
Murphy, D. M., and T. Koop. 2005. “Review of the vapour pressures of ice and supercooled water for atmospheric applications.” Q. J. R. Meteorol. Soc. 131 (608): 1539–1565. https://doi.org/10.1256/qj.04.94.
Murphy, F. W. 1982. “Effects of microstructure and pore fluids on the acoustic properties of granular sedimentary materials.” Ph.D. dissertation, Dept. of Geophysics, Stanford Univ.
NIST. 2017. “Thermophysical properties of fluid systems.” Accessed September 30, 2018. https://webbook.nist.gov/chemistry/fluid/.
Nomura, S., M. Tomooka, and R. Funase. 2017. “Initial design and analysis of a system extracting and collecting water from temporarily captured orbiters.” In Proc., 10th Symp. on Space Resource Utilization, AIAA SciTech Forum, 0651. Reston, VA: American Institute of Aeronautics and Astronautics. https://doi.org/10.2514/6.2017-0651.
Overstreet, R., and W. F. Giauque. 1937. “Ammonia. The heat capacity and vapor pressure of solid and liquid. Heat of vaporization. The entropy values from thermal and spectroscopic data.” J. Am. Chem. Soc. 59 (2): 254–259. https://doi.org/10.1021/ja01281a008.
Parker, E. N. 2006. “Shielding space travelers.” Sci. Am. 294 (3): 40–47. https://doi.org/10.1038/scientificamerican0306-40.
Presley, M. A., and P. R. Christensen. 1997. “Thermal conductivity measurements of particulate materials. 2: Results.” J. Geophys. Res. Planets 102 (E3): 6551–6566. https://doi.org/10.1029/96JE03303.
Reiss, P. 2018. “A combined model of heat and mass transfer for the in situ extraction of volatile water from lunar regolith.” Icarus 306 (May): 1–15. https://doi.org/10.1016/j.icarus.2018.01.020.
Ruiz-Agudo, E., J. D. Martín-Ramos, and C. Rodriguez-Navarro. 2007. “Mechanism and kinetics of dehydration of epsomite crystals formed in the presence of organic additives.” J. Phys. Chem. B 111 (1): 41–52. https://doi.org/10.1021/jp064460b.
Sanders, G. B., W. E. Larson, K. R. Sacksteder, and C. Mclemore. 2008. “NASA in-situ resource utilization (ISRU) project-development & implementation.” In Proc., AIAA Space 2008, 7853. Reston, VA: American Institute of Aeronautics and Astronautics. https://doi.org/10.2514/6.2008-7853.
Scott, R. F., and H.-Y. Ko. 1968. “Transient rocket-engine gas flow in soil.” AIAA J. 6 (2): 258–264. https://doi.org/10.2514/3.4487.
Sekanina, Z. 2009. “Crystallization of gas-laden amorphous water ice, activated by heat transport to its subsurface reservoirs, as trigger of huge explosions of Comet 17P/Holmes.” Int. Comet Q. 31: 99–124.
Sercel, J. C., C. B. Dreyer, A. Abbud-Madrid, D. Britt, R. Jedicke, L. Gertsch, and S. G. Love. 2016. “A coordinated research program to develop the technology to optical mine asteroids.” In Proc., Earth and Space 2016 Conf. Reston, VA: ASCE. https://doi.org/10.1061/9780784479971.048.
Siegler, M., O. Aharonson, E. Carey, M. Choukroun, T. Hudson, N. Schorghofer, and S. Xu. 2012. “Measurements of thermal properties of icy Mars regolith analogs.” J. Geophys. Res. Planets 117 (E3): E03001. https://doi.org/10.1029/2011JE003938.
Sowers, G. F. 2016. “A cislunar transportation system fueled by lunar resources.” Space Policy 37 (Aug): 103–109. https://doi.org/10.1016/j.spacepol.2016.07.004.
Sumarokov, V., A. Jeżowski, and P. Stachowiak. 2009. “The influence of the disordered dipole subsystem on the thermal conductivity of solid CO at low temperatures.” Low Temp. Phys. 35 (4): 343–347. https://doi.org/10.1063/1.3117966.
Sumarokov, V. V., P. Stachowiak, and A. Jeżowski. 2003. “Low-temperature thermal conductivity of solid carbon dioxide.” Low Temp. Phys. 29 (5): 449–450. https://doi.org/10.1063/1.1542510.
Summers, P. 2012. “2D heat equation modeled by Crank-Nicolson method.” Accessed December 23, 2019. http://wiki.tomabel.org/images/c/c2/Paul_Summers_Final_Write_up.pdf.
Vasavada, A. R., J. L. Bandfield, B. T. Greenhagen, P. O. Hayne, M. A. Siegler, J.-P. Williams, and D. A. Paige. 2012. “Lunar equatorial surface temperatures and regolith properties from the diviner lunar radiometer experiment.” J. Geophys. Res. Planets 117 (E12): E00H18. https://doi.org/10.1029/2011JE003987.
Vasavada, A. R., D. A. Paige, and S. E. Wood. 1999. “Near-surface temperatures on Mercury and the Moon and the stability of polar ice deposits.” Icarus 141 (2): 179–193. https://doi.org/10.1006/icar.1999.6175.
Walton, O. R. 2007. Adhesion of lunar dust. Cleveland: NASA Glenn Research Center.
Winter, D. F., and J. M. Saari. 1969. “A particulate thermophysical model of the lunar soil.” Astrophys. J. 156: 1135–1151. https://doi.org/10.1086/150041.
Zacny, K., P. Metzger, K. Luczek, J. Mantovani, R. Mueller, and J. Spring. 2016. “The world is not enough (WINE): Harvesting local resources for eternal exploration of space.” In Proc., AIAA Space 2016, 5279. Reston, VA: American Institute of Aeronautics and Astronautics. https://doi.org/10.2514/6.2016-5279.

Information & Authors

Information

Published In

Go to Journal of Aerospace Engineering
Journal of Aerospace Engineering
Volume 33Issue 6November 2020

History

Received: Jul 4, 2019
Accepted: Mar 6, 2020
Published online: Aug 4, 2020
Published in print: Nov 1, 2020
Discussion open until: Jan 4, 2021

Authors

Affiliations

Planetary Scientist, Florida Space Institute, Univ. of Central Florida, 12354 Research Pkwy., Partnership 1 Bldg., Suite 214, Orlando, FL 32826-0650 (corresponding author). ORCID: https://orcid.org/0000-0002-6871-5358. Email: [email protected]
Kris Zacny, Ph.D., A.M.ASCE [email protected]
Vice President, Honeybee Robotics, 398 W Washington Blvd. Suite 200, Pasadena, CA 91103. Email: [email protected]
Phillip Morrison [email protected]
Senior Systems Engineer, Honeybee Robotics, 398 W Washington Blvd. Suite 200, Pasadena, CA 91103. Email: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share